The Island Chumash: Behavioral Ecology of a Maritime Society 9780520931435

Colonized as early as 13,500 years ago, the Northern Channel Islands of California offer some of the earliest evidence o

163 52 5MB

English Pages 310 [312] Year 2005

Report DMCA / Copyright

DOWNLOAD FILE

Polecaj historie

The Island Chumash: Behavioral Ecology of a Maritime Society
 9780520931435

Table of contents :
Contents
Preface
1. The Island Chumash
2. Human Behavioral Ecology and Maritime Societies
3. Environmental Context
4. Cultural Context
5. Historic Island Communities
6. Terminal Pleistocene to Middle Holocene Records
7. Late Holocene Record
8. Synthesis
References
Index

Citation preview

The Island Chumash

This page intentionally left blank

The Island Chumash Behavioral Ecology of a Maritime Society

Douglas J. Kennett

UNIVERSITY OF CALIFORNIA PRESS

Berkeley / Los Angeles / London

University of California Press Berkeley and Los Angeles, California University of California Press, Ltd. London, England © 2005 by the Regents of the University of California Library of Congress Cataloging-in-Publication Data Kennett, Douglas J. The island Chumash : behavioral ecology of a maritime society / Douglas J. Kennett. p. cm. Includes bibliographical references and index. ISBN 0-520-24302-1 (cloth : alk. paper) 1. Chumash Indians.

2. Indians of North America—California. I. Title.

E99.C815K46 2004 979.4004'9758—dc22 2004006985 Manufactured in China 10 09 08 07 06 05 10 9 8 7 6 5 4 3 2 1 The paper used in this publication meets the minimum requirements of ANSI/NISO Z39.48-1992 (R 1997) (Permanence of Paper).

For my parents, Diana and James, for their unconditional love and support

This page intentionally left blank

Contents

Preface 1. The Island Chumash Study Area Climate Change and Emergent Cultural Complexity

2. Human Behavioral Ecology and Maritime Societies Maritime Foraging Strategies Diet Choice in Maritime Settings Return Rates for Marine Resources Central Place Foraging and Maritime Foragers Intensification and the Ideal Free Distribution Competition and the Formation of Social Hierarchies in Coastal Settings Summary

3. Environmental Context General Physiography Geology Climate Hydrology Terrestrial Resources: Spatial Distribution Marine Resources: Spatial Distribution Seasonal Variability Short-Term Climatic Variability Paleoenvironment

ix 1 4 6 10 14 16 20 29 32 36 39 41 42 42 46 48 49 54 58 60 61

Sea Level Marine Climate History Terrestrial Climate History Summary

4. Cultural Context The Ethnohistoric Record The Prehistoric Record Summary

5. Historic Island Communities Historic Island Villages Santa Cruz Island Santa Rosa Island San Miguel Island Geographic Analysis Viewshed Analysis Rank-Size Analysis Summary

6. Terminal Pleistocene to Middle Holocene Records Terminal Pleistocene Record Early Holocene Record Middle Holocene Record

7. Late Holocene Record Population Growth and Demographic Expansion Territoriality and Warfare Economic Intensificaton Increases in Trade and Exchange Emergent Sociopolitical Complexity

8. Synthesis

62 64 69 71 72 72 80 90 91 93 93 97 104 104 105 108 111 112 113 122 128 154 155 180 187 198 209

Diet Breadth Central Place Foraging Intensification and the Ideal Free Distribution Competition and the Formation of Social Hierarchies Human Behavioral Ecology and Maritime Societies

217 217 224 229 233 236

References

239

Index

291

Preface

This book explores the evolutionary history of the island Chumash using a series of models derived from neo-Darwinian behavioral ecology. The Chumash occupied the four northern Channel Islands— Anacapa, Santa Cruz, Santa Rosa, and San Miguel—that form a chain off the Southern California coast. These islands are best known for their early colonization history (~13,500 to 12,500 calendar years before present) and unusually complex maritime society that was first documented by Europeans in October of 1542 when Juan Rodríguez Cabrillo sailed into the Santa Barbara Channel region for the first time. Each of these islands is unique ecologically, and they are truly magical places to visit. They also contain some of the best-preserved archaeological records in North America, owing to the absence of both burrowing animals, and, thanks to Channel Islands National Park, the urban sprawl that has impacted much of the coastal landscape in Southern California. The allure of these islands, sometimes visible from the mainland across the Santa Barbara Channel, overwhelmed me as a graduate student at the University of California, Santa Barbara, and my first visit to Santa Cruz Island with Michael Glassow in 1993 stimulated a pleasant interruption to a career largely devoted to Mesoamerican archaeology. At the core of this book is my Ph.D. work conducted on the islands between 1994 and 1998. However, the volume before you differs substantially from this original work in two ways. First, it provides a more synthetic summary of island prehistory and ethnohistory based on work conducted on the northern Channel Islands by a large number of ix

x

PREFACE

archaeologists and enthnohistorians. In particular, I highlight the recent and outstanding work of Jeanne Arnold, Roger Colten, Jon Erlandson, Michael Glassow, John Johnson, Pat Lambert, Ann Munns, Jennifer Perry, Scott Pletka, Torben Rick, Jan Timbrook, René Vellanoweth, and Phillip Walker. This includes three important volumes that have been published since the completion of my Ph.D. work: The Origins of a Pacific Coast Chiefdom: The Chumash of the Channel Islands (2001, edited by Jeanne Arnold); Catalysts to Complexity: Late Holocene Societies of the California Coast (2002, edited by Jon Erlandson and Terry Jones), and Cultural Affiliation and Lineal Descent of Chumash Peoples in the Channel Islands and Santa Monica Mountains (1999, edited by S. McLendon and J. R. Johnson). This final volume has been particularly important for establishing a dialogue with the living descendants of the island Chumash that was sorely, and regretfully, lacking when I conducted my research. Second, I employ much more formal models from human behavioral ecology (HBE) to explore the evolutionary history of the island Chumash. HBE is successfully providing a general conceptual framework for the analysis of a range of human behaviors in living and prehistoric societies from many different environmental and social settings. These models also provide a set of hypotheses, or predictions, regarding diet choice, economic intensification, and the development of sociopolitical complexity that are evaluated based on the available archaeological and ethnohistoric records from the islands. In some cases the data are relatively consistent with model predictions, whereas in others the data are inconsistent and point to fruitful avenues for future research. Ultimately, however, this analysis supports the conclusion that the social and political complexity evident at historic contact was ultimately a product of individual behavioral responses, both competitive and cooperative, to demographic expansion, human-induced impacts to marine habitats, and periods of rapid paleoclimatic change. As with any project of this magnitude it is difficult to reconstruct the complex web of intellectual influences and their affect on my thinking regarding the evolutionary history of the island Chumash. Michael Glassow initially introduced me to the archaeology of the northern Channel Islands, and Jon Erlandson, along with Don Morris, encouraged me to pursue my Ph.D. on these beautiful islands. While conducting my research at the University of California, Santa Barbara, I benefited enormously from discussions with Mark Aldenderfer, Michael Glassow, Michael Jochim, John Kantner, Barbara Voorhies, and Phillip

PREFACE

xi

Walker, several of whom served patiently on my Ph.D. committee. On the topic of the prehistory and ecology of the northern Channel Islands, I have also benefited greatly over the years from discussions with Larry Agenbroad, Jeanne Arnold, Roger Colten, Robert DeLong, Michael Glassow, James Kennett, Terry Jones, John Erlandson, John Johnson, Pat Lambert, Don Morris, Jennifer Perry, Torben Rick, and Phillip Walker. Special thanks goes to Jon Erlandson and Torben Rick for many stimulating conversations regarding the prehistory of the northern Channel Islands and for reading and extensively commenting on various drafts of this manuscript. My theoretical approach to island Chumash prehistory is ecological and evolutionary and has been influenced by a large number of scholars. In particular, Mark Aldenderfer, Jon Erlandson, Michael Jochim, John Kantner, Daniel Larson, Sarah B. McClure, Torben Rick, and Bruce Winterhalder are acknowledged for many stimulating conversations about evolutionary/ecological theory and its application to archaeological problems. This multidisciplinary project would not have been possible without funding from a number of sources. The National Science Foundation (Grant #SBR-9521974) provided the initial funding required to establish my research program on the northern Channel Islands. Channel Islands National Park provided transportation to the islands and all of the logistical support necessary to conduct the fieldwork. The National Park Service also provided financial support for both field and laboratory studies. Indeed, this project has its origins in a small National Park Service grant that I received to survey Cañada Verde on the north coast of Santa Rosa Island. The success of this early project ultimately led to the establishment of a cooperative agreement between Channel Islands National Park and UC Santa Barbara that supported several students and improved the quality of this project immensely (Grant #1443CA8120-96-003). A countless number of enthusiastic undergraduate and graduate students helped with the laboratory and fieldwork needed to complete this project. Field studies were conducted with assistance from Lisa Cipolla, Robert Clifford, Christina Conlee, Juliet Christy, Michael Cruz, Jelmer Eerkens, Chris Forester, Holly Gunn, Susan Harris, Corina Kellner, Liz Klarich, Scott Lehman, Alina Meneken Maldonado, Dan Momii, Peter Paige, Robin Palmer, Lubo Popov, Adrienne Reese, Brian Stokes, and Dawn and Kevin Vaughn. Laboratory work was conducted with help from Patrick Collins, Christina Conlee, Holly Gunn, Scott Lehman, Miguel Moran, Tina Oglesby, Robin Palmer, Lubo Popov, Adrienne Reese, Brixen Reich, Lisa Roulette, and Kevin Vaughn. I also thank

xii

PREFACE

John Johnson, Linda Agren, and Tim Hazeltine at the Santa Barbara Museum of Natural History for providing me access to the archaeological collections and sharing their extensive knowledge. Thanks also to Leal Mertes and Ben Waltenberger (Department of Geography, UCSB) for sharing the geographic datasets needed to complete this study. I am indebted to Joan Murdoch and the staff at the Humanities and Social Science Computing Facility for unlimited access to the computer equipment needed to complete the project and to my father, James P. Kennett, for access to his laboratory to complete the stable isotopic work presented in this book. A large number of people also helped with the formulation, preparation, and production of this book. I thank Blake Edgar and the staff at UC Press for patiently waiting on various drafts of this book and Mark Raab, two anonymous reviewers, and the editorial committee at UC Press, all of whom provided valuable comments that improved this book immensely. Graphical assistance was provided by Jacob Bartruff (Maps 2, 3, 6, 7, 8, 9, 10, 11, 12, 13, 14, and 15), Chris Goldsmith (Figure 9), and Rusty van Rossman (Maps 1 and 5 and Figures 7, 17, 19, 24, 25, 28). The following photographs are included with the permission of the photographers: Figures 5C,D, 6A,B,C,E,F,H, 18, 20B, 29A,B, Sam Spaulding; and Figure 16, Jon Erlandson. Thank you all for allowing me to use these materials. Finally, I thank my family for their support throughout the duration of this project and particularly my wife, Sarah B. McClure, who helped in all stages of this project, from fieldwork to proofreading. I would not have completed this work without her love and enthusiastic view of both the modern and the prehistoric world.

The Island Chumash

Long ago before there were white people here, a wizard at San Gabriel caused a great drought and famine that lasted for five years. María’s grandfather Estévan and his wife lived through this period. This sorcerer had a tablet or stone on which he painted many figures of men and women bleeding from the mouth and falling down. He took this out into the hills and exposed it to the sun, praying for sickness. For five years there was no rain, and many people died of hunger. The seeds the women had stored came to an end, and there were no acorns or islay. Even the shells along the shore had only sand in them. When the men went out to get mescal only a few came back—the rest died of hunger—and the women put hot water on their breasts so they would have some milk to keep husbands or brothers alive. Finally some other sorcerers spied on the man who was causing all this and saw him going into the hills. When he returned they said, “Let’s follow his tracks and see what he has been doing.” They followed the tracks and found the tablet or whatever it was with the evil figures on it, and they threw it into the water. The drought ended and it rained. How did those people know how to do those things? They were wiser than the American in those days (Blackburn 1975, 276).

chapter 1

The Island Chumash

Twenty miles off the coast of southern California, a chain of four islands extends from east to west along the southern border of the Santa Barbara Channel—the northern Channel Islands (map 1). People colonized the larger islands in this chain as early as 13,500 years ago and archaeological evidence indicates that they were permanently occupied by at least 7,500 BP.* At the time of European contact (AD 1542), island Chumash people were heavily dependent upon maritime resources, lived in relatively large coastal communities governed by chiefs, and had active exchange and marriage alliances with their relatives on the mainland coast (Johnson 1988, 2000; Kennett and Kennett 2000). Indeed, some archaeologists consider the people that lived on this section of the California coast to be among the most socially and politically complex hunter-gatherers in the world (Arnold 2001; Erlandson and Rick 2002a). The primary goal of this book is to explore the evolutionary history of these people with an emphasis on stability and change in subsistence and settlement strategies, economic intensification, competitive and cooperative behavior, and the ultimate development of the social and political complexity that was evident at historic contact. Human behavioral ecology (HBE) provides the theoretical framework for studying cultural evolutionary processes in this maritime setting. As the Chumash myth in the epigraph suggests, the conditions

*All dates presented in this volume are calibrated year before present (BP).

1

map 1. Map of the northern Channel Islands and adjacent mainland (drafted by R. van Rossman).

THE ISLAND CHUMASH

3

(both social and environmental) in the Santa Barbara Channel region were not constant through time and people were continually confronted with new situations that they had to contend with, and these decisions ultimately had strong evolutionary consequences. HBE is well equipped to explain behavioral variability and adaptive design within changing ecological contexts.* (Bettinger and Richerson 1996; Smith and Winterhalder 1992; Winterhalder and Smith 1992, 2000). It is firmly grounded in neo-Darwinian principles, particularly natural selection theory, and furnishes an expanding set of models that can be tested with ethnographic or archaeological data within different environmental and social settings (Bettinger and Richerson 1996; Broughton and O’Connell 1999; Winterhalder and Smith 1992, 2000). During the past 20 years, HBE has provided a solid theoretical framework for studying human foraging behavior (Begossi 1992; Broughton 1994a, 1999; Madsen and Schmitt 1998; Smith 1991; Ugan and Bright 2001), and the repertoire of topics addressed has expanded to include various non–subsistence-related behaviors such as land tenure and the development of social and political hierarchies (see chapter 2). Despite the success of HBE for exploring and explaining changes in human behavior, it is rarely used by archaeologists working in maritime settings (for exceptions see Broughton 1999; Hildebrandt and Jones 1992). This is partially related to the tendency for archaeologists to emphasize terrestrial resources and to discount the important role that maritime or aquatic habitats and resources played in the evolutionary history of our species (Pálsson 1988). Some archaeologists have even argued that aquatic resources were a “last resort” for foragers living in coastal settings, and the relatively late prehistoric use of these resources worldwide is often cited as evidence for their low value (Binford 1968; Osborn 1977; Pálsson 1988). This is inconsistent with ethnographic and archaeological evidence indicating that maritime environments supported some of the largest and most complex hunter-gatherer groups on earth (Erlandson 2001; Pálsson 1988; Yesner 1980, 1987). Indeed, one study suggests that the fatty acids found in marine foods such as fish played an essential role in the early development of the human brain (Gibbons 2002) and that aquatic and maritime adaptations were important for the demographic and *Ecological context is broadly defined here to include environmental, economic, social, and political contingencies.

4

THE ISLAND CHUMASH

geographic expansion of anatomically modern humans after 150,000 years ago (Erlandson 2001). Nevertheless, archaeologists working in maritime settings are left in a theoretical vacuum without the tools necessary to explore cultural evolutionary processes. These processes are complex, and it is argued here that HBE is well positioned to fill this theoretical void.

Study Area When Juan Rodríguez Cabrillo sailed into the Santa Barbara Channel in AD 1542, much of the Chumash population was concentrated along the mainland coast, but large numbers of people also lived on the northern Channel Islands. Ethnohistorical accounts indicate that they lived in large coastal villages, were heavily dependent upon fishing, produced a variety of trade items, and participated in an extensive interregional exchange network (Johnson 1988). The plank canoe was a vital innovation that was used for open ocean fishing and intervillage commerce (Arnold 1995; Gamble 2002). Food and nonfood items were exchanged between individuals living in coastal and island villages (King 1990). Islanders produced a variety of trade items, including shell beads that were considered a medium of exchange throughout Chumash territory (Arnold and Munns 1994; King 1990). Aspects of this exchange system were controlled by elite individuals who commanded some economic and political authority (Johnson 2000). The evolutionary history of Chumash society is explored in this book based on the rich archaeological records on the northern Channel Islands of California. Studies suggest that some of these islands were colonized as early as 13,500 BP (Erlandson et al. 1996a, 1996b; Johnson et al. 2000; Rick et al. 2001a), providing the first evidence for human occupation of coastal habitats in North America (see Jones et al. 2002). The archaeological record clearly shows permanent occupation of all the larger Channel Islands by 7,500 years ago and suggests that many of the cultural traditions recorded in the area by Spanish explorers were well established by at least 650 years ago. In this book, I explore the emergence and dominance of the human behavior associated with these cultural traditions. In a general sense, islands provide well-bounded areas to study cultural evolutionary processes and consequently have received much attention by archaeologists (Burley and Dickinson 2001; Keegan and

THE ISLAND CHUMASH

5

Diamond 1987; Kirch 2000; Patton 1996). The offshore islands of California are particularly important because they contain the longest and best-preserved archaeological sequences available for study along the west coast of North America. Thanks to Channel Islands National Park and The Nature Conservancy, the islands are permanently protected from the urban sprawl that is impacting much of southern California’s coastal landscape. Archaeological deposits are also relatively undisturbed compared with contemporary sites on the mainland because the pocket gopher (Thomomys spp.), an incessant burrower that mixes archaeological sequences (Erlandson 1984), never colonized the Channel Islands. Well-preserved archaeological deposits and the long, continuous record of human occupation make the Channel Islands one of the best areas in western North America to study foraging strategies, adaptive variability, and evolutionary processes. Although there are many benefits to studying prehistoric behavioral strategies on the northern Channel Islands, there are also some fundamental limitations, particularly when examining the late Holocene development of sociopolitical complexity. This is because the largest and most important villages (economically and politically) were positioned along the mainland coast where a confluence of key environmental variables promoted higher population levels (Erlandson 1994). Because mainland archaeological sequences are relatively poorly preserved, the higher-resolution records of the northern Channel Islands are crucial as a proxy for what was occurring throughout the region. On the other hand, the behavioral strategies used by islanders during the Holocene cannot be decoupled from the opportunities and constraints imposed by economic and political developments that occurred more broadly in the region. In this book, I synthesize environmental, ethnohistorical, and archaeological data from the northern Channel Islands collected during the past century (Arnold 1987, 2001; Colten 1995, 2001; Erlandson 1994; Erlandson and Colten 1991; Erlandson and Glassow 1997; Erlandson and Jones 2002; Gamble 2002; Gamble et al. 2001; Glassow 1977, 1980, 1993a, 1997, 2000; Glassow and Wilcoxon 1988; Heizer and Elsasser 1956; Johnson 1982, 1988, 1993, 2000, 2001; King 1990; Perry 2003; Rick et al. 2001a; Rogers 1929). These studies are augmented with my own sustained work on the islands since 1993 (figure 1A). The bulk of these data were collected during my Ph.D. research, and the data-gathering methods used are presented in greater detail elsewhere (Kennett 1998). The study included surveys on Santa Cruz, Santa Rosa, and San Miguel islands

6

THE ISLAND CHUMASH

A

B

figure 1. (A) Photograph of Michael Glassow (left) and James Kennett (right) discussing the geologic context of chert microblade quarries on eastern Santa Cruz Island. Photo taken during the author’s first visit to the northern Channel Islands in 1993 (photo by D. Kennett looking west across Santa Cruz Island). (B) Don Morris excavating column sample from seacliff exposure at CA-SRI-15, western Santa Rosa Island (photo by D. Kennett).

coupled with small-scale sampling of archaeological sites. Many of the samples were taken from wave-cut exposures in seacliffs to minimize the impact on these important archaeological deposits (figure 1B). These data, along with published information on other islands sites (~1,900 sites), were incorporated into a digital spatial database (Geographic Information System or GIS) to explore changes in settlement location and land use through time. Artifact assemblages and radiocarbon dates were used to determine when sites were occupied. Quantitative analyses of faunal and tool assemblages provided information regarding changes in subsistence, technology, and social organization through time. Stable oxygen and carbon isotopic analyses of marine shells from key sites were used to determine the season of shellfish harvesting and to infer the duration and season of site use.

Climate Change and Emergent Cultural Complexity Archaeological data indicate that socially and politically complex huntergatherer societies were well established on the southern California coast, including the northern Channel Islands, by at least 650 years ago. Archaeologists working in southern California have long suggested that changes in climate during the Holocene played an important role in the development of greater sociopolitical complexity (Arnold 1987, 1991, 1992a, 1992b, 1993, 1997, 2001; Arnold et al. 1997; Colten 1993, 1994,

THE ISLAND CHUMASH

7

1995; Erlandson 1997a, 1997b; Glassow et al. 1988, 1994; Johnson 2000; Jones and Kennett 1999; Jones et al. 1999; Kennett 1998; Kennett and Kennett 2000; Lambert 1994, 1997; Lambert and Walker 1991). The exploration of climatic influences is not surprising given the quality of the available climatic and archaeological records and the relative sensitivity of the region to major climatic perturbations (Behl and Kennett 1996; Cannariato et al. 1999; Hendy and Kennett 1999, 2000; Heusser 1978; Heusser and Sirocko 1997; Kennett and Ingram 1995a, 1995b; Kennett and Kennett 2000; Pisias 1978, 1979). Nevertheless, considerable debate has emerged about the details of the timing and nature of sociopolitical evolution on the southern California coast in relation to environmental change (Arnold 1992a; Arnold et al. 1997; Erlandson 2002a; Kennett 1998; King 1990; Raab and Bradford 1997; Raab and Larson 1997; Raab et al. 1995). Central to this debate are the specific environmental and biotic changes that may have triggered sociopolitical and economic developments. In particular, this debate has focused on the relative roles of changing terrestrial and marine ecosystems on cultural development between 800 and 650 years ago (Arnold 1997; Arnold et al. 1997; Jones and Kennett 1999; Jones et al. 1999; Lambert 1997; Lambert and Walker 1991; Raab and Larson 1997). Arnold (1992a, 1997) and others (Arnold and Tissot 1993; Arnold et al. 1997; Colten 1993, 1994, 1995; Colten and Arnold 1998, 2000) have focused on sociopolitical and economic responses to marine and terrestrial climatic stresses on the northern Channel Islands, particularly a reported interval of elevated sea-surface temperature and low marine productivity between 800 to 650 years ago. Elevated sea-surface temperatures were inferred from a longstanding marine paleoclimatic sequence for the region and faunal assemblages from Santa Cruz Island (Arnold 1992b; Arnold and Tissot 1993; Colten 1994, 1995; Pisias 1978, 1979). In contrast, Raab and Larson (1997) suggested that punctuated cultural changes during this interval were stimulated by increasing violence and competition between individuals resulting from widespread drought and the associated reduction of terrestrial resources (also see Jones and Kennett 1999; Jones et al. 1999; Kennett and Kennett 2000; Yatsko 2000). This debate is considered within an evolutionary and ecological framework, informed by HBE, and in light of new paleoclimatic data for the region (Kennett and Kennett 2000). The paleoclimatic data are used to evaluate the environmental context for cultural stability and change during the Holocene. Most important, they indicate that the

8

THE ISLAND CHUMASH

table 1. Chronology for Santa Barbara Channel region including northern Channel Islands

Terminal Pleistocene Early Holocene Early Period Middle Period Late Middle Period Middle/Late Transition Late Period Historic

BC–AD

BP

BC 11,000–8000 BC 8000–6120 BC 6120–490 BC 490–AD 1150 AD 660–980 AD 1150–1300 AD 1300–1782 AD 1782–

13,000–9,950 9,950–8,060 8,060–2,440 2,440–800 1,290–970 800–650 650–168 168–

sources: Arnold 1992a; Erlandson and Colten 1991; Kennett 1998; King 1990.

interval from 800 to 650 years ago, prior to the dominance of sociopolitical complexity in the region, was marked by noticeably cool, highly variable marine conditions associated with high marine productivity. These new paleoenvironmental data are interpreted along with broad patterns of cultural evolution during the Holocene (10,000 to 650 years ago), including increasing population levels, intensified fishing, and decreased settlement mobility. A behavioral–ecological model is proposed for the emergence of sociopolitical complexity that emphasizes competitive and cooperative responses to strong climatic variability, including sustained terrestrial drought and high marine productivity that mark the interval between 1,500 to 650 years ago. Larger and more sedentary settlements emerged on the northern Channel Islands during this period in the context of increased regional violence (Lambert 1994, 1997; Lambert and Walker 1991). Violence was exacerbated by the introduction of the bow and arrow sometime between 1,500 and 1,300 years ago. Islanders responded to these highly unstable environmental and social conditions in unique ways, competing violently for available resources in some cases and apparently cooperating in others. Reduced settlement mobility, intensified fishing, and the production of exchange items emerged as alternative strategies for dealing with environmental and social instability. Cooperative strategies, such as trade, associated with the new economic system became dominant after 650 years ago as violent interaction decreased regionally. It is within this context that certain individuals were able to enhance their social status and at times control aspects of the political and economic relationships for their own benefit. The basis for this hypothesis is outlined in this volume.

THE ISLAND CHUMASH

9

The ages in this study (table 1) are expressed as calibrated calendar years before present (BP). This is somewhat unconventional in the study of Channel Islands prehistory where calendar years BC/AD or radiocarbon years before present are commonly used. The decision to use calibrated calendar years before present was largely driven by the broad use of this dating system by paleoclimatologists and the belief that the use of a religious calendrical system is inappropriate within this particular context. A more detailed description of the chronology is presented in chapter 4, but table 1 summarizes the chronological framework used in this study with cross-references between BP and BC/AD dates.

chapter 2

Human Behavioral Ecology and Maritime Societies

As early as the 1920s, David Banks Rogers had identified general trends in population growth, intensified economic strategies (e.g., fishing), and cultural elaboration evident in the archaeological record of the Santa Barbara Channel region (Rogers 1929). Since that landmark study, archaeologists working in the region have continually refined cultural chronologies and asked more specific questions about when and why these developments occurred (Arnold 1987; Erlandson and Jones 2002; Gamble et al. 2001; Glassow 1997; King 1990). Many of these theories have invoked environmental, demographic, and social variables to explain the origins of the economic, social, and political strategies observed at historic contact (Arnold 1987, 1991, 1992a, 1993, 1995, 1996, 2001; Erlandson 1994; Glassow and Wilcoxon 1988; Glassow et al. 1988; Johnson 2000; King 1990). Some emphasize “prime movers,” such as population growth or environmental change, to explain cultural developments (Glassow et al. 1988), whereas others are founded on more complex ecological and social interactions (Arnold 1987, 1993, 2001; Raab and Larson 1997). All of these are processual or adaptationist theories (Binford 1989, 2001; see Kantner 1996), in the sense that they invoke group-level adaptive responses to internally or externally generated stimuli (i.e., population increase or environmental change). Except for Arnold (2001; also see Arnold and Green 2002), who has focused on elite individuals as important evolutionary agents, these models also largely ignore the role of individual decision making and behavioral variability. 10

HBE AND MARITIME SOCIETIES

11

Human behavioral ecology (HBE) provides an alternative to adaptationist approaches and is employed here to explore cultural stability and change on the northern Channel Islands of California. The approach offers a well-developed set of models for studying behavioral variability and individual adaptations within varying ecological contexts (Bettinger and Richerson 1996; Smith and Winterhalder 1992; Winterhalder and Smith 1992). In essence, HBE is the study of evolution and adaptive design of human behavior within specific ecological contexts (broadly defined) and is useful for formulating hypotheses and identifying important environmental and social selective pressures of the past (Winterhalder and Smith 2000, 51). The approach is firmly grounded in neo-Darwinian principles, particularly natural selection theory (Winterhalder and Smith 1992), and provides an expanding set of models that can be tested with ethnographic or archaeological data (Bettinger and Richerson 1996; Broughton and O’Connell 1999; Winterhalder and Smith 2000). A review of HBE in the field of anthropology indicates both that it is a progressive and developing research tradition and that it is making large contributions to our understanding of human behavioral diversity and design (Winterhalder and Smith 2000). Starting in the 1970s, the first generation of behavioral ecologists adapted optimal foraging models developed in the biological sciences (Charnov 1976; MacArthur 1960; MacArthur and Pianka 1966) to questions regarding hunter-gatherer prey selection and land use (Hames and Vickers 1982; Hawkes et al. 1982; O’Connell and Hawkes 1981; Simms 1987; Smith and Winterhalder 1981; Winterhalder 1977; Winterhalder and Smith 1981). The interest in foraging strategies continues and has become more refined (Barlow and Metcalfe 1996; Barlow et al. 1993; Begossi 1992; Bird 1997; Bird and Bliege Bird 1997, 2000; Bliege Bird and Bird 1997; Broughton 1994b, 1997, 1999; Cannon 2000, 2003; Cashdan 1990, 1992; Madsen and Schmitt 1998; Metcalfe and Barlow 1992; Nagaoka 2002a, 2002b; Smith 1991; Ugan and Bright 2001; Zeanah 2000), but the repertoire of topics addressed within an HBE framework has also expanded considerably to include: conservation biology (Aswani 1998; Smith and Wishnie 2000; Winterhalder and Lu 1997); agroecology (Tucker 2001); plant and animal domestication and the transition to agriculture (Alvard and Kuznar 2001; Barlow 1997, 2002; Gremillion 1996, n.d.; Kennett and Winterhalder n.d.; Kennett et al. n.d.b; Piperno and Pearsall 1998; Winterhalder and Goland 1993, 1997); land tenure (Aswani 1999; Cashdan 1983; Eerkens 1999; Kennett and Clifford 2004); exchange (Winterhalder 1997; Winterhalder et al.

12

HBE AND MARITIME SOCIETIES

1999); colonization and migration (Aswani and Graves 1998; Kennett et al. n.d.a); costly signaling and cultural complexity (Bliege Bird and Bird 1997; Bliege Bird et al. 2001; Boone 1992; Hawkes and Bliege Bird 2002; Neiman 1997; Smith and Bliege Bird 2000; Smith et al. 2003; Sosis 2000); and even prehistoric cannibalism (Kantner 1999a, 1999b). In archaeology, HBE is one of several Darwinian approaches used to explore cultural evolutionary processes and there is considerable debate as to the direction and cohesiveness of the paradigm (Bamforth 2002; Barton and Clark 1997; Bettinger and Richerson 1996; Boone and Smith 1998; Broughton and O’Connell 1999; Dunnell 1971, 1980; Maschner 1996; Schiffer 1996; Shennan 1989, 2002; Smith 2000; Teltser 1995a, 1995b). Dunnell’s (1971, 1980) selectionist archaeology is extreme in its position that natural selection is the dominant cultural evolutionary force (also see Lyman and O’Brien 1998; Neff 2000; O’Brien and Holland 1992, 1995; O’Brien and Lyman 2002; Teltser 1995a). More moderate Darwinian approaches give thought to environmental and social contextual information to varying degrees (Barton and Clark 1997; Bettinger 1991; Boone 1992; Kantner 1996; Maschner 1996; Maschner and Patton 1996; Shennan 1989, 2002). Behavioral ecologists consider natural selection the fundamental mechanism shaping human behavior without compromising the social and environmental context in which selection operates. One of the most appealing aspects of HBE is its attention to the ecological interaction between humans and their social and natural environments (Broughton and O’Connell 1999). In a sense, behavioral ecology provides the evolutionary and adaptive mechanisms (evolution by natural selection) that were often missing in cultural ecological studies (e.g., Jochim 1976, 1981; Netting 1977; Steward 1955). Cultural ecology was grounded in the functionalist thought of the 1950s (Trigger 1989), focusing primarily on group adaptation and seeking direct (causal) associations between cultural and environmental structure. Although this approach has been repeatedly criticized as “normative” (Brumfiel 1992, 1994; Mithen 1989; Wobst 1978), some archaeologists continue to use cultural ecological principles, explicitly or implicitly, to explore cultural systems around the world. HBE differs from cultural ecology in its emphasis on individual behavioral variability and evolutionary processes, particularly selection, that shape human societies (Kelly 1995). Visible adaptations at the group level are a product of selection operating on individuals (methodological individualism; Mithen 1989; Nettle 1997). Essentially, HBE changes the focus of

HBE AND MARITIME SOCIETIES

13

research from group to individual behavioral strategies and provides a mechanism for cultural persistence and change that is absent in cultural ecological studies. HBE is employed here for investigating prehistoric human behavior and cultural change in a specific maritime setting: the northern Channel Islands of California. It is also an attempt to demonstrate the applicability and flexibility of HBE for exploring maritime foraging strategies and cultural evolutionary processes in coastal settings more generally. Evidence for systematic human use of coastal and aquatic resources worldwide extends back roughly 150,000 years, and these new adaptive strategies likely played an important role in the expansion of anatomically modern humans out of Africa and into Eurasia, Australia, and the Americas (Erlandson 2001, 2002b). The use of coastal habitats appears to have intensified in many areas during the past 10,000 years, and more sophisticated seafaring techniques and specialized fishing and marine mammal hunting technologies developed independently in several areas. Historical and archaeological data from a variety of regions suggest that large population concentrations occurred in coastal settings and that maritime hunter-gatherers were often more sedentary and culturally complex than their interior neighbors (Birdsell 1953; Erlandson 2001; Pálsson 1988; Yesner 1980). In the past, some archaeologists have incorrectly described all aquatic and marine resources as marginal and of low nutritional and caloric value (e.g., Osborn 1977). In some regions marine foods may be ranked relatively low in comparison to terrestrial alternatives, but this is not always the case and all marine foods cannot be easily categorized as high or low ranked. HBE provides a framework for exploring this variability. Ethnographers commonly use HBE models to interpret and predict hunter-gatherer foraging behavior in terrestrial habitats (e.g., Hawkes et al. 1982, 1991; Simms 1987). Interest in maritime foraging strategies has lagged behind, possibly due to the complexities associated with the spatial distribution of marine resources and the difficulties involved with prey detection and mobility in these habitats (Aswani 1998; Kelly 1995). Smith’s (1991) study of Inujjuamiut foragers in the Canadian Arctic is one exception, as is Beckerman’s (1983, 1991) work on optimal group size among Barí fishermen in Columbia and Sosis’ (2000, 2001, 2002) work on Ifaluk Atoll in Micronesia. Fishing and shellfishing strategies are also receiving increasing attention by behavioral ecologists (Aswani 1998, 1999; Begossi 1992; Bird and Bliege Bird 2000; de Bour 2000; de Bour and Longamane 1996) and a variety of behavioral strategies, beyond

14

HBE AND MARITIME SOCIETIES

foraging, are being addressed in maritime settings, including food sharing (Bliege Bird and Bird 1997) and costly signaling (Bliege Bird et al. 2001; Smith and Bliege Bird 2000; Sosis 2000). Many of the underlying microeconomic principles of HBE are employed to varying degrees by archaeologists working in maritime settings (Erlandson 2001; Hildebrandt and Jones 1992; Raab 1992), and Broughton (1994a, 1994b, 1999) has explicitly tested the diet-breadth model with archaeological faunal assemblages from the San Francisco Bay Area. Some anthropologists and archaeologists have questioned the use of HBE, particularly optimal foraging theory (OFT; Dwyer 1985; Erlandson 1991b; Jochim 1981; Rick et al. 2001a), because it oversimplifies the complexities of human behavior and the historical contingencies involved in cultural evolutionary processes. The purpose of HBE is not to reduce all aspects of human behavior into foraging equations but to explore specific questions with a set of models that can be tested with ethnographic or archaeological data. In this study, the primary questions addressed with the archaeological record on California’s northern Channel Islands are: (1) How and why did subsistence strategies change through time (e.g., intensification)? (2) How did settlement patterns and land use change over time? (3) What role did demographic change, competition, and cooperation have in shaping these patterns? and (4) How were these changes related to the emergence of social hierarchies evident at historic contact? To address these issues, I use a nested series of HBE models (diet choice, central place foraging, the ideal free distribution) that provide a set of explicit predictions to be tested with archaeological data from the northern Channel Islands of California. Each of these models is outlined in the next section with special attention to maritime societies in general.

Maritime Foraging Strategies For the vast majority of human history, people have foraged for naturally occurring wild resources. Many different foraging strategies have been documented among living people and the archaeological record indicates that this is but a small fraction of the diversity that once existed (Bettinger 1991; Jochim 1976; Kelly 1995; Lee 1968; Price and Brown 1985). Between about 11,000 and 8,000 BP, in several parts of the world, some foraging populations started adopting food production or were replaced by agriculturalists (Bellwood 2001; Kennett and

HBE AND MARITIME SOCIETIES

15

Winterhalder n.d.), but hunting, gathering, and fishing persisted in certain regions such as California’s Channel Islands. Although agriculture was never adopted in these regions, the archaeological record commonly indicates that foragers intensified their subsistence pursuits during the Holocene. In coastal settings, this often involved the development of more sophisticated fishing technologies and intensified exploitation of near- and offshore fisheries. Maritime foragers once occupied a variety of different coastal habitats from the tropics to higher and lower latitudes (e.g., Arctic, Tierra del Fuego). Coastal habitats vary greatly, but these ecological systems provided foraging peoples with some common opportunities and constraints. Qualities shared by many coastal environments include highresource biomass, large resource diversity, and great ecological stability (Yesner 1980). High primary productivity along coasts results from the mixing of surface waters with nutrient-rich, deeper ocean water—a process that is promoted by ocean current systems and coastal upwelling. Due to this high productivity, a wide variety of resources are concentrated along coastlines (shellfish, fish, sea mammals) and migratory birds are often attracted to these rich ecological zones seasonally. Marine systems also have a great deal of inertia and, because of this, are commonly more stable than adjacent terrestrial environments. Although coastal environments are generally productive, resources are often unevenly distributed or patchy. Coasts can be convoluted in ways that create large bays and sheltered harbors, or they may be open and exposed to heavy surf and gusty wind. Rocky intertidal zones and estuaries commonly punctuate coastlines and can be up to ten times more productive than other coastal habitats (e.g., beaches; Lieth and Whittaker 1975; Odum 1971; Yesner 1980). Coastal upwelling of nutrient-rich water also results in highly localized marine productivity, and submarine canyons can funnel nutrient-rich water to create highly productive patches along the coast (e.g., kelp beds). Other important variables include tidal range, nearshore topography (steep or shallow), and the length of coastline available in a foraging area. Coastal environments also vary with respect to available terrestrial resources. For example, the highly productive coastal zone of southern Peru is juxtaposed against one of the driest deserts in the world (Keefer et al. 1998, 2003; Moseley 1975; Sandweiss 2003; Sandweiss et al. 1996, 2001). The virtual absence of terrestrial resources in this ecological context constrained foraging behavior in a way that was different from other terrestrial landscapes containing abundant and diverse resources. Fundamental HBE principles predict

16

HBE AND MARITIME SOCIETIES

that foragers living in coastal environments weigh the costs and benefits of the known subsistence alternatives. Individuals will tend to act rationally and economically in their judgments regarding resource or patch selection (Bettinger 1991), but subsistence decisions are constrained by the spatial and temporal availability of resources and the subsistence and settlement activities employed by other people in the region.

Diet Choice in Maritime Settings The diet-breadth model provides a point of departure for exploring prehistoric changes in resource selection on the northern Channel Islands and coastal environments more generally. It is one of several models, developed in the biological sciences (e.g., MacArthur and Pianka 1966), that exist for the analysis of food choice. The model predicts how many different food types yield the best average energetic return rate under specific conditions of resource density (Kaplan and Hill 1992; Stephens and Krebs 1986). Application of the diet-breadth model to human foraging decisions has a relatively long history and continues to be tested in a variety of ethnographic situations (de Boer 2000; Hill et al. 1987; Winterhalder 1981a, 1981b, 1986; Winterhalder and Smith 2000). The results of these studies suggest that the fundamental microeconomic principles underlying the model are robust and that they make a number of useful predictions about food choice that are sometimes counterintuitive (Winterhalder 1986). It has also been used by archaeologists, with varying degrees of success, to explore prehistoric dietary choices and changes in diet through time, particularly population-dependent increases in diet breadth and intensified use of some food types (Broughton 1999; Butler 2000, 2001; Hildebrandt and Jones 1992; Keene 1981; Kelly 1995; Piperno and Pearsall 1998; Raab 1992; Simms 1987). Feeding behavior is complex, particularly among humans, and the diet-breadth model simply provides a starting point for exploring dietary decisions in the past and present. Human foragers generally live in environments that have a variety of potential subsistence resources that vary in abundance and value (protein, calories, carbohydrates, etc.). Wild plant and animal foods also vary with respect to the time and energy needed to extract and process them. The diet-breadth model predicts that a forager will choose the combination of foods that maximizes the net intake of energy (calories), that is, the number of calories acquired minus the energy expended acquiring them

HBE AND MARITIME SOCIETIES

Excluded

Included

Return Rate

17

Search Time Optimum

Handling Time

1

2

3

4

5

6

7

8

Resource Ranking figure 2. Graphical representation of the diet-breadth model (drafted by D. Kennett).

(search and handling time; figure 2). Highly ranked resources are those that provide large caloric gains with little energy expenditure. A forager will never pass over a high-ranked resource for a lower-ranked one. If resources are scarce, then the time spent in search of them (search time) may outweigh overall dietary utility. Therefore, the use of lower-ranked resources is contingent upon the availability of resources that initially had a higher rank. The model can also accommodate suboptimal foraging behavior by certain members of a group in the sense that it predicts that evolutionary forces will favor a tendency toward optimal behaviors. These may be manifested as “rules of thumb” that develop through trial-and-error experimentation with subsistence resources (Winterhalder and Smith 2000). Application of the diet-breadth model to prehistoric contexts is no simple matter, and the model must be adjusted to fit some of the limitations imposed by the archaeological record (Jochim 1981;

18

HBE AND MARITIME SOCIETIES

Winterhalder and Smith 2000). Many of the initial applications of the diet-breadth model suffered from the “bigger is better” misunderstanding (Erlandson 2001); larger animals were considered to be higher ranked simply because of their overall body size. In coastal settings, this relegated shellfish and other small aquatic organisms to the position of “starvation foods” (Osborn 1977). Simplistic characterizations of marine resource value are certainly unwarranted, because higherranked resources are not necessarily larger, but reflect size, abundance, and costs associated with exploiting them (Madsen and Schmitt 1998). This points to one of the fundamental problems with using this type of model—archaeologists cannot directly measure rates of return or determine procurement, processing, and manufacturing costs associated with exploiting specific resources (Jochim 1981; Lindström 1996). The costs and benefits of exploiting certain types of resources must be inferred from ethnographic studies where this kind of economic data is collected and quantified and/or from experimental studies that estimate resource return rates for labor investment in producing various technologies (table 2). Ethnographers are beginning to collect these data in coastal settings (Bird and Bliege Bird 1997, 2000; de Boer 2000), and there have been several experimental studies using different fishing technologies and shellfish harvesting strategies (Erlandson 1988a; Jones and Richman 1995; Lindström 1996; see below). Once reasonable return rates are established for different food types, the diet-breadth model can be used to make several predictions about prehistoric foraging behavior that can be tested with archaeological data. The model predicts that as the abundance of higher-ranked prey species increases, the variety of resource types will decrease. It follows that if the abundance of high-ranked resources decreases, thus increasing search time, diet breadth will increase. This means that potential prey types will enter a forager’s diet based on the abundance of higherranked resources and not simply because of their immediate value. The availability of high-ranked resources is often population dependent. In other words, as the density of the human population increases the availability of the highest-ranked prey species will usually decrease. Climate change may also decrease the availability of high-ranked prey as was likely the case at the end of the Pleistocene epoch throughout much of the New and Old Worlds (Piperno and Pearsall 1998). Formal applications of the diet-breadth model in coastal settings have been used to explain: (1) inclusion of lower-ranked prey in the San Francisco Bay Area as large game animals (deer and sea mammals)

HBE AND MARITIME SOCIETIES

19

table 2. Post-encounter return rates in calories per hour for various terrestrial plant and animal resources from Great Basin environments Resource deer sheep antelope jackrabbit gopher rabbit pollen, cattail squirrel squirrel waterbird, ducks seeds, gambel oak seeds, tansymustard seeds, pinyon pine roots, bitterroot seeds, salina wild rye seeds, shadscale seeds, shadscale seeds, bulrush seeds, barnyard grass seeds, peppergrass seeds, sunflower seeds, bluegrass seeds, wild rye seeds, ricegrass seeds, reed canary grass seeds, scratchgrass seeds, foxtail barley seeds, sedge roots, cattail roots, bulrush seeds, saltgrass seeds, pickleweed seeds, squirreltail grass

Scientific Name

Cal/person/hour

Odocoileus hemionus Ovis canadensis Antilocapra americana Lepus sp. Thomomys sp. Sylvilagus sp. Typha latifolia Spermophilus sp. Citellus sp. Anas sp. Quercus gambelli Descurainia pinnata Pinus monophylla Lewisia rediviva Elymus salinas Atriplex nuttalli Atriplex confertifolia Scirpus sp. Echinochloa crusgalli Lepidium fremontii Helianthus annuus Poa sp. Elymus cinereus Oryzopsis hymenoides Phalaris arundinacea Muhlenbergia asperifolia Hordeum jubatum Carex sp. Typha latifolia Scirpus sp. Distichlis stricta Allenrolfea occidentalis Sitanion hystrix

17,971–31,450 17,971–31,450 15,725–31,450 13,475–15,400 8,983–10,780 8,983–9,800 2,750–9,360 5,390–6,341 2,837–3,593 1,975–2,709 1,488 1,307 841–1,408 1,237 921–1,238 1,200 1,033 302–1,699 702 537 467–504 91–418 266–473 92–301 261–321 162–294 138–273 202 128–267 160–257 60–146 90–150 91

source: Simms 1987.

became depleted prehistorically (Broughton 1994a, 1994b, 1997, 1999); (2) use of smaller marine snails as larger shellfish species (Haliotis) became increasingly rare on the southern Channel Islands of California (Raab 1992); (3) changes in the targeted fish species through time, from large to small, in the Cook Islands of Polynesia (Butler 2001); and (4) the reduction in mainland sea mammal rookeries in California and the ultimate development of watercraft for exploiting offshore breeding colonies (Hildebrandt and Jones 1992; Jones and Hildebrandt 1995; Porcasi et al.

20

HBE AND MARITIME SOCIETIES

2000). Many of these studies document gradual increases in diet breadth through time as higher-ranked prey items became depleted, usually due to intensified human predation. This may be the case in certain contexts, but tendencies toward optimization are often contingent upon environmental and social contexts that are stochastic or punctuated. Although the diet-breadth model is primarily used to predict prey choice, these dietary choices also have ramifications for settlement decisions. If higher-ranked resources become depleted, foragers will choose a broad diet, or move their settlement (if possible), rather than spend large amounts of time searching for higher-ranked foods (Piperno n.d.). Reductions in search time may also prompt greater investment in storage and food processing and can promote residential stability or greater sedentism. One of the assumptions of the diet-breadth contingency model is that individuals make decisions during a limited period of time in very specific, spatially limited, environmental and cultural contexts. The sum of these decisions across space and through time results in the longterm subsistence shifts that may be evident in the archaeological record.

Return Rates for Marine Resources Estimated energetic return rates versus labor for various resources are a fundamental component of modeling subsistence change using the diet-breath model. Several studies have compiled return rates for terrestrial plant and animal resources (Piperno and Pearsall 1998; Simms 1987; see table 2). These return rates are used here as a point of comparison for estimated rates of return for the most common aquatic/marine resources (shellfish, fish, and sea mammals) based on several studies (Bird and Bliege Bird 2000; Lindström 1996; Smith 1991). The resource structure in coastal settings is complex and highly varied worldwide, and there is considerable disagreement regarding the dietary value of marine prey species. In addition to terrestrial plants and animals, coastal environments contain shellfish, fish, sea mammals, seabirds, and amphibians and also attract a variety of waterfowl seasonally. Modern foragers exploit a wide range of marine resources (Bird and Bliege Bird 1997, 2000; Bliege Bird et al. 2001; de Boer 2000; Hockey et al. 1988), and foraging decisions in coastal habitats depend upon the availability and character (e.g., clumped, dispersed) of resources in both marine and terrestrial environments and the costs of procuring them given the available technology.

HBE AND MARITIME SOCIETIES

21

shellfishing Shellfish are often the most visible marine resources in prehistoric coastal midden deposits, and their dietary importance is a subject of great debate (Erlandson 2001). The general category shellfish includes a variety of organisms (crab, shrimp, sea urchins, etc.), but prehistoric shell middens are often dominated by marine mollusks (bivalves and gastropods). These organisms range in size (generally small) and often occur in large aggregations (or beds) relatively close to shore or easily accessible in the intertidal zone. The location of these beds is highly predictable, and these organisms are available throughout much of the year as a spatially clumped and highly productive resource. Seasonal fluctuations in meat weight can occur in some species (Claassen 1986; de Boer 2000), and inclement weather, seasonal tidal patterns, or bacterial infestations (red tide) can restrict their access to foragers (Bird and Bliege Bird 2000). Sophisticated technologies are not required to procure mollusks, and all members of a society (men, women, children, and the elderly) can participate in this subsistence pursuit (Bird and Bliege Bird 1997, 2000; de Boer 2000; de Boer and Longamane 1996; Meehan 1982; Moss 1993). There is considerable variation between species, but shellfish tend to be high in protein and low in fat, carbohydrates, and calories (table 3; Erlandson 1988a, 2001; Sidwell 1981; Watt and Merrill 1975). Seasonal variations in the carbohydrate content and nutritional value are also documented in some species of shellfish (Claassen 1986). The direct relationship between carbohydrates and calories has led to the erroneous characterization of shellfish as a universally low-ranked resource (Osborn 1977). Shellfish beds in highly productive marine environments (e.g., Peru and California) can be densely packed, and the exploitation of these habitats would certainly be energy efficient. Several studies provide information regarding the return rates on the collection of shellfish. Based on experimental work, Jones and Richman (1995) determined that mussel beds on the central coast of California produced about 500 cal/person/hour. In a classic ethnographic study of shellfishing among the Anbara of the Northern Territory of Australia (Arnhem Land), Meehan (1977, 524) determined that skilled women collecting shellfish (Tapes hiantina) had return rates in the 1,000 cal/person/hour range. Bird and Bliege Bird (2000) quantified the relative return rates for different shellfish species from the reefs of the Meriam Islands of Australia (Torres Strait). The profitability for adults collecting shellfish was ranked according to energy gained per unit of

22

HBE AND MARITIME SOCIETIES

table 3. Nutritional information for plants and animals found in the Santa Barbara Channel region Scientific Name Resource H2O Protein

Fat

Carb. Kcal

Other Nutrients

Marine Mammals, Fish, Sharks Phoca vitulina Genyonemus lineatus Mustelus vulgaris Notorynchus maculatus Rhinobatus sp. Sardinops sp. Sebastes sp.

Harbor seal Croaker



26





143

Fe (20 mg)

79.7

18

0.8

0

79

No data

Shark

77.1

19.7

0.9

1.1

91

P (309 mg)

Shark

67.9

15.3

13.1

2.5

189

No data

Shark Sardine Bocaccio

76.2 70.7 79.5

16.2 19.2 18.9

6.4 8.6 1.5

0 0 0

122 160 89

No data K (560 mg) K (411 mg)

Mussel Cockle

74.6 78.8

14.4 13.5

2.2 0.7

3.3 4.7

95 79

K (315 mg) Fe (16 mg)

Abalone Chiton

75.8 60

18.7 22

0.5 16.3

3.4 0

98 234

P (191 mg) No data

Oyster

81

9.6

2.5

5.4

82

P (178 mg)

Clam

79.4

13.5

1

3.5

77

No data

Clam

80

13

1.2

4.1

79

No data

Clam

81.9

11.2

1.4

4

74

Ca (607 mg)

Shellfish Mytilus sp. Clinocardium nuttalli Haliotis sp. Mopalia muscosa Ostrea lurida Protothaca staminea Saxidomus nuttalli Tivela stultorum

source: Erlandson 1988a.

handling time (cal/person/hour). These data indicate that the profitability of different shellfish in this context varies greatly (table 4). Some species (Tridacna gigas) have return rates that are comparable to large game animals in other environments (see table 2), while the profitability of others falls more in the range of low-ranked plants. Based on this study, the return rates for most shellfish are comparable to mediumsized game or small game/plants. One of the significant results of this study was that adults will pass over those shellfish species encountered that are below the overall reef flat collecting return rate (Bird and Bliege Bird 2000, 467), thus providing support for the diet-breadth contingency model.

HBE AND MARITIME SOCIETIES

23

table 4. Return rates in calories per hour for select shellfish species from the Meriam Islands Scientific Name

Habitat

Cal/person/hour

Range (1 sigma)

Tridacna gigas Hippopus Tridacna spp. Large Trochus Lambis Cypraea Small Trochus Tridacna crocea Strombus Asaphis Nerita

Reef flat Reef flat Reef flat Reef flat Reef flat Reef flat Reef flat Reef flat Reef flat Rocky shore Rocky shore

14,100 6,200 3,800 3,900 3,000 2,100 950 600 500 400 300

16,250–10,000 8,900–5,000 2,500–5,000 3,000–3,900 3,200–2,500 2,100–1,800 1,000–800 650–550 600–400 425–375 275–325

source: Bird and Bliege Bird 2000.

fishing The viability of fishing relative to other subsistence strategies also has received considerable attention over the years, particularly as it relates to the broad-spectrum revolution during the Early Holocene and the ultimate emergence of more complex societies in coastal settings worldwide (Erlandson 2001; Moseley 1975; Osborn 1977; Pearlman 1980). This is due, in part, to the great resource potential that many fisheries have if appropriate fishing technologies exist. Some archaeologists consider fish to be inferior to other terrestrial resources because of their elusive nature and the sophisticated technology required for capturing and retrieving them (Kelly 1996). Fish are sometimes elusive and difficult to catch, but they can also be an incredibly productive and predictable resource that can easily be taken with relatively simple technologies (Erlandson 2001; Rick et al. 2001a). They can also be dried and stored easily for future use. It is also possible that the labor investment in the production of fishing technology (e.g., nets, weirs, hooks, and lines) also contributed to the seemingly late development of fishing in human history (Osborn 1977). Erlandson (2001) reviewed the history of aquatic adaptations and has questioned the idea that fishing is exclusively a late phenomena. The earliest evidence for fishing comes from the site of Katanda (Yellen 1998; Yellen et al. 1995), where barbed harpoons and fish remains are dated to ~80,000 years ago. Erlandson (2001, 334) noted that current evidence suggests that the first appearance of

24

HBE AND MARITIME SOCIETIES

anatomically modern humans outside of Africa (~90,000 years ago— Qafzeh and Skhul caves in coastal Israel)* is temporally coincident with the first evidence for more sophisticated fishing technologies. There certainly is clear evidence for the intensified use of marine fisheries during the past 10,000 years (Bailey and Craighead 2003; Kennett and Kennett 2000; Moseley 1975; Spriggs 1997); however, fishing has always played an important subsistence role for some anatomically modern humans living in coastal contexts. Behavioral ecologists ask the question, under what conditions were fish a cost-effective resource to exploit? (Lindström 1996, 114). Similar to shellfish, fish tend to be high in protein, vitamins, and minerals but low in carbohydrates (see table 3; Erlandson 1988a; Sidwell 1981; Watt and Merrill 1975). Fish flesh is easily metabolized and high consumption rates in some modern populations appear to be correlated with lower incidents of disease and longer life spans (Erlandson 2001). Return rates for fishing are dependent upon the overall productivity and character of the targeted fish species and the technology available at the time of capture or harvest. A variety of sophisticated fishing techniques and devices are known from ethnographic and ethnohistorical studies around the world (Aswani 1998; Lindström 1996; Winterhalder 1981b). Lindström (1996) conducted the most comprehensive study of return rates for fishing with ethnohistorical and experimental data from the Truckee River basin on the western edge of the Great Basin (California/Nevada). Although these data were collected from a riverine context, they provide at least a proxy for fishing in maritime settings. Lindström calculated return rates for different kinds of fish (tiny to large) using different types of fishing technology (table 5). These estimates do not include the time invested in producing and maintaining fishing equipment. The key observation is that return rates for fish are highly variable and depend on the technology used, the season of capture (spawn or nonspawn), and the size of fish. It is also worth comparing these return rates to those for terrestrial resources listed in table 2. In certain contexts and with certain types of technology, fish can clearly compete with large- and mediumsized game and almost always have higher return rates than plants (Rick et al. 2001a).

*A tilde (~) is used here and throughout the remainder of the text as an indication of approximation of the specified age.

table 5. List of resource return rates for different fishing strategies from the Truckee River basin of Great Basin (western edge) Fish Size

Species

Attended

Capture

Spawn

Technology

Cal/person/hour

Small Large Large Large Large Small Small Large Medium Medium Small Medium Tiny Medium Small Medium Large Tiny Tiny Tiny Medium Tiny

TC/TS C/LCT C/LCT C/LCT C/LCT TC/TS TC/TS C/LCT MW/TC/TS MW/TC/TS TC/TS MW/TC/TS MS/PS/LRS/SD MW/TC/TS TC/TS MS/TC/TS C/LCT MS/PS/LRS/SD MS/PS/LRS/SD MS/PS/LRS/SD MW/TC/TS MS/PS/LRS/SD

Yes No No Yes No No No Yes No No Yes Yes Yes No No Yes Yes No No Yes Yes No

Mass Mass Mass Mass Mass Mass Mass Individual Mass Mass Mass Mass Mass Mass Mass Individual Individual Mass Mass Mass Individual Mass

Yes/no Yes/no Yes/no Yes Yes/no Yes/no Yes/no Yes Yes/no Yes/no Yes Yes Yes/no Yes/no Yes/no Yes Yes Yes/no Yes/no Yes No Yes/no

Basket scoop Basket trapa Gill net Bag, dip, or lift netb Multiple hook and line Basket trapa Gill net Weir and spear Basket trapa Gill net Bag, dip, or lift netb Bag, dip, or lift netb Basket scoop Multiple hook and line Multiple hook and line Spear/harpoonb Spear/harpoon or hook and lineb Basket trapa Gill net Bag, dip, or lift netb Spear/harpoon or hook and lineb Bag, dip, or lift netb

360,400–243,800 84,994–34,147 73,908–29,694 67,996–27,318 51,512–20,696 51,486–34,829 27,723–18,754 25,563–10,270 24,740–19,561 21,513–16,020 21,200–14,341 19,791–15,649 16,790–1,976 14,994–11,855 10,922–7,388 7,440–5,883 3,301–1,327 2,399–282 1,292–152 988–116 961–760 509–60

source: Lindström 1996. These rates of return vary greatly depending upon season and capture technology used. Fish species: TC, tui chub (Gila bicolor); TS, Tahoe sucker (Catostomus tahoensis); C, Cui-ui (Chasmistes cujus); LCT, Lahontan cutthroat trout (Salmo clarki henshawi); MW, mountain whitefish (Prosopium williamsoni); MS, Mountain sucker (Catostomus platyrhynchus); PS, Paiute sculpin (Cottus beldingi); LRS, Lahontan redside (Richardsonius egregius); SD, speckled dace (Rhinichthys osculus). Fish size: Large, 3.4–1.37 kg; medium, 0.86–0.68 kg; small, 0.34–0.23 kg; tiny, 0.017–0.002 kg. a With or without weir/platform or from bank. b With platform/weir or from bank.

26

HBE AND MARITIME SOCIETIES

sea mammal hunting The importance of marine mammal hunting for maritime people varies greatly and the relative role that these animals had in subsistence economies along the west coast of North America is a subject of a current and spirited debate (Colten and Arnold 1998, 2000; Erlandson et al. 1998; Hildebrandt and Jones 1992; Jones and Hildebrandt 1995; Lyman 1995; Porcasi et al. 2000; Walker et al. 2000). Marine mammals range in size from relatively small Sea Otters to large whales weighing over 100,000 kg (Erlandson 2001). The large body size of many marine mammals can make them a highly profitable prey type simply from the perspective of their high meat yield. They are comparable to land mammals in nutritional value and generally provide more calories per kilogram than shellfish and some fish (table 3). Marine mammals also provide essential nutrients (Fe, P, K), and many are high in fat that numerous indigenous coastal dwellers rendered into oil (used for food and fuel; see Erlandson 2001). In addition, many secondary products (hides, bones, teeth, ivory, and baleen) provided valuable raw materials with some (e.g., otter pelts) more highly prized than others. Based simply on body size, in combination with other secondary benefits, the profitability of marine mammals would appear to be exceedingly high. However, return rates are contingent on pursuit time and handling costs associated with capturing, killing, and butchering these large and sometimes dangerous animals. Marine mammal behavioral patterns also vary greatly and contribute to the relative costs associated with human exploitation. For instance, Gray Whales migrate up and down the west coast of North America in relatively large pods, whereas other whales (e.g., Humpback Whales) can be more solitary (Erlandson 2001). Certain types of seals and sea lions (e.g., Northern and Southern Fur Seals) must breed on dry land and often aggregate in large, spatially limited coastal rookeries that are highly vulnerable to human predators. Other marine mammals give birth in water (e.g., Sea Otters) or have pups that can swim immediately after birth (e.g., California Harbor Seals). The availability of many marine mammals is also highly seasonal (e.g., weather patterns, rookeries), and the supply of animals at certain times of the year may be well above the immediate needs of the local group. However, there are social and economic contexts during which large numbers of animals were taken for their valuable secondary products (e.g., hides, blubber; Erlandson 2001). Predation, by humans or

HBE AND MARITIME SOCIETIES

27

other animals, can also have rapid and deleterious effects on marine mammal populations, and even the presence of human predators can deter certain marine animals from hauling out or establishing breeding populations. Capturing some of these animals (e.g., whales) requires substantial watercraft and a relatively sophisticated technology that requires continual maintenance (Erlandson 2001). Marine mammal hunting is relatively uncommon today and ethnographers rarely discuss hunting practices in terms of return rates. Smith’s (1991) work with Inujjuamiut foragers from the Canadian Arctic is one exception that provides some insight into the economic decisions made by coastal hunters faced with a variety of prey species, including marine mammals, and limited time to pursue, capture, and process them. Smith (1991) collected detailed information on encounter rates, handling time, and search time for a variety of prey species hunted by the Inujjuamiut (terrestrial and marine), including Ringed Seals, Bearded Seals, and Beluga Whales. Based on these data, it is clear that the rank order of marine mammals (and other prey species) shifted over relatively short periods (e.g., seasons) and the inclusion of prey items was highly contingent upon the availability and productivity of higher-ranked resources. However, within specific seasons and within specific hunt types the dietbreadth contingency model was a reasonable predictor of prey choice. Inujjuamiut foragers primarily hunt marine mammals in summer and fall when they are most accessible (after the ice breaks up) and other higher-ranked terrestrial species (Caribou) are not as readily available. Table 6 summarizes the relative return rates of the marine mammal species taken during summer and fall (Smith 1991, 215–216). The return rates presented are not directly comparable to other studies because the Inujjuamiut use nontraditional technologies (e.g., rifles, freighter canoes, and outboard motors), and handling rates were calculated in a more sophisticated way (Smith 1991, 169–191). Marine mammals were hunted near shore during both of these seasons in a similar fashion— two to three men in a freighter canoe armed with rifles and harpoons. During both of these seasons marine mammals were among the highestranked prey items for male hunters, but their relative ranking changed based on a number of contingencies. These contingencies include: (1) the abundance and vulnerability of other prey species in the region (Canadian Geese in summer molt and therefore unable to fly); (2) encounter rates with various marine mammals (Ringed Seal higher in summer than fall, Bearded Seal highest in fall, Beluga Whale higher in fall); and (3) changes in the value of each species (Ringed Seals doubling

28

HBE AND MARITIME SOCIETIES

table 6. Return rates for various marine mammal species hunted by the Inujjuamiut in the Canadian Arctic Scientific Name Erignathus barbatus Phoca hispida Delphinapterus leucas

Common Name

Weight (kg)

Edible (kg)

kcala

Cal/person/ hour

Bearded Seal Ringed Seal Beluga Whale

207 33 499

55.2 12 101.7

70,610 20,390 169,630

8,480 4,550 7,450

a

Edible portion Smith 1991. Based on work with the Inujjuamiut in the Canadian Arctic.

SOURCE :

in value—fatter, providing more calories). Two species of marine mammal (Harp Seal and Walrus) were virtually never pursued because of either the high costs of traveling to offshore islands (Walrus) or behavior that makes them difficult to capture (Harp Seal). In addition, return rates from hunting Beluga Whale, were well below the mean return rate for higher-ranked resources in fall and, because of this, would not be expected in the optimal diet during fall months. Smith (1991, 217) argued that this may have resulted from a small sample size or contingencies specific to a particular hunt. It is also possible that whale hunters (men) ranked prey with a different currency, perhaps prestige rather than caloric value, as emphasized in costly signaling theory (see below; Smith and Bliege Bird 2000; Smith et al. 2003). The complexities of predicting prey choice with the diet-breadth model in this ethnographic case have direct implications for archaeologists trying to apply this model to the archaeological record. Some archaeologists in California argue that broadscale patterns of marine mammal exploitation during the Holocene are consistent with the basic predictions of the diet-breath model (Hildebrandt and Jones 1992; Jones and Hildebrandt 1995; Porcasi et al. 2000). Based on a variety of dated site components from the coast, Hildebrandt and Jones (1992) documented a decline through time in the relative frequency of larger marine mammal species (Steller Sea Lion, California Sea Lion, and Northern Fur Seal) and a relative increase in smaller species that are more difficult to capture (Sea Otters and Harbor Seals). They argued that population-dependent decreases in high-ranked prey species were responsible for the dietary inclusion of lower-ranked species (Hildebrandt and Jones 1992, 383). Criticisms of this model highlight some of the complexities of applying prey-choice models over spatially extensive areas with differing social and environmental contingencies

HBE AND MARITIME SOCIETIES

29

(Colten and Arnold 1998, 2000). However, studies of faunal assemblages at sites where large marine mammals were abundant when settlements were first established suggest that these animals were preyed upon heavily and that their availability was depressed through time compared to other lower-ranked species (Broughton 1999; Porcasi et al. 2000). Inconsistencies between local and regional datasets are likely explained by natural differences in the availability of these larger animal species, limited access to localized resources within another group’s territory, or constraints imposed by central place foraging. Constraints imposed by central place foraging are considered in the next section.

Central Place Foraging and Maritime Foragers The distribution of resources in some environments can be relatively continuous and undifferentiated, and the diet-breadth model may do reasonably well in predicting prey choice (Binford 1983; Kelly 1983, 1995; Winterhalder 1986). However, resource distribution (terrestrial and marine) in coastal habitats is often patchy and discontinuous—shellfish occur in highly productive clumps or beds, sea mammals haul out in groups or cluster in rookeries, and many fish periodically school or are associated with discrete types of marine habitats (reefs, kelp beds, etc.). In these contexts, simple diet-breadth models are often poor predictors of overall prey choice. For instance, in the Inujjuamiut case described above, Smith (1991) determined that the diet-breadth model was a good predictor of prey choice within specific resource patches or types of hunts. In other words, in patchy environments diet-breadth models are likely to do well at the local, rather than regional, level. This means that archaeologists need to be aware of the spatial variability in resources prehistorically and consider the environment in the immediate vicinity of each settlement. Indeed, diet-breadth models are often excellent predictors of changes in diet breadth through time at individual archaeological sites (Broughton 1999). The concept of resource patchiness is well defined in HBE, a resource patch being a discrete and bounded location on a landscape that has economic value to the organism being studied (Winterhalder 1994). The type, size, productivity, and distribution of resources affect behavior, human or otherwise. Foragers are confronted with a variety of patch types with respect to total obtainable energy. As with resources, resource patches can be ranked according to the net intake of energy minus

30

HBE AND MARITIME SOCIETIES

foraging and processing time. Foraging returns begin to decline once a resource patch is used, and a resource patch will be abandoned once the rate of return equals the rate obtainable in the surrounding environment (Charnov 1976). Optimal uses of a patchy environment are potentially different given a specific set of selective pressures (e.g., population size, mobility, level of exchange), and decreases in the overall abundance of food within an environment can make patches that were originally excluded more attractive to foragers (Bettinger 1991). Ethnographic and archaeological data suggest that maritime foragers, when compared to their terrestrially focused neighbors, tend to live in larger, more stable communities—often on or near the coast (Ames 2002; Erlandson 2001; Yesner 1980). Economically, one would expect maritime foragers to establish settlements in locations (patches) that provide the largest subsistence rewards (e.g., estuaries), and the archaeological record in many coastal areas generally supports this prediction (Erlandson 2001). In other words, maritime foragers tend to position themselves centrally and collect resources logistically (e.g., Ames 2002; Binford 1980, 2001); that is, smaller foraging parties venture out for short periods of time, capture and process prey, and return these food items to a central location. Because maritime foragers often collect resources logistically and leave large accumulations of debris at residential bases, central place foraging (CPF) theory is particularly promising for archaeologists working in coastal settings. CPF was first described by Orians and Pearson (1979) and has been used successfully in the biological sciences to explain foraging behavior in patchy environments (Stephens and Krebs 1986). Variants of this model have also been employed by ethnographers and archaeologists to analyze foraging behavior in a variety of ecological settings (Bettinger et al. 1997; Bird 1997; Bird and Bliege Bird 1997; Cannon 2003; Gremillion n.d.; Nagaoka 2002a). Optimal diet choice is constrained significantly when foraging from a central location due to costs associated with round-trip travel and natural limitations on load size. The resources exploited logistically must yield sufficient energy to offset these costs. Greater distances traveled to resource patches require larger packages of energy to be returned to the central place (figure 3). Therefore, longer trips require higher rewards than shorter trips and rewards are constrained by limitations on load size. Coastal habitats provide an assortment of protein-rich resources, but may not be positioned near plant communities that provide seasonally high rewards and necessary carbohydrates (Speth and Spielman 1983).

HBE AND MARITIME SOCIETIES

31

Return Rate (kcal) A=High B=Medium C=Low

Net Return (kcal)

A B

Minimum Return Needed

C

Distance from Central Place figure 3. Graphical representation of the central place foraging model showing the relationship between net return rate from foraging versus the distance traveled to a foraging area. As the forager expends more time traveling to the resource patch the net return rate declines (drafted by D. Kennett).

Maritime foragers with a wide diet breadth may also encounter scheduling conflicts between spatially separated coastal resources that become available at the same time. Division of labor (age and sex) could be used to solve these spatial incongruities (Jochim 1988; Kelly 1995; Meehan 1977; Moss 1993; Smith 1991; Zeanah 2000), but individuals would be forced to decide where to place residential communities, how often to move these settlements, and which resources to exploit residentially and logistically. The CPF model predicts that foragers will establish settlements in locations that maximize average central place foraging returns relative to transportation costs associated with logistical forays. Sets of resources will be exploited from central places that maximize foraging returns in light of pursuit, handling, and transport costs. Therefore, in the absence of highly productive terrestrial resource patches, maritime foragers would be expected to establish settlements in the most productive coastal locations. Logistical mobility would be used from these central locations to

32

HBE AND MARITIME SOCIETIES

exploit shellfish beds, near- and offshore fisheries, sea mammal rookeries, and productive terrestrial plant and animal communities. Transportation of bulky plant foods over land would be relatively expensive, and load sizes would be limited by the size of burden baskets (25–26 L or 36 kg; Barlow et al. 1993; Bettinger et al. 1997). Boats reduce pursuit and transportation costs significantly in maritime settings and can carry loads up to 2,000 kg (Ames 2002). Therefore, even rudimentary boats would promote settlement in coastal areas. In these settings, residential mobility may be used to capitalize on seasonal foraging opportunities in highly productive and distant resource patches (e.g., sea mammal rookeries or seed-bearing trees). Experimental work on the energetics of resource procurement and transportation suggests that logistical procurement involving food processing is generally more costly than residential procurement (Bettinger et al. 1997). This is particularly the case in terrestrial ecosystems for which transportation costs are high and load size is limited (Barlow and Metcalfe 1996; Barlow et al. 1993; Jones and Madsen 1989). Under conditions of low population density and unlimited access to spatially separated resources, foragers may make several residential moves annually rather than exploit resources logistically, even if these resources were profitable after processing and transport (Bettinger et al. 1997). Residential movement in coastal settings may occur when seasonally available plant foods in the interior are in need of harvesting. In coastal California, residential moves into the interior to collect acorns are well documented (Landberg 1965). However, under conditions of high population density, it is likely that more permanent settlements would be established along the coast and seasonally productive plant resources would be collected logistically or acquired via exchange. Decreased residential mobility also promotes more intensified use of localized resources, resource depletion, and increases in diet breadth. These temporal changes are best addressed within the framework of the ideal free distribution.

Intensification and the Ideal Free Distribution Archaeological data from a variety of coastal settings around the world suggest intensified use of maritime resources during the past 10,000 years (Bailey and Craighead 2003; Erlandson 2001; Kennett and Kennett 2000; Kennett and Voorhies 1996; Spriggs 1997). This is partially related to the stabilization of sea level and the development of productive coastal habi-

HBE AND MARITIME SOCIETIES

33

tats between ~10,000 and 7,000 BP (e.g., estuaries in some regions). Intensified use of coastal resources is also coincident with climatic changes during the Pleistocene to Holocene transition, the extinction of many large animal species in the New and Old World at this time (Piperno and Pearsall 1998), and the diversification of foraging and farming strategies worldwide (broad-spectrum revolution). The large number of coastal sites dating after 7,000 years is also related to the inundation and destruction of coastal areas because of earlier marine transgression. However, the development of more sophisticated fishing technologies and the intensified use of maritime resources, particularly fish, also has occurred in some parts of the world within the past 7,000 years, after major climatic changes associated with the Pleistocene–Holocene transition and the stabilization of sea level. Resource intensification is a process whereby the total productivity of a resource patch is increased but at the expense of decreases in net foraging returns (Boserup 1965; Broughton 1999). Intensification models are often density dependent in that they predict that foraging efficiency decreases through time as populations increase. The mechanism for change is resource depression caused by increased predation pressure. Reduced foraging returns from easy prey (low investment–high yield) stimulate increases in diet breadth and an emphasis on highercost prey items. These types of models have been used to help explain the appearance of more intensive economic strategies during the Late Holocene (Basgall 1987; Beaton 1991; Bettinger 1991; Broughton 1994a; Butler 2000; Grayson 1991; Janetski 1997; Piperno and Pearsall 1998; Porcasi et al. 2000). There is also a spatial component to resource intensification. If resource zones vary with respect to overall productivity, people will tend to aggregate near those patches that provide the highest overall return rates (packing; Binford 1968, 1983). However, the overall productivity and suitability of these resource patches can change, in turn stimulating movement of some individuals into adjacent areas or more intensified use of the same patch of land. Short- and long-term climatic changes may alter the distribution and availability of resources within the patch, as would changing subsistence practices and technology (e.g., foraging or farming), fluctuations in the density of human occupants, and habitat degradation resulting from unsustainable exploitation. Dense populations often deplete resources rapidly, but low-density use can also affect the availability and distribution of other plant and animal species on which a population depends (Sutherland 1996). However, mutualistic

34

HBE AND MARITIME SOCIETIES

and beneficial relationships between humans, animals, and plants may also increase subsistence rewards within a resource patch (Rindos 1984). Another consequence of large populations is interference with the subsistence pursuits of others that impacts the overall suitability of a resource patch or habitat. Interference can result from fighting, stealing, or control of resources or patches by individuals (Sutherland 1996). Territorial or despotic behavior by individuals or groups also affects the suitability of a resource patch and stimulates the movement of people into adjacent, less desirable resource zones. In larger-scale human societies (e.g., chiefdoms or states), other social and political factors may stimulate the use of more marginal resource patches. These factors include the control of large areas by a small number of elite individuals, unreasonable taxation or tribute systems, and the persistence, or threat, of warfare. The ideal free distribution model (IFD) provides a potential explanatory framework for predicting when individuals will move to less desirable resource patches based on density-dependent changes in the suitability of the habitats available to them (figure 4). In this model, resource patches are ranked by their quality, as assessed by the fitness of an occupant or a related measure of suitability (see Kennett et al. n.d.a; Winterhalder and Kennett n.d.). In the IFD habitat quality is density dependent and declines with increasing population density through the effects of interference or scramble competition. One of the assumptions of the model is that individuals will elect to reside in the best habitat available to them (the ideal of the model name) and that they are unrestricted in their movement to effect that choice (the free of the name). In effect, they are competitors of equal ability and access to resources. Under these conditions, individuals will settle first in the best habitat available. With increasing density due to immigration or in situ growth, suitability within the patch drops. When it is diminished to the level of the second-ranked resource patch, further population growth will be divided between them. Decreased suitability could be related to interference caused by other individuals, depletion of resources, or other density-dependent variables that negatively impact the patch and its overall suitability. In coastal settings this could include density-dependent decreases of larger-bodied sea mammals or depletion of local shellfish beds due to intensive stripping. The model predicts that once the average payoff for individuals in the best patch is equal to that for the second-best patch (figure 4, Point A), then populations will continue to grow slowly in the best patch as individuals now begin occupying the second-best patch. The equilibrium distribution is a consequence of the

HBE AND MARITIME SOCIETIES

+

H1

35

Best Habitat

H3



SUITABILITY

H2

Worst Habitat A

0



B

POPULATION DENSITY

+

figure 4. The ideal free distribution model (based on data from Sutherland 1996; drafted by D. Kennett).

marginal equalization of habitat suitability: When this is the case, no individual has any incentive to relocate. The ideal despotic distribution is a variant of the IFD highlighting differential access to resources. If interference arises among competitors of unequal abilities, or if by establishing territories initial or superior competitors can protect themselves from density-dependent habitat deterioration by defending better resource opportunities, then the inferior competitors and those without territories are pushed to poorer habitats. Compared to the IFD, a despotic distribution will equilibrate with disproportionate numbers or densities in the lower-ranked habitats. This makes intuitive sense: By garnering disproportionate resources in the best habitats, the better competitors push the inferior competitors, sooner, into habitats of lesser suitability. Because of this, the use of lower-ranked resource patches has been documented as a buffering strategy—at least among a variety of bird species (Brown 1969; Meire and Kuyken 1984; Moser 1988). In fact, in many empirical studies the ideal free distribution serves as a null hypothesis in order to measure the effects of interference competition and unequal resource access (Sutherland 1996). Whatever form it takes, the IFD shows how an incremental quantitative change in one variable (e.g., population density) may lead to

36

HBE AND MARITIME SOCIETIES

qualitative changes in another (e.g., the range of habitats occupied; Kennett et al. n.d.). Moving from the IFD to the despotic variant shows the qualitative changes of increasing magnitude. As with most HBE models, there are few limits on what kinds of variables one might accommodate in the IFD. For instance, climate change might shift the relative suitability (vertical position, thus relative ranking) of the curves. Habitats or subsistence practices highly susceptible to density-dependent degradation will have steep downward slopes; those that generally are not so sensitive to population density will have more shallow slopes. As noted, economies of scale in subsistence practice may cause the slope of the curve to be positive. The consequences of social inequality and economic exploitation for dispersion and habitat use can be represented in despotic versions. Some of the best resource patches along coastal stretches are near estuaries where the mixing of fresh water and salt water produce rich ecological systems and fresh drinking water is also close at hand and easily transported by boat. Habitats supporting large sea mammal colonies, kelp forests, or extensive shellbeds would also be ideal settings for early settlements. The IFD model predicts that the earliest settlements would be located in these areas. Depletion of larger sea mammals, fish, or easily accessible shellfish would stimulate increases in diet breadth and the expansion of populations into adjacent locations, if the payoffs of moving into these areas are equivalent to the original resource patch. Resource depletion in coastal settings could also be caused by increases in local population densities or environmental change (e.g., infilling of estuaries). Once the available resource patches are inhabited, population growth will depend on the ecological sustainability of the resource patch. In some cases, more intensive use of certain maritime resources may result, with fish being one of the prime candidates as an intensifiable resource.

Competition and the Formation of Social Hierarchies in Coastal Settings Group sizes and population densities along coasts tend to be larger than adjacent interior areas because of the productivity of marine ecosystems, their juxtaposition with terrestrial habitats which increases the diversity of resources available, and the linear distribution of subsistence resources (Erlandson 2001; Yesner 1980). Environmental or social circumscription also appear to be correlated with settlement packing

HBE AND MARITIME SOCIETIES

37

and high population densities in many coastal settings (Binford 1968; Moseley 1975). Economic, social, and political differentiation of individuals in large coastal communities is common and the interconnectedness of individuals in and between communities is often well developed as it is in other complex societies (Kantner 2002). Complex decision-making hierarchies and permanent leadership often accompany increased differentiation between individuals. This can be manifested in lineage ownership of rich shellfish beds or fishing grounds and the technology needed to effectively exploit these resource patches (boats, weirs, etc.). Leaders often fulfill vital roles in redistributing resources, organizing communities to defend territories, and brokering exchange relationships/alliances with other communities. Statistical analysis of modern societies from The Ethnographic Atlas (Murdock 1967) show a positive correlation between coastal resource use and a variety of measures for social complexity (e.g., hierarchy, territoriality, stratification; Pálsson 1988). Complex group formations should not be surprising in coastal contexts because successful foraging expeditions (e.g., fishing or marine mammal exploitation) are often dependent upon cooperation between several members of a group and large catches must be effectively distributed through social networks. Manufacture, use, and maintenance of boats also often require a coordinated group effort that must be sustained for long periods of time. The archaeological record in coastal California, and other maritime settings (Yesner 1980), suggests increased population density and the formation of social hierarchies during the Late Holocene (Arnold 2001; Erlandson and Jones 2002; Kennett and Kennett 2000; Raab and Larson 1997). Groups of people became both larger and more sedentary and more differentially distributed along the coast. In these and other contexts, the size and structure of groups is mediated by selfinterested individuals trying to maximize access to limited resources (Boone 1992). The HBE approach espoused here predicts that the formation of communities will result from individuals weighing the costs and benefits of living in a large group compared with other alternatives. The benefits of living in a group would include access to mates, more efficient harvesting of resources (economies of scale), and defense and control of resources. In coastal settings this may include the protection of intertidal resources, nearshore fishing grounds, or sea mammal colonies (see Thomas 2001). Costs may include interference and competition between individuals for land/food and ultimately the degradation of local ecosystems. In general, one would predict that individuals

38

HBE AND MARITIME SOCIETIES

will live in groups as long as the net rewards from doing so are greater or equal to the rewards of living alone or in smaller, more dispersed groups. The development of inequality involves the exploitation of some individuals within a group and this phenomenon is particularly interesting given the assumption that people will act relatively rationally and with self-interested motivations. High-ranking individuals clearly receive more tangible social, political, and economic advantages in hierarchical societies, but the average per capita rewards may be well above the gains acquired if individuals lived apart from the group (Arnold 2001). An individual’s ability to leave a group also depends on both the availability and suitability of other habitats and the competitive and cooperative actions of other people in the region. Individuals may live in large groups, even under severely disadvantaged conditions, if population densities are high and unoccupied territories are lacking (Boone 1992). Within this context some individuals or lineages may seize control of locally available resources and subjugate other members of the group. In this situation, the disadvantaged individuals are still self-interested because being subjugated was the best choice given a limited range of possibilities (Boone 1992). Social hierarchies are therefore density dependent and at least partially related to resource structure and territorial behavior. Dyson-Hudson and Smith (1978) argued that strong forms of territorial defense are favored in habitats that contain dense resources predictable in time and space. In these contexts, the benefits of controlling access to resources exceed the costs of defense and are therefore economically defendable. Many coastal environments contain densely packed resources that are available throughout the year and thus economically defendable. The presence or absence of other people makes certain resource patches (clumps) more or less attractive, and individuals entering an environment that is occupied by other people may choose a less desirable place to settle because of overcrowding at prime locations. The ideal free distribution model presented in the previous section predicts that resource patches providing the highest rewards will be occupied first. Population-dependent degradation of the best resource patches and interference through scramble competition will result in the occupation or colonization of adjacent habitats. In coastal settings, once the most economically defendable locations are occupied a more despotic distribution is predicted. Therefore, individual fitness will vary and the population will be unevenly distributed along the coastline.

HBE AND MARITIME SOCIETIES

39

Economically defendable resources are also the greatest source of competitive aggression, and escalating offensive and defensive strategies may develop if individuals perceive net economic gains (Boone 1992). Under these conditions, larger groups may be favored to acquire, maintain, and defend access to resources. This provides a potential mechanism for the formation of social hierarchies. As group size increases through in-migration or inherent population growth, competition for resources forces some people to accept more marginal access to resources within the vicinity of the settlement. In coastal settings this may involve greater travel costs to fishing grounds or intertidal habitats that cause overall decreases in net rates of return. The model predicts that groups will fission when local rates of return are equal to, or lesser than, net return rates in adjacent unoccupied territories. Social hierarchies would not be expected to develop if viable opportunities to relocate existed. Emigration may not be economically feasible under highly circumscribed conditions (environmental or social; Carneiro 1970, 1978, 1988), caused by population growth and environmental infilling, and it is under these conditions that hierarchical social organization should begin to develop (Boone 1992, 317).

Summary HBE provides a robust framework for exploring prehistoric human behavior (Kelly 1995; Winterhalder and Smith 1992). Theoretically, the approach is well grounded in neo-Darwinian theory, relying heavily upon the principles of variation, optimization, and selection. The approach is dissimilar to selection-based evolutionary archaeology (Dunnell 1989; Teltser 1995a) in its emphasis on ecological context; that is, natural selection cannot be separated from a bounded set of changing selection pressures imposed by environmental, historical, and social contexts. Human biological and cultural evolution are extremely complex, and certain aspects of the evolutionary process are beyond the current methods of analysis (Bettinger and Richerson 1996; Jochim 1981). Taking a Darwinian stance requires an acute sense of variability on many different timescales. The data requirements necessary to document such variability in the archaeological record are immense. A thorough understanding of the natural environment is necessary, particularly the spatial and temporal distributions of subsistence resources. Detailed

40

HBE AND MARITIME SOCIETIES

paleoclimatic reconstructions for a region can be used to make inferences regarding shifts in the spatial and temporal distribution of resources and how these changes would affect selection pressures acting on individuals. Historical context is also crucial. When did people first colonize an area? What are the general demographic trends for the region? What are the key historical events (i.e., technological innovations) that could have changed the ecological context? In the following chapter, I discuss the temporal and spatial distribution of resources on the northern Channel Islands of California. Paleoenvironmental reconstructions are used to make inferences about changes in the natural environment that would have altered the selective pressures operating on the prehistoric human inhabitants of the northern Channel Islands. Chapter 3 is followed by a general discussion of the prehistory of the Santa Barbara Channel region (chapter 4). The primary objective of these two chapters (chapters 3 and 4) is to provide the ecological context for the economic, social, and political developments visible in the archaeological record on the northern Channel Islands through time.

chapter 3

Environmental Context

Exploring the prehistory of any region using an HBE approach requires a detailed understanding of the natural environment. This chapter describes the physical and biotic character of the northern Channel Islands, emphasizing the spatial and temporal distribution of subsistence and nonsubsistence resources. As in most of coastal California, the marine environment surrounding these islands provided rich and varied subsistence resources to the prehistoric people who once occupied them. In contrast to the marine system, the terrestrial environments on these islands are generally depauperate. However, terrestrial habitats did provide a range of subsistence and nonsubsistence resources that were crucial to prehistoric island life. Many archaeologists working in the Santa Barbara region treat the northern Channel Islands as one ecological unit, usually relative to the mainland and interior parts of “Chumash territory” (e.g., King 1976; Walker and DeNiro 1986). The primary aim of this chapter is to show that significant physiographic and environmental differences exist between and within each of the northern Channel Islands and that subsistence resources, although rich, are patchy and often unpredictable. The uneven distribution of terrestrial and marine resources is also exacerbated by seasonal, interannual, decadal, and longer-term climatic variability. Individual behavioral strategies that emerged and became dominant during the Holocene were certainly mediated, in part, by this changing set of ecological opportunities and constraints.

41

42

E N V I R O N M E N TA L C O N T E X T

General Physiography The northern Channel Islands, a continuation of the Santa Monica Mountains, extend east–west for approximately 88 km (55 miles) along the southern margin of the Santa Barbara Channel. Four islands— Anacapa (2.9 km2), Santa Cruz (249 km2), Santa Rosa (217 km2), and San Miguel (37 km2)—comprise this island chain (map 2). These islands are located between 32 and 44 km (20 and 44 miles) offshore and were never connected to the mainland, even during the low–sea level stands of the Pleistocene epoch (Junger and Johnson 1980; Wenner and Johnson 1980). Anacapa is the smallest of these islands, consisting of three small islets and a shear, rocky coastline, making it relatively inaccessible (figure 5A). Santa Cruz is the largest and most topographically rugged of the islands, with three mountain ranges and associated ridge systems dominating the landscape, the highest peaks approaching 800 m (figure 5B). Two of these ranges, oriented east–west, are separated by a distinctive central valley, and a third, El Montañon, runs from north to south, separating the eastern end of the island from the more rugged terrain that dominates the west. Compared with Santa Cruz, the outer islands of Santa Rosa and San Miguel are less rugged and topographically variable. A small, central mountain range runs east–west through Santa Rosa, and it has peaks reaching only 450 m (Soledad Peak; figure 5C). The northern part of Santa Rosa Island is dominated by a series of relatively flat, uplifted Quaternary terraces, incised by streams flowing out of the central range. Terrain on the south side of the island is more precipitous compared to the north coast, and the distance between the coast and the central range is compressed. San Miguel is the least mountainous of the northern Channel Islands (figure 5D). Dune fields dominate the landscape and two small hills, Green Mountain and San Miguel Peak, form a subtle boundary between the northern and southern sides of the island.

Geology Landscape formation on the northern Channel Islands is strongly influenced by the rich and diverse geologic history of the region (Atwater 1998; Dibblee and Ehrenspeck 1998, 2000; Palmer 1997; Weaver 1969; Weigand 1998). The chemical composition and erosional qualities of

SAN

TA BARBARA CHANNEL

Harris Pt.

Profile Pt. Frasier Pt.

1 Pt. Bennett

2

San Miguel Island

Cardwell Pt.

Brockway Pt.

Bechers Bay

Sandy Pt.

Diablo Pt.

Cavern Pt.

Carrington Pt. Christy Beach 1

San Pedro Pt.

4 3 2

4

Skunk Pt. Santa Rosa Is. Fault

Mountain Peaks 1 Green Mountain (817 ft)

2 San Miguel Hill (831 ft) 3 Soledad Peak (1,574 ft) 4 Diablo Peak (2,434 ft)

Anacapa Island Morse Pt.

3

Punta Arena

Santa Cruz Island Fault (Central Valley)

El Montanon

East Pt.

Santa Cruz Island

Cluster Pt. South Pt.

Santa Rosa Island

Meteorological Stations 1

Christy Ranch

2 Stanton Ranch 3 Prisoners Harbor 4

Navy Base Fault

N

0

10

20 Kilometers

W

E S

map 2. Topographic map of the northern Channel Islands showing major landmarks discussed in text (map produced by D. Kennett with the assistance of Jacob Bartruff).

44

E N V I R O N M E N TA L C O N T E X T

A

B

C

D

figure 5. Photographs of the northern Channel Islands. (A) Eastern Anacapa Island looking west (photo by D. Kennett); (B) eastern Santa Cruz Island looking west (photo by D. Kennett); (C) northern Santa Rosa Island taken from Soledad Peak (photo by S. Spaulding); (D) western San Miguel Island looking west toward Point Bennett (photo by S. Spaulding).

different substrates and rock types provide the basis for soil formation and ultimately the distribution of terrestrial plant and animal communities on these islands (Butterworth et al. 1993; Jones and Grice 1993; Junak et al. 1995). Uplift, faulting, and substrate permeability generate topographic and hydrologic differences on all the islands. Certain geologic formations were also of great importance in providing raw materials critical for prehistoric stone tool production (Arnold 1987; Arnold et al. 2001; Conlee 2000; Erlandson et al. 1997; Kennett 1998; Kennett and Conlee 2002; Walker and Snethkamp 1984). Weaver (1969) mapped 18 major geologic substrates on Santa Cruz Island ranging from pre-Tertiary bedrock formations to more recently uplifted Quaternary terrace deposits. A fault, running east–west through the central valley, forms a major geologic boundary on the island (see map 2; Atwater 1998; Boles 1997; Pinter et al. 1998; Sorlien and Pinter 1997). Mid-Tertiary volcanic and volcaniclastic formations—overlain by Monterey Shale and Quaternary terrace deposits—dominate the landscape north of the fault (Weaver 1969; Weigand 1997). These mid-

E N V I R O N M E N TA L C O N T E X T

45

Tertiary volcanic rocks are similar in composition to the volcanic formations that predominate on Anacapa Island (Weigand 1993, 1997; Weigand and Savage 2000). A variety of Tertiary sedimentary and metamorphic substrates exist south of the fault, along with ancient (pre-Tertiary) metamorphic and plutonic basement rock (Gordon and Weigand 1997; Molesworth and Sloan 1998; Shapiro 1998). On the eastern end of the island, rhythmically bedded diatomaceous, tuffaceous, and siliceous Miocene shales and cherts (Monterey Formation) overlie the mid-Tertiary volcanic formations that predominate on the northern side of the island (Weaver and Meyer 1969). These cherts were used prehistorically to make stone tools (Arnold 1985, 1990a; Kennett 1998). The Tertiary sedimentary and metamorphic stratigraphic sequences that flank the south side of Santa Cruz continue on the north and south side of the central range on Santa Rosa (Dibblee and Ehrenspeck 1998, 2000). Stratigraphic sequences ranging in age from the Eocene to the Middle Miocene are present and exposed on various parts of the island (Jennings 1977). The Santa Rosa Island fault forms a natural boundary between the younger Miocene sediments to the north and older Miocene formations and volcanics to the south (see map 2; Dibblee and Ehrenspeck 2000; Weaver 1969). Deeply incised drainages on the south side expose Tertiary sandstone and shale as well as midTertiary clastics and volcaniclastics with volcanic intrusives. These include the Sespe, Vaqueros, Rincon, and Monterey formations (Avila and Weaver 1969; Chinn and Weigand 1994). Upper Miocene intrusive basalts are also well exposed in the south-central portion of the island, northeast of Soledad Mountain. Uplifted Quaternary marine and nonmarine terrace deposits occur north of the Santa Rosa fault and are well known for containing pygmy mammoth remains (Mammuthus exilis), a small elephant species that became extinct at the end of the Pleistocene (Agenbroad 1998, 2000a, 2000b; Orr 1968; Thaler 1998; Woolley 1998). Quaternary terraces predominate on San Miguel and co-occur with stabilized and active dune deposits. Volcanic and Monterey sedimentary formations, including chalcedonic cherts, are exposed in small areas on the eastern side of the island near Cardwell Point (Erlandson et al. 1997). Pleistocene eolianite sandstones on the northwest side of the island have produced some mammoth bones (Gray and Harz 1998). Large dune fields flank the north side of the island, oriented northwest with the prevailing wind that persistently sweeps the island. These dune

46

E N V I R O N M E N TA L C O N T E X T

fields are superimposed on massive Quaternary caliche deposits, marine limestones, and beach rock (Johnson 1972). The volcanic and Monterey formations on these islands provided raw materials for a range of prehistoric flaked and groundstone tool industries during the Holocene (Arnold 1987; Arnold et al. 2001; Conlee 2000; Erlandson et al. 1997; Kennett and Conlee 2002; Walker and Snethkamp 1984). Volcanic formations are found on all of the islands but are most extensive on Anacapa and Santa Cruz. Crude cores, flakes, and cobble tools, of Holocene age, made from island volcanics are common in prehistoric deposits on all of the islands. Volcanic porphyries exposed on the northwest coast of San Miguel Island were used intensively after 1,300 BP to produce stone mortars and pestles (Conlee 2000; Kennett and Conlee 2002; Walker and Snethkamp 1984). Monterey chert formations on eastern Santa Cruz Island produced some of the most valuable flake stone material on the northern Channel Islands. Chert is exposed west of El Montañon along a geologic contact between the volcanic and Monterey formations and in small pockets across the eastern end of the island. This chert was used intensively during the Late Holocene to produce small trapezoidal and triangular-shaped microblades (Arnold 1985, 1987, 1990a; Arnold et al. 2001; Kennett 1998), an important component of a bead manufacturing industry that developed on the islands during that time. Chalcedonic chert is also present on the eastern side of San Miguel Island near Cardwell Point and was used well into the Late Holocene (Erlandson et al. 1997). Poor quality chalcedonic chert also occurs on various parts of Santa Rosa Island.

Climate The Santa Barbara Channel region has a Mediterranean climate marked by cool, wet winters and warm, dry summers (Smith 1952). Due to the ameliorating effect of maritime conditions, summers and winters along the Santa Barbara coast and offshore islands are generally mild relative to more extreme climatic conditions in the interior of California. In general, January through March are the wettest and coolest months of the year, with rainstorms blowing across the Southern California Bight. Precipitation from these storms is highly variable regionally. As frontal activity dies down in April and May, strong northwest winds blow in the Santa Barbara Channel impacting areas west of Santa Cruz (Fagan

E N V I R O N M E N TA L C O N T E X T

47

table 7. Average monthly air temperature and precipitation measured at Stanton Ranch Precipitation Month January February March April May June July August September October November December Mean

Air Temperature

mm

in.

C

F

113.03 114.3 84.84 31.75 7.87 0.09 0.25 1.01 7.37 13.46 40.89 85.34 500.6

4.45 4.5 3.34 1.25 0.31 2.28 0.01 0.04 0.29 0.53 1.61 3.36 19.7

11.9 12.9 13.2 14.6 16.1 17.5 19.7 20.9 19.3 17.2 14.3 11.7 15.8

53.4 55.2 55.7 58.2 60.9 63.5 67.5 69.7 66.8 63.8 57.8 53.1 60.4

Monthly precipitation measured at Stanton Ranch between 1904 and 1993. Air temperature taken between 1961 and 1971. source: Junak et al. 1995.

1993). Storm fronts are rare between June and September, and there is little rain during this season. Coastal sea fog is common during summer months (mid-May to July) but dissipates starting in October when storm fronts again begin to reach Southern California. Historic climate records for Santa Cruz indicate that variations in topographic relief, exposure to wind, and coastal sea fog have a large effect on local climate (Junak et al. 1995). The largest air temperature variations on the island are in the central valley because it is sheltered from the ameliorating effects of wind and coastal sea fog (Junak et al. 1995). Average air temperatures recorded at Stanton Ranch between 1904 and 1993 ranged from 11.7°C (December) to 20.9°C (August), and precipitation for the same interval ranged from 0.09 mm (June) to 113.03 mm (January) (table 7). Annual precipitation levels at four different locations on Santa Cruz (measured between 1972 and 1982) indicate that precipitation is highly varied spatially (table 8), with the lowest averages occurring on the western end of the island at Christy Ranch. Comprehensive meteorological data are not available for Anacapa and Santa Rosa islands. Rainfall averages on Anacapa probably approximate those collected at Prisoners Harbor on eastern (central) Santa Cruz, whereas averages on Santa Rosa are more probably similar to

48

E N V I R O N M E N TA L C O N T E X T

table 8. Mean annual precipitation measured in mm (inches) at four stations on Santa Cruz Island from 1972 to 1982 (see map 2 for locations) Year

Stanton Ranch

Christy Ranch

1972–73 1973–74 1974–75 1975–76 1976–77 1977–78 1978–79 1979–80 1980–81 1981–82 Mean SD

537 423 380 174 328 1,131 645 648 408 426 508 256

297 279 265 124 248 732 347 342 288 289 321 157

(21.1) (16.6) (15.0) (6.8) (12.9) (43.8) (25.4) (25.5) (16.1) (16.8) (20.0) (10.1)

(11.7) (11.0) (10.4) (4.9) (9.8) (28.8) (13.7) (13.5) (11.3) (11.4) (12.6) (6.2)

Navy Base 462 451 327 158 323 1,034 559 495 340 355 450 234

(18.2) (17.7) (12.9) (6.2) (12.7) (40.7) (22.0) (19.5) (13.4) (14.0) (17.7) (9.2)

Prisoners Harbor 584 330 378 171 295 763 447 624 337 305 423 181

(23.0) (13.0) (14.9) (6.7) (11.6) (30.0) (17.6) (24.6) (13.3) (12.0) (16.7) (7.1)

source: Junak et al. 1995.

those on western Santa Cruz. San Miguel, the most seaward of the islands, has lower overall air temperatures when compared to Santa Cruz (11.9°C–15.5°C), and these temperatures exhibit less variation (Huckins 1995). Climatic conditions on San Miguel and western Santa Rosa are also heavily influenced by strong northwesterly winds and coastal sea fog. Northwesterly winds are strongest between May and September and have a drying effect on the outer islands that promotes dune formation. Coastal sea fog blankets these islands during spring and summer months and can contribute measurable amounts of precipitation.

Hydrology Compared with the mainland coast, fresh drinking water is restricted on the northern Channel Islands, particularly on the smaller islands of Anacapa and San Miguel. Map 3 shows the 44 largest watersheds on the northern Channel Islands. Because of its size, Santa Cruz has the greatest number of drainages; the largest and most reliable being Cañada de Puerto and Cañada Christy (Bremner 1932). Cañada de Puerto flows into Prisoners Harbor on the north coast of the island, and Cañada Christy drains east–west through the central valley of the island and enters the ocean on the western end of the island near Christy Beach. Fewer drainages exist on the southern and western sides of the island, but they tend to be larger than those on the north coast. Many of these

E N V I R O N M E N TA L C O N T E X T

49

streams are intermittent, flowing only during the wettest months of the year. Santa Rosa Island has several streams and springs that provide reliable water throughout the year. The largest and most reliable creeks are found on the north coast where drainages are long and watersheds are laterally extensive (figure 6A). Of these, Lobo Canyon, Cañada Verde, Arlington Canyon, and Cañada Tecolote provide perennial streams that continue to flow even during severe drought years. Perennial streams also flow in Old Ranch, Water, and Ranch canyons on the eastern side, but only a few reliable streams exist on the south and southwestern sides of the island. There are no substantial drainages on Anacapa, but a few small springs exist on the largest of the three islets. More drinking water is available on San Miguel, where the three most substantial drainages provide at least a trickle of water throughout the year. Verdant vegetation along the north coast of the island is supported by water seeps, and larger springs and creeks show a noticeable flow at contacts between dune deposits and underlying and impermeable caliche beach formations (Johnson 1972; Power 1979). Toward the western end of the island, several of these springs are quite vigorous and continue to flow during dry years. Water availability on the island also appears to be tied to the amount of vegetation on the island, with additional springs appearing when vegetation is most abundant. Drinking water is currently limited on the south side of San Miguel.

Terrestrial Resources: Spatial Distribution A variety of terrestrial microenvironments exist on each of the northern Channel Islands, and the distribution of economically valuable plants and animals is variable across space. Pygmy mammoths (M. exilis) were present on these islands prior to 12,000 years ago (Agenbroad 1998, 2000a, 2000b; Orr 1968; Roth 1993), but direct evidence that these animals were hunted does not exist (Erlandson 1994). Tantalizing new data are closing the temporal gap between the extinction of these mammoths and the appearance of humans between 13,000 and 12,000 years ago (Agenbroad 1998, 2000a, 2000b; Johnson et al. 2000). The largest land mammal available to islanders during the Holocene was a diminutive species of fox (Urocyon littoralis; Collins 1991a, 1991b; Fausett 1993), and it is possible that humans introduced these animals onto the islands sometime during the Holocene (Vellanoweth 1998). Given the

SAN

TA BARBARA CHANNEL 21

2

18 3

1

22 17

19

25 15 2

13

San Miguel Island

11 2 14

16

11

3

4 10

3

7 23

1 5

6

5 12 10

6 8

4

20 14 12

24 13 16

9

8

1 9 15

7

Santa Cruz Island

Santa Rosa Island

Rocky Headlands

N

0 W

E

10

20 Kilometers

Rocky Intertidal Sandy Beach

S

E N V I R O N M E N TA L C O N T E X T

51

absence of large game animals, various plant communities were certainly the most attractive terrestrial resources. Ethnohistorically, the Chumash people living in the Santa Barbara Channel region collected a wide variety of plants from different ecological zones (Timbrook 1990, 1993). Plants provided an important source of food and were also of high value for medicinal and ceremonial purposes (Timbrook 1980, 1982a, 1982b, 1986, 1987, 1990; Walker and Hudson 1993). Past distributions of plants across the northern Channel Islands is unknown because of the introduction of exotic grasses and historic alterations to the native landscape (Chess et al. 2000; Colvin and Gliessman 2000; Corry and McEachern 2000; Klinger and Messer 2000). Nevertheless, in spite of modern influences, establishment of the natural distribution of plant communities in the region is difficult because the Chumash continually altered the distribution and productivity of plant resources with periodic burning (Timbrook et al. 1982). Plant communities on the northern Channel Islands are less diverse compared with the mainland, with maritime influences tending to promote plant species more closely related to central California (Timbrook 1993). There are some island endemics, but fewer than half of the plant species available on the mainland occur on the islands (Philbrick 1980; Schoenher et al. 1999; Timbrook 1993). Philbrick and Haller (1977) described nine plant communities currently found on the northern Channel Islands: southern beach and dune, coastal bluff scrub, coastal map 3. Map showing major watersheds and variations in coastal habitat on the northern Channel Islands. Watersheds were tabulated using a digital elevation model (DEM) for the islands and the hydrologic modeling tools in Arc/Info 7.1. Watersheds were ranked based on the total surface area contributing to a particular drainage (map produced by D. Kennett with the assistance of J. Bartruff). Santa Cruz Island drainages: 1. Cañada de la Calera, 2. Cañada Christy, 3. Laguna Canyon, 4. Willows Canyon, 5. Pozo Canyon, 6. Cañada de los Sauces, 7. Scorpion Canyon, 8. Coches Prietos Canyon, 9. Smugglers Canyon, 10. Alamos Canyon, 11. Johnson’s Canyon, 12. Unnamed (Twin Harbors 1), 13. Montañon Canyon, 14. Orizaba Canyon, 15. Unnamed (Trident Cove), 16. Unnamed (China Harbor), 17. Hazard’s Canyon, 18. Unnamed (Profile Point), 19. Diablo Canyon, 20. Unnamed (Dick’s Cove), 21. Unnamed (Lady’s Harbor), 22. Unnamed (Ruby Rock), 23. Unnamed (Twin Harbors 2), 24. Cañada de la Calera 2, 25. Valdez Canyon; Santa Rosa Island drainages: 1. Old Ranch Canyon, 2. Cañada Soledad, 3. Water Canyon, 4. Cañada Tecolote, 5. Arlington Canyon, 6. Cañada Verde, 7. Jolla Vieja Canyon, 8. Ranch Canyon, 9. Wreck Canyon, 10. Unnamed Canyon (China Camp 1), 11. Dry Canyon, 12. Unnamed (China Camp 2), 13. Cañada Lobos, 14. Unnamed Canyon (Bee Rock), 15. San Augustine Canyon; San Miguel Island drainages: 1. Willow Canyon, 2. Otter Creek, 3. Nidever Canyon.

A

B

C

D

E

F

G

H

figure 6. Photographs of different environs on the northern Channel Islands: (A) Perennial stream at the mouth of Cañada Verde, north coast of Santa Rosa Island (photo by S. Spaulding); (B) grassland, chaparral, and scattered scrub oaks on the eastern end of Santa Cruz Island (photo by S. Spaulding); (C) grassland and coastal sage scrub habitats on northern Santa Rosa Island (photo by S. Spaulding); (D) Torrey Pine grove on eastern Santa Rosa Island looking across channel to Santa Cruz Island (photo by D. Kennett); (E) California sea lions on beach near Point Bennett, western San Miguel Island (photo by S. Spaulding); (F) offshore kelp-bed habitat along the northern side of Smugglers Harbor, eastern Santa Cruz Island (photo by S. Spaulding); (G) sandy beach and rocky intertidal habitats near Ford Point, southern Santa Rosa Island (photo by D. Kennett); (H) remnant estuary at the mouth of Old Ranch Canyon, eastern Santa Rosa Island (photo by S. Spaulding).

E N V I R O N M E N TA L C O N T E X T

53

sage scrub, island chaparral, valley and foothill grassland, oak woodland, southern riparian woodland, pine (coniferous) forest, and coastal marsh. Junak et al. (1995) defined five additional communities on Santa Cruz: coyote brush scrub, freshwater seeps and springs, vernal ponds, riparian herbaceous vegetation, and mule fat scrub. Of these plant communities, grasslands, coastal bluff scrub, coastal sage scrub, and oak/pine woodland provided the greatest resource potential for the prehistoric people of the islands. Island plant communities are patchy, and the abundance and types of flora vary between islands. Geologic and elevation differences coupled with groundwater availability and wind exposure create a variety of plant microhabitats (Junak et al. 1995). The larger islands of Santa Cruz and Santa Rosa support a much higher diversity of floral communities than the smaller islands. Southern coastal dune, coastal bluff scrub, coastal sage scrub, and native grass communities occur on San Miguel, interspersed between windswept sand dunes. Anacapa is covered with a mix of coastal sage scrub and coastal bluff associates, and small areas support native grasses (Philbrick and Haller 1977). Large stands of oak and pine are absent on both of these smaller islands. Santa Cruz exhibits the greatest floral diversity owing to its size and topographic relief (figure 6B; Jones et al. 1993; Junak et al. 1995). Grasslands are widespread on both the east and the west ends, and currently they are composed of introduced species, but seed-bearing native bunch grasses are becoming more common on the western part of the island (Cobb and Mertes 2000). Coastal bluff scrub predominates on seacliffs and coastal slopes, and coastal sage scrub dominates on dry, rocky south-facing slopes across the island. Chaparral, island woodland, and southern coastal oak woodland tend to favor north-facing slopes, and the sheltered central valley of the island supports particularly large stands of oak woodland (Quercus agrifolia) and bishop pine forest (Pinus muricata) (Junak et al. 1995; Minnich 1980; Ostoja and Klinger 2000; Walter and Taha 2000). Santa Rosa also supports a wide variety of plants potentially of great value to prehistoric people. As on San Miguel, wind is one of the primary variables shaping the distribution of plant communities and much of the island is covered with grassland, coastal sage scrub, and coyote brush scrub. Grasslands dominate the flat Quaternary terraces on the north side of the island, but microhabitats are present in well-watered areas sheltered from the prevailing northwesterly wind (figure 6C). Coastal sage scrub grows on south- and east-facing slopes of drainages

54

E N V I R O N M E N TA L C O N T E X T

and even in the lee of small hills on the north side of the island. Pines (P. muricata) and oaks (Q. agrifolia) are also common on Santa Rosa but are restricted to larger, more sheltered drainages. A small stand of torrey pines (P. torreyana subsp. insularis)—representing one of only two isolated stands of these pine trees in California (San Diego and Santa Rosa)—is present on the northeastern end of the island (figure 6D). Pollen from a sediment core taken nearby indicates their presence on the island at least during the Middle and Late Holocene (Cole and Lui 1994). This species of pine has distinctively long needles and large edible nuts (Biondi et al. 1997; Cole and Lui 1994; Cole and Wahl 1996, 2000). Acorns and wild cherry, as well as a wide variety of other fruits, small seeds, and bulbs, were available to people on the islands, particularly on Santa Cruz and Santa Rosa. Descriptions of plant use by the islanders are generally underrepresented in ethnohistorical descriptions, and some accounts suggest that plant foods were less important to island populations. However, Timbrook (1993) has argued that the use of plants by islanders during the contact period was similar to plant use by mainland people. The depauperate flora was limiting and forced islanders to collect locally available plant foods and subsidize shortfalls by importing plants from the mainland. Important plants included islay, manzanita, mangle berries, tarweed seeds, and tubers (blue dick bulbs, brodiaea, Dichelostemma capitatum) (Timbrook 1993, 51). Chia sage (Salvia columbariae), of much importance to mainland populations because of its high protein content (Timbrook 1986), was also present on Santa Cruz and Santa Rosa. Periodic burning of sagebrush communities on the mainland increased the distribution and productivity of chia and this likely occurred on the islands as well (Timbrook 1993, 51; Timbrook et al. 1982).

Marine Resources: Spatial Distribution The marine environment surrounding the northern Channel Islands is highly productive, although the spatial composition and productivity of marine plants and animals is not uniform. Offshore habitats associated with each island vary because of their geographic position along the coast and marine productivity linked to water temperature variations and regional ocean circulation (Engle 1993, 1994). Present-day watermass circulation in this region is complex, and a number of small- and large-scale phenomena alter marine productivity. The region is influenced by two major current systems (figure 7): the southward-flowing

E N V I R O N M E N TA L C O N T E X T

55

figure 7. Major current systems along the Southern California coast (drafted by R. van Rossaman).

California Current (cold, low salinity) and the northward-moving California Countercurrent (warm, saline) (Browne 1994; Hickey 1992). The cold California Current and the associated Southern California Eddy are the dominant influence along the Southern California coast, including the Santa Barbara Channel (Hickey 1992; Wickham 1975). The northward-flowing California Countercurrent transports warm equatorial water along the Southern California coast and strongly interfaces with the California Current and California Eddy south of Point Conception and in the Santa Barbara Channel (see map 2 for location). One branch travels north between the Santa Barbara mainland and the northern Channel Islands, and the other flows south of the islands. In the Santa Barbara Channel, the countercurrent dominates during summer and early fall. Localized upwelling of nutrient-rich Pacific intermediate waters along the California coast during spring and summer is driven by temperature differences between air masses over water and land (Bakun 1990; Dorman and Palmer 1981). Strong northwesterly winds are generated during these months as air masses in western North America heat up relative to the air over the Pacific. These strong winds blow

56

E N V I R O N M E N TA L C O N T E X T

figure 8. Average monthly sea-surface temperatures surrounding Anacapa, Santa Cruz, Santa Rosa, and San Miguel islands between 1982 and 1991. Notice the ~2°C temperature difference between Anacapa and San Miguel islands (drafted by D. Kennett; data collected by J. Engle).

nearshore waters offshore and are replaced by cold, nutrient-rich Pacific intermediate waters. Localized upwelling is most intense adjacent to the mainland and along the south coasts of the northern Channel Islands. Upwelling close to the coast increases nutrients and decreases surface-water temperatures in portions of the California Current to as low as 8°C. (Bernal and McGowan 1981; Brink et al. 1983; Hickey 1992; Huyer 1983; Mooers and Robinson 1984; O’Brian 1983). The influence of the cold California Current, coupled with seasonal upwelling, provides the foundation for high marine productivity and the rich nearshore fishery in the region. A surface-water temperature gradient of ~2°C (warm to cool) occurs between Anacapa, the farthest island east, and the westernmost island of San Miguel (figure 8). This directly affects the distribution and composition of algae, kelp beds, benthic biota, shellfish, fish, and sea mammals around each island (Engle 1993, 1994; Murray et al. 1980; Neushul et al. 1967; Seapy and Littler 1980; Stewart et al. 1993). Warm-water fish are common in the waters surrounding Anacapa and Santa Cruz, whereas cold-water fish are favored

E N V I R O N M E N TA L C O N T E X T

57

around the outer islands of Santa Rosa and San Miguel (Engle 1993). Warm-water pelagic fish (e.g., bonito) are more abundant seasonally in the vicinity of the eastern islands (Landberg 1965, 1975). Geographic proximity to the California Current also influences the distribution of sea mammals on the islands. Harbor seals live and breed along the shores of all the islands (Bartholomew 1967; Le Boeuf and Bonnel 1980; Odell 1971), although they favor the cool, productive waters surrounding Santa Rosa and San Miguel. The high primary and secondary productivity around San Miguel supports one of the largest sea mammal rookeries on the west coast of North America (Stewart et al. 1993). Most of these animals are concentrated on the western end of the island at Point Bennett (figure 6E), where harbor seals, northern fur seals, California sea lions, and elephant seals all have viable breeding colonies. Each year, large numbers of northern fur seals, California sea lions, and elephant seals visit San Miguel at various times to molt and breed (Le Boeuf and Bonnel 1980), and archaeological evidence indicates that southern fur seals were also more common at this location prehistorically (Walker and Snethkamp 1984; Walker et al. 2000). Although seasonal changes occur in the species that inhabit the rookery, large concentrations of animals are found throughout the year. Kelp forests off various parts of the northern Channel Islands provide a sheltered habitat for at least 125 different species of fish (figure 6F; Landberg 1965; Love 1996). Pedro Fages, traveling through the Santa Barbara Channel in 1775, noted that “the fishing is so good, and so great is the variety of fish, known in other seas, that this industry alone would suffice to provide sustenance to all the settlers which this vast stretch of country could receive” (Fages 1937, 35). The distribution and productivity of kelp forests around the northern Channel Islands are not only related to high nutrients and island size but also to the depth and character of subtidal habitats (Engle 1994). Size of subtidal shelf, wave exposure, substrate type (geologic formation), and bottom relief all influence where kelp and related faunas are distributed. The thickest kelp forests are present on the south sides of Santa Cruz and Santa Rosa, but San Miguel has the largest stands of kelp relative to its short coastline (Ehorn and Cornell 1979a, 1979b, 1980). Extensive kelp beds are also present on the northwestern side of San Miguel and along the northwest coast of Santa Rosa. Nearshore coastal habitats also vary greatly on the islands (figure 6G; also see map 3; Littler 1980). The rocky intertidal zone supports several

58

E N V I R O N M E N TA L C O N T E X T

species of mollusks, including California mussels (Mytilus californianus), black abalone (Haliotis cracherodii) and red abalone (H. rufescens). A variety of other mollusks (Tivela stultorum, Protothaca staminea, and Saxidomus nuttalli) are also found in island beach and estuarine habitats. Large quantities of Olivella biplicata, a small gastropod used prehistorically to make beads, are also found associated with island beach habitats. The longest stretches of beach are present on northern San Miguel, southwestern and eastern Santa Rosa, and southern and western Santa Cruz (map 3). One reasonably sized remnant estuary is situated on eastern Santa Rosa at the mouth of Old Ranch Canyon (figure 6H; see map 3 for location) and several smaller coastal wetland environments are found on Santa Cruz.

Seasonal Variability Seasonal fluctuations in the abundance and distribution of terrestrial and marine organisms occur on these islands and were important for shaping human subsistence decisions and strategies (figure 9). The distribution and productivity of plant communities on the islands are highly dependent upon annual rainfall (Junak et al. 1995). Native grasslands and acorn-bearing oak trees are more productive during wet years. In general, plants living in grassland and coastal sagebrush communities produce edible seeds, roots, and tubers during summer months (May–August). Blue dick bulbs (D. capitatum) are available throughout the year and are particularly conspicuous between March and May because of their small purple flowers. There are ethnohistoric records of people on the islands collecting blue dick bulbs during these months (Timbrook 1993). Acorns (Q. agrifolia) and pinenuts (P. muricata and P. torreyana) are available on Santa Cruz and Santa Rosa between October and December. Many of the marine resources available to past island peoples are present throughout the year. Mollusks are found in the rocky intertidal zone throughout the year except at times when small microorganisms make them poisonous (red tide). Kelp bed and rocky shore fish are available throughout the year, and fishing in the past was probably only impeded during the worst winter storms. Large numbers of schooling fish enter the channel during summer and fall months (sardines, yellowtail, bonito), sometimes venturing close to shore in search of food. This pulse was of great importance to late prehistoric

E N V I R O N M E N TA L C O N T E X T

59

WET Feb

Sa

R n ta

M

os a )

Bi

Loons, Geese & S rd s (

uz

&

De

c

Jan

ch

p h a nt S eals

ers

)

C

pt

st

Tu

ne

Se

s& oot s, R d e e S

Ju

ns Ha Lio rbor S eals, Sea

Augu

May

&

M

re Fish

Oct

Pela gic Fish

El eph ant Se or als mo ran be ts, P rs elicans & Rookery

N ears

o llu s ks

&

ho

Aco rns ( S

Nov

l Apri

a n ta

Cr

E le

ar

c ot

July

DRY

figure 9. Seasonal availability of resources on the northern Channel Islands (drafted by C. Goldsmith).

Chumash living on the coastal mainland who were consummate fisherfolk (Bernard 2001; Landberg 1965). Sea mammals are also more readily available during summer months (May–August) when large concentrations of animals are found at rookeries on San Miguel (DeLong and Melin 2000; Melin and DeLong 2000; Stewart et al. 1993). Although male northern fur seals and California sea lions leave the rookery in September, the females and pups remain on San Miguel throughout much of the year. Harbor seals are present throughout the annual cycle but breed in rookeries on San Miguel, Santa Rosa, and Santa Cruz during summer. Elephant seals visit rookeries on San Miguel to breed during winter months (January and February), and female elephant seals revisit during summer months to molt.

60

E N V I R O N M E N TA L C O N T E X T

Short-Term Climatic Variability Significant interannual, decadal, and century-scale climatic changes impact the marine and terrestrial resource base in the region. The dynamic interplay between atmospheric and oceanographic phenomena make short- and long-term climatic patterns very complex. Fluctuations in rainfall directly impact the distribution and productivity of native grasses and oak woodland (Junak et al. 1995). Changes in upwelling and the periodic impacts of El Niño/Southern Oscillation (ENSO) can drastically affect overall marine productivity and the availability of mollusks and nearshore fish. On the decadal timescale, ENSOs produce the largest environmental changes in Southern California. These climatic anomalies are driven by major oceanic and atmospheric changes in the equatorial Pacific. Pacific tradewinds are generally strong enough to cause surface-water flow toward the western Pacific. However, these tradewinds periodically weaken and cause eastward movement of warm surface waters along the equator (Ramage 1986). These warm currents flow southward from Ecuador along the Peruvian coast and northward along the west coast of North America. Thus, much of western North America is influenced by warm El Niño water. In California, warm water displaces cold water and reduces upwelling along the coast. Strong El Niño events have occurred every 7 to 19 years since recording began in 1944. Average monthly sea-surface temperatures recorded around the islands capture the effect of these anomalies in the Santa Barbara Channel region (see figure 8). During El Niño years, sea-surface temperatures along the central and Southern California coast are elevated throughout the annual cycle. Upwelling of cool, nutrient-rich water is reduced and causes overall decreases in marine productivity. Between 1982 and 1992, surface waters surrounding the islands were relatively cool except during the El Niño events that occurred in 1982, 1983, 1984, 1990, and 1992. Warmest sea-surface temperatures occurred between 1982 and 1984 and were associated with one of the strongest El Niño years in recent history. Warm sea-surface temperatures and reductions in upwelling disrupt kelp growth, and strong tropical storms cause physical disturbance to kelp beds (Dayton et al. 1992). These disturbances profoundly impact nearshore fisheries, particularly the ecological relationships between kelp bed and crust-dominated (i.e., sea urchins) communities (Ambrose et al. 1993; Dayton et al. 1992; Tegner and Dayton 1987,

E N V I R O N M E N TA L C O N T E X T

61

1991). Increases in water temperature kill sensitive marine species and impact the reproductive success of others (Engle 1994). Sea mammal populations are severely affected during these events (DeLong and Melin 2000; Stewart et al. 1993). California sea lion and fur seal pups suffer high mortality rates because females cannot obtain enough food to maintain lactation (DeLong and Melin 2000; Ono et al. 1993). Warmer-water species of fish are also introduced during these events. Atmospheric responses to El Niño/Southern Oscillation events are complex. In the Santa Barbara Channel region, increased precipitation often occurs during strong to very strong El Niños. This is because equatorial storms tend to track farther north during these intervals as water temperatures increase in the eastern Pacific. Weak and moderate El Niño events are not as clearly linked to increased precipitation levels in Southern California. For instance, the moderate El Niño of 1976–1977 was associated with one of the driest periods of the past 50 years (Ramage 1986). Intermittent drought is also part of the natural ecological cycle in the region, as are associated brush fires. Grass, sage, and chaparral seeds often lie dormant during drought years and germinate after fires or seasonally high rainfall. During La Niña intervals (between El Niños), drought is more common and terrestrial productivity lower.

Paleoenvironment Centennial- through millennial-scale paleoclimatic fluctuations are evident in climate records for the Santa Barbara Channel region spanning the past 12,000 years (Axelrod 1967; Behl and Kennett 1996; Cole and Lui 1994; Heusser 1978; Heusser and Sirocko 1997; Ingram and Kennett 1995; Kennett and Ingram 1995a, 1995b; Kennett and Kennett 2000; Pisias 1978, 1979). The short-term climatic variability evident historically (e.g., El Niño/La Niña cycles) likely occurred throughout much of the Holocene, superimposed on larger-scale climatic changes. However, some studies suggest that ENSOs were less frequent during the Early and Middle Holocene (Sandweiss et al. 1996). Proxy climate records are required for making inferences about the distributions of plant and animal communities on and around the islands during the prehistoric past and for reconstructing the selective environments that influenced changes in human behavior during the Holocene. Independent of the interests of archaeologists, paleoclimatologists are interested in the climate history of the region because it is situated

62

E N V I R O N M E N TA L C O N T E X T

at an ecological boundary between two large ocean current systems and is highly sensitive to global and local climate change (Kennett and Ingram 1995a; Pisias 1978). The unique orientation of the Santa Barbara Channel shelters an offshore basin (Santa Barbara Basin) that preserves remarkably detailed records of Holocene climate change. Limited sediment bioturbation and high sedimentation rates, caused by runoff and high primary productivity (plankton), preserve annual laminated sediments (varves). Deep-sea cores taken from Santa Barbara Basin provide some of the highest-resolution marine climate records in the world (Heusser 1978; Kennett and Kennett 2000; Pisias 1978). These climate records are described in this section within the context of regional and global climate change.

Sea Level Regional climatic changes in Southern California during the past 12,000 years developed in the context of broader global climatic fluctuations and postglacial sea level rise (Clark et al. 2002; Fairbanks 1989). Changes in sea level had a major impact on shoreline and wetland formation along the Southern California coast and offshore islands (Bickel 1978; Erlandson 1994; Inman 1983). During the last glacial maximum (~18,000 ka), sea level was 121 m below the current high-level stand (figure 10; Fairbanks 1989). At that time, the northern Channel Islands were amalgamated and formed a large landmass known as Santarosae Island (Orr 1968). A land bridge between the mainland and Santarosae never existed, but the distance from the eastern end of the island to the mainland was reduced to ~5 km (~3 miles) (Junger and Johnson 1980; Wenner and Johnson 1980). After ~18,000 BP, sea level worldwide rose until between 5,000 and 6,000 BP (figure 10a,b). Sea level reconstructions for the California coast indicate rapid inundation of the coastline until 7,000 BP and more gradual increases in sea level until 5,000 to 6,000 BP (figure 10c; Inman 1983). At the beginning of the Holocene (~10,000 years ago), sea level was still 20–40 m below current levels. The bathymetry surrounding the northern Channel Islands reveals that large areas of Santarosae Island were inundated between 18,000 and 7,000 BP, reducing the island’s land by about half (figure 10d; also see Porcasi et al. 1999). Rapid sea level rise certainly disrupted marine habitats and reduced overall terrestrial productivity on the islands. By the

E N V I R O N M E N TA L C O N T E X T

63

Present-Day Landmass 100 - 600 m 0 - 40 m > 600 m 40 - 100 m

Santa Barbara Point Conception Ventura

Anacapa

San Miguel Santa Cruz

d

Santa Rosa

figure 10. Global and local sea level changes over the past 18,000 years (kyr  thousand years): (a, b) Well-established global sea level record based on oxygen isotopic shifts in corals from the Caribbean (data from Fairbanks 1989); (c) estimated sea level changes for the California coast (data from Inman 1983); (d) bathymetry of the Santa Barbara Channel region (drafted by D. Kennett).

Early Holocene the channels between Anacapa, Santa Cruz, Santa Rosa, and San Miguel had been flooded and increases in sea level changed the base level of drainages along the coast that were down-cut and scoured out during the glacial maximum. In some parts of California, estuaries and marshes formed in the mouths of larger drainages (Bickel 1978). By ~7,000 BP, extremely productive estuaries punctuated the Santa Barbara

64

E N V I R O N M E N TA L C O N T E X T

coast and were extensively used by prehistoric peoples (Erlandson 1994). The distribution of archaeological sites at the mouth of Old Ranch Canyon on Santa Rosa indicates that a reasonably large and productive estuary formed at around 7,000 years ago. These have since largely filled in. Remnant estuaries and small wetlands are found on Santa Cruz Island at Prisoners Harbor, Scorpion Anchorage, Cañada de los Sauces, and Cañada de Malva Real (see maps 2 and 3 for locations; Junak et al. 1995).

Marine Climate History Changes in sea-surface temperature (SST) and marine productivity during the past 11,000 years are evident in several proxy climate records from the coast of California (Heusser et al. 1985; Pisias 1978, 1979; Van Geen et al. 1992). High-resolution Holocene climate records have also been established for coastal Southern California (Kennett and Kennett 2000). This Holocene record represents the upper 17 m of a 200-m core, a Late Quaternary sequence spanning the past 160,000 years (893A), drilled in Santa Barbara Basin as part of the Ocean Drilling Program (map 4, no. 1; Behl and Kennett 1996; Cannariato et al. 1999; Hendy and Kennett 1999, 2000; Ingram and Kennett 1995; Kennett and Ingram 1995a, 1995b). The Holocene (11,000 BP to present) is represented by a 17-m sequence of laminated sediments deposited at an average rate of ~155 cm/1,000 years. Climatic change through the Holocene is inferred from oxygen isotopic (18O) analysis of two planktonic foraminiferal species: Globigerina bulloides, a surface dweller, and Neogloboquadrina pachyderma, which lives near the base of the thermocline (~60 m below surface). The Holocene age model is based on 20 AMS 14C dates converted to calendar years using a reservoir age of 230  35 years (Ingram and Southon 1996; Kennett et al. 1997). This has provided one of the highest-resolution marine Holocene climate sequences in the world: 25year intervals for the past 3,000 years and 50-year increments for the remainder of the Holocene. The climate record from Santa Barbara Basin is of high quality because of rapid sedimentation rates, lack of bioturbation, continuous abundance of foraminifera, and a highly sensitive environmental setting (Kennett and Kennett 2000). The new marine climate sequence demonstrates that the Holocene was not climatically stable, but instead it exhibits millennial-scale oscillations in SST (figure 11A; Kennett and Kennett 2000). Compared with the Late Pleistocene (Kennett and Ingram 1995a), Holocene SSTs

E N V I R O N M E N TA L C O N T E X T

124°

122°

120°

118°

116°

65

114°

3

40°

Nevada

2

d va Ne rra Sie

Lake Tahoe

N

9

38° San Francisco Bay

40°

38°

Mono Lake

e hit

W

a

ns

Pa

36°

tai un Mo

8

Tulare Lake

36°

ic

cif

11

n

ea

Oc

34°

7

10 1 6

34°

5

Salton Sea

Channel Islands 4

map 4. Map of California indicating the location of the primary paleoclimatic records discussed. (1) Ocean Drilling Project, Hole 893a (Kennett and Kennett 2000); (2) Leonard rockshelter pollen sequence (Byrne et al. 1979); (3) Ruby Valley pollen record (Thompson 1992); (4) archaeological pollen sequences (Masters and Gallegos 1997); (5) San Joaquin Marsh pollen spectra (Davis 1992); (6) Santa Rosa Island pollen sequence (Cole and Lui 1994); (7) Union pollen spectra (Morgan et al. 1991); (8) bristlecone pine tree-ring record (La Marche 1973, 1974); (9) Lake Tahoe submerged tree-stump record (Lindström 1990); (10) Point Reyes pollen spectra (Rypins et al. 1989) (drafted by D. Kennett).

were warm (12.5ºC on average), but distinctive cold- and warm-water cycles are present in the record. Nine of these cycles in SST are evident during the Holocene (figure 11A). Warm-water intervals occurred between 11,000 and 9,600 BP; 8,200 and 6,300 BP; 5,900 and 3,800 BP; and 2,900 and 1,500 BP; along with general warming from 500 BP to the present. Cold-water intervals developed between 9,600 and 8,200 BP; 6,300 and 5,900 BP; 3,800 and 2,900 BP; and 1,500 and 500 BP. The

0

W

1

C

2 3

AGE (ka)

9

10

SST (°C)

11

12

13

Vertical Stratification/Productivity 14

15

∆δ

+

18O (pachyderma-bulloides)

-0.4 -0.2 0

0.2 0.4 0.6 0.8 1

Coefficient of Variance

-15 -10

0

5

10

15

0 1

-

Highly 2 Variable

N

3

+

C

-5

4

4

5

W

-

5

6

C

+

6

7

-

8 9 10 11 12

Highly Variable

+

W

+

C

+ + + -

W

A

0.6 0.4 0.2

0 -0.2 -0.4 -0.6

δ18O (PDB)

7 8 9 10

B

+

PRODUCTIVITY



C

11 12

E N V I R O N M E N TA L C O N T E X T

67

coldest SSTs during the Holocene occurred between 1,500 and 500 years ago and the warmest between 5,900 and 3,800 BP. Fluctuations in marine productivity also occurred during the Holocene (figure 11B), often synchronous with cold- and warm-water episodes. Changes in marine productivity have been inferred using a marine-productivity index. This index is based on inferred temperature differences between surface waters (as measured by the oxygen isotopic composition of surface-dwelling G. bulloides) and waters at the base of the thermocline (as measured by the oxygen isotopic composition of N. pachyderma, which inhabits the thermocline). Modern studies indicate that the isotopic difference between G. bulloides and N. pachyderma reflects the thickness of the upper mixed layer and therefore provides a proxy for the strength of upper-water-column stability, the intensity of upwelling, and overall marine productivity (Pak et al. 1997). During the Holocene, inferred warming in surface waters was often associated with cooling at the thermocline, and vice versa, suggesting episodic thickening and thinning of the mixed layer and associated depth changes in the thermocline. During cool episodes, little or no vertical temperature gradient exists between the surface and thermoclinal species. This suggests that the upwelling of nutrient-rich deep ocean waters was then especially intense. Inferred changes in marine productivity during the Holocene were also generally associated with changes in SST. Vertical mixing and inferred high marine productivity were greatest between 3,800 and 2,800 BP and again between 1,000 and 500 BP. Reduced vertical mixing and lower marine productivity were sustained between 6,000 and 3,800 BP and again between 2,800 and 1,000 BP. As sea level stabilized along the California coast at ~5,500 BP, climatic conditions became increasingly unstable. Inferred average annual SST varied by 2C to 3C between about 11,000 and 5,500 BP,

figure 11. New Holocene climate record from the Santa Barbara Basin. (A) sea-surface-temperature curve is based on the oxygen isotopic composition of G. bulloides (surface-dwelling species of foraminifera) from varved sediments in the Santa Barbara Basin. Sea-surface temperatures were estimated based on Bemis et al. (1998). Bar on left side of figure represents warm (w) and cold (c) water cycles through the Holocene; (B) vertical stratification/productivity record based on differing oxygen isotopic composition of G. bulloides and N. pachyderma (deeper-dwelling species of foraminifera). Bar on the left of figure shows intervals inferred as high () or low () productivity during the Holocene (drafted by D. Kennett).

Pollen Zones 0

AGE (ka)

3

9

10

11

12

13

14

warm/dry

15

17

6

9

21

22 0

4

cool/wet

1

3 4

2 3 4

Altithermal warm/dry

5

warm/dry

6

7 8

20

2

2 3

Medithermal cool/wet

19

1

4 5

18

1

cool/wet

1 2

PRECIPITATION (cm)

SST (°C)

Union Project

Antevs (1955) Point Conception

7

Drying

8

Anathermal cool/wet

9

warm/wet 10 11

10

Morgan et al. 1991

A

B

C

D

11 12

12

0.6

0.4

0.2

0

-0.2 -0.4 -0.6

δ18O (PDB)

E N V I R O N M E N TA L C O N T E X T

69

but varied from 3C to 5C from ca. 5,500 BP to present. A statistical test of variance (coefficient of variation, standard deviation/average) on SST (50-year intervals) indicates increasing climatic instability during the Late Holocene (figure 11C). During the latter half of the Holocene, climatic conditions were particularly unstable between 4,000 and 3,000 years ago and again between 1,500 and 800 years ago.

Terrestrial Climate History Fluctuations in terrestrial environments also are inferred in California during the Holocene (11,000 BP-present) and appear to be tied more broadly to climatic changes across western North America (Thompson et al. 1993; Whitlock 1992; Whitlock and Bartlein 1997). Pollen evidence indicates that pine forests were dominant in coastal California during the last glacial episode, whereas oaks became predominant during the Holocene (Axelrod 1967, 1981; Johnson 1977; West and Erlandson 1994). This trend was associated with regional warming from glacial to interglacial conditions. Compared with the dramatic regional floral shifts that occurred in California between glacials and interglacials (Heusser and Sirocko 1997), Holocene pollen spectra provide evidence that plant distributions in coastal Southern California were similar to those of the present day (Adam et al. 1981; Byrne 1979). Nevertheless, environmental conditions across western North America did fluctuate during the Holocene. Based on limited investigations in the Great Basin, Antevs (1948, 1952, 1955) argued that the Middle Holocene (7,000–4,500 BP) was warm and dry across much of western North America, the so-called Altithermal or climatic optimum (figure 12A). This was preceded by the Anathermal (10,000–7,000 BP) and followed by the Medithermal (4,500 years ago to present), both intervals consisting of generally cool and wet climatic conditions. More recent work indicates that dry conditions during the Middle Holocene

figure 12. Terrestrial climate records for the Holocene shown against the new sea-surface-temperature record from the Santa Barbara Basin. (A) Holocene climate phases (data from Antevs 1955); (B) composite climate sequence based on pollen data from the Union Project north of Point Conception (data from Morgan et al. 1991); (C) sea-surface-temperature record from the Santa Barbara Basin; (D) bristle cone pine record from interior California (data from La Marche 1973, 1974) (drafted by D. Kennett).

70

E N V I R O N M E N TA L C O N T E X T

existed in the Great Basin and precipitated significant decreases in lake levels and changes in plant distributions (see map 4, nos. 2, 3, and 9; Bright 1966; Byrne et al. 1979; Hansen 1947; Madsen and Rhode 1990; Mehringer 1985; Thompson 1992). Dry Altithermal conditions had less of an impact on coastal California environments during the Middle Holocene. Relatively dry conditions in the Santa Barbara Channel region are suggested by high percentages of Chenopodium and Ambrosia pollen in estuarine deposits on Santa Rosa Island between 5,200 and 3,250 BP (map 4, no. 6; Cole and Lui 1994). Based on pollen from sediment columns extracted north of Point Conception (map 4, no. 7), Morgan et al. (1991) also argued for dry conditions peaking in the Middle Holocene along the Santa Barbara coast (7,600–4,800 BP; figure 12B). In Santa Barbara Basin (ODP Hole 893A; map 4, no. 1), frequency changes of pine and oak (Heusser and Sirocko 1997) exhibit no distinct trends during the Middle and Late Holocene and climatic interpretations are inconclusive. Pollen spectra from estuarine and archaeological deposits in coastal San Diego County (map 4, no. 4) indicate relatively stable environmental conditions during the Holocene (Masters and Gallegos 1997). Changes in stream flow into San Joaquin Marsh (map 4, no. 5) suggest cyclical fluctuations in rainfall during the Middle and Late Holocene (Davis 1992). The Middle Holocene sedimentary history of this wetland environment indicates a rapid flooding event at ca. 5,000 BP and more sustained stream flow between 4,000 and 3,500 BP, suggesting moist conditions during this interval. Late Holocene tree ring and lake level records suggest dry conditions in Southern California between approximately 1,100 and 600 BP (map 4, nos. 10 and 11; Larson and Michaelson 1989; Stine 1994). Many of the available Holocene pollen and lake level records are not continuous, and it is likely that terrestrial environmental conditions were more cyclical, tracking the millennial-scale global climate changes now evident in a variety of proxy records worldwide (Kennett and Kennett 2000). The relationship between marine and terrestrial climatic conditions on the California coast are complex, but historical data suggest that these two climate systems are closely interrelated today (Jones and Kennett 1999). Based on evidence from the Late Holocene, it appears that especially cold SSTs in the Santa Barbara Basin were correlated with intervals of low precipitation over parts of western North America (Kennett and Kennett 2000). A comparison of the Santa Barbara Basin core data with the bristle cone pine record from

E N V I R O N M E N TA L C O N T E X T

71

the White Mountains of eastern California suggest a broad correlation between cool SSTs and drier conditions during the past 4,000 years (figure 12D; see map 4, no. 8). During the past 4,000 years, cold SSTs and low precipitation levels dominate between 4,000 and 2,300 BP and again between 1,500 and 500 BP; warm SSTs and more precipitation are evident between 2,300 and 1,500 BP and again after 500 years ago. Cool SSTs between ~1,500 and 500 years ago also correlate with lower precipitation levels evident in a shorter tree ring record from the coastal ranges of Southern California (Kennett and Kennett 2000; Larson and Michaelson 1989). Several other lines of evidence also indicate dry conditions during this period of time (Jones et al. 1999; Raab and Larson 1997; Stine 1994). Prior to 4,000 BP the correlation between SST in the Santa Barbara Channel region and the bristlecone pine record is not clear, possibly reflecting an overall shift in the mode of climatic change in Southern California.

Summary The distribution of potential subsistence resources on the northern Channel Islands is patchy and highly variable temporally. Terrestrial resources are depauperate relative to the mainland coast. No large terrestrial mammals were available prehistorically, and the distribution of grasses, sagescrub, and oak woodland is patchy and productivity is temporally unpredictable. Except in some of the larger drainages on Santa Rosa and Santa Cruz, water availability is also limited. Although marine resources are rich, their distribution and productivity also varies through space and time. The availability of marine and terrestrial resources is dependent upon short- and long-term environmental changes that are evident in long-term proxy climate records for the region. These data provide an environmental context for the analysis of foraging strategies on the islands and the emergence of sociopolitical complexity during the Holocene. In chapter 4, the broad cultural and historical developments in the Santa Barbara Channel region are explored.

chapter 4

Cultural Context

This chapter reviews the ethnohistoric and prehistoric records for the Santa Barbara Channel region to provide a broad context for evaluating the cultural developments on the northern Channel Islands during the past 13,000 years. Compared with other native Californians, the contact-period Chumash had a relatively complex sociopolitical life and the origins of this complexity are currently the subject of a spirited intellectual debate (e.g., Arnold 1991, 1992a, 1993, 1997, 2001; Arnold and Green 2002; Erlandson and Rick 2002a; Gamble and Russell 2002; Gamble et al. 2001, 2002; Johnson 2000; Kennett and Kennett 2000; King 1990; Raab and Larson 1997). In this chapter, the nature of Chumash sociopolitical complexity is discussed and the debate regarding when and why it emerged is outlined.

The Ethnohistoric Record In October of 1542, Juan Rodríguez Cabrillo entered the Santa Barbara Channel region and became the first documented European to contact the Chumash people occupying the coast between Malibu and San Simeon (see map 1), including the rugged mountains of the interior and the northern Channel Islands. Cabrillo, and later Sebastián Vizcaíno (ca. 1602), were exploring Alta California for sheltered harbors to facilitate the Maníla Galleon trade between New Spain (Mexico) and the Philippines (Johnson 1982; Wagner 1929). The chronicles of these visits, although brief, provide a tantalizing glimpse of traditional Chumash 72

CULTURAL CONTEXT

73

lifeways in the 1500s. After a hiatus of largely undocumented exploration, Gaspar de Portolá led an overland expedition into Southern California in 1769 that ultimately provided the foundation for the first Franciscan missions and the beginning of the mission period (Johnson 1988; Johnson and McLendon 2000). Much of the existing ethnohistoric information for the Santa Barbara Channel region comes from observations made by members of the Portolá expedition and early Franciscan missionaries. Although the impact of early Spanish exploration and the introduction of European diseases during the protohistoric period (AD 1542– 1769) is poorly known and a subject of current debate (e.g., Erlandson and Bartoy 1995, 1996; Erlandson et al. 2001; Johnson 1988; Preston 1996, 2002), the devastation of epidemics and the cultural changes associated with missionization are well documented (Johnson 1988; Larson et al. 1994). Regardless of the potential demographic shifts and changes in sociopolitical organization caused by European disruption, the ethnohistoric record provides a valuable starting point for investigating the prehistoric record in the region. Nevertheless, the available ethnohistoric data provide only a glimpse of Chumash society recorded during a period of severe disruption, and it cannot be assumed that these accounts characterize the cultural milieu deep into the prehistoric past.

chumash population levels Population estimates for the Chumash region, based on Portolá’s expedition and baptismal records, range from 8,000 to 25,000 (Brown 1967; Cook 1964, 1976; Johnson 1988; Kroeber 1925). Given the accounts of the Portolá expedition, Cooke (1940) estimated that 5,000 people were living along the mainland coast south of Point Conception, with another 1,000 living on the coast north of Point Conception and 3,000 occupying the broad expanses of the interior. When population assessments made by Portolá’s expedition are compared with baptismal records for each village, the number of baptisms is usually smaller, a pattern that is at least partially related to high mortality rates caused by introduced European diseases (Brown 1967). Johnson (1988) argued that Portolá’s estimations were inaccurate and that population declines during the mission period were overstated. Regardless, the largest Chumash villages and the highest population densities occurred along the mainland coast between Malibu and Point Conception (map 5).

map 5. Distribution of Chumash villages on the northern Channel Islands and adjacent mainland. Relative settlement sizes are based on the number of baptisms at the missions in the region (data from Johnson 1982, 1988; drafted by R. van Rossman).

CULTURAL CONTEXT

75

Precontact island populations are thought to have approached 3,000, a conservative estimate derived from both baptismal records (1,270 people) and hypothesized impacts of introduced European diseases (Johnson 1982; Johnson and McLendon 2000).

subsistence Ethnohistoric data indicate that people living in the Santa Barbara Channel region had a mixed subsistence economy consisting of a variety of food items. Of these, marine foods, acorns, and large, terrestrial game animals (e.g., deer) were the most important. Interior Chumash people were less reliant upon marine resources, focusing on readily available and abundant acorns and terrestrial animals. Seasonally, people from the interior visited the coast to acquire marine foods and other items through trade with coastal dwellers. People living along the mainland coast and offshore islands used a variety of marine resources (e.g., sea mammals, shellfish), but early ethnohistoric accounts suggest that fishing was the most important subsistence pursuit (Landberg 1965). The plank canoe was important for open-ocean fishing, and the Chumash had a well-developed and diverse fishing technology (fishhooks, nets, harpoons, etc.; Hudson et al. 1978). Acorns were also intensively collected along the mainland coast to complement proteinrich marine resources. The combination of terrestrial land mammals, rich marine foods, and concentrated, acorn-bearing oak groves supported large coastal populations. Islanders were more heavily dependent upon marine resources, but seasonally collected available seeds and tubers and traded craft items (beads, etc.) for acorns and other plant foods from the mainland to supplement their carbohydrate-poor diets (Timbrook 1993).

settlement The Chumash people living along the Santa Barbara Channel were described by early Spanish explorers as living in relatively large, permanent villages. In 1793, Archibald Menzies noted, “The Natives live in Villages of from 20 to 40 huts each which are crowded together and much larger than any we saw about the Settlements to the Northward” (Eastwood 1924, 325). Both the seasonal and the long-term stability of these population centers are poorly known and variations in community mobility within Chumash territory were never systematically

76

CULTURAL CONTEXT

documented. Much of the ethnohistoric information regarding settlement comes from the mainland coast, with only a few accounts of settlement on the northern Channel Islands, the mainland interior, or the coast north of Point Conception. Regional variation in community mobility must have existed, with the composition and location of villages changing in response to differing environmental and social contexts. Landberg (1965) inferred seasonal settlement shifts in Chumash territory based on early journals of Spanish explorers who passed through the area and the interrogatorios kept by early missionaries. He argued that during rainy winter months people aggregated in relatively large villages and relied upon stores of acorns and fish, supplemented by available plant foods (e.g., sprouts). These were probably the principal historic villages described by Spanish explorers and missionaries. Starting in late spring and extending through fall, Landberg argued that Chumash populations were more dispersed and mobile. In spring, as winter rains became less frequent and plant foods more abundant (bulbs, roots, and tubers), some people dispersed and established small campsites closer to hunting and collecting grounds. As water supplies continued to dwindle during summer months, people remained dispersed, possibly aggregating for short periods of time at the end of summer when schooling fish (e.g., tuna) were abundant in the Santa Barbara Channel. Community mobility increased again in fall months as people moved to oak groves in the coastal range and to the more rugged interior. Large quantities of acorns were collected at this time and transported to larger villages and stored for winter months. Much of Landberg’s model is based on information acquired from the mainland coast south of Point Conception. He argued that the permanence of villages was dependent upon the seasonal availability of resources and the abundance of resources in local habitats. With regard to coastal settlement, Landberg states that “No doubt the nearness of most food sources as well as the stability of their food supply enabled the greater part of the Santa Barbara coastal population to remain permanently based in large villages” (Landberg 1965, 112). He argued that villages on the northern Channel Islands were also relatively stable but that islanders focused primarily on marine resources because of the limited array of terrestrial fauna and flora. Based on analogy with other California Indian groups, Landberg suggested that the people living in the interior of Chumash territory were more mobile than coastal groups.

CULTURAL CONTEXT

77

sexual division of labor Like many hunter-gatherer societies, Chumash people had a distinct sexual division of labor during the protohistoric and historic periods. There was probably tremendous variability among households, but, in general, young and middle-aged men living on the mainland coast and islands were primarily involved in open-water fishing and sea mammal hunting (Blackburn 1975; Landberg 1965). Women were responsible for much of the food processing (e.g., acorns and fish) but also specialized in gathering wild plants and shellfish (Landberg 1965; Walker and Hollimon 1989). Acorn collecting was an important economic pursuit primarily practiced by women, but elderly men sometimes participated in the acorn harvest (Walker and Hollimon 1989). Males and females often produced subsistence-related equipment specific to the task at hand. For instance, men specialized in making fishing equipment, whereas women produced baskets used for collecting plant foods. Some men also belonged to the Brotherhood of the Tomol, a group of male specialists who manufactured, owned, and operated plank canoes used for fishing and cross-channel exchange (Arnold 1995; Gamble 2002; Hudson et al. 1978). Both men and women were involved in the bead-manufacturing industry that was of great importance to islanders during the late and historic periods (Arnold 1987; Heizer 1955).

sociopolitical organization Chumash people were organized into extended family units (clans) governed by an older man or woman with high status and prestige (Johnson 1988). Villages along the coast were composed of multiple clans, and these groups were sometimes organized hierarchically. Most clans were organized matrilinealy, but the most elite lineages were organized patrilineally (Johnson 1988). It is generally assumed that the Chumash were organized hierarchically (simple chiefdoms), and ascribed leadership roles were the norm by historic contact (Arnold 1992a, 1993, 1995; Gamble 1991; Gamble et al. 2001; King 1990; Raab 1994; Raab and Larson 1997). Much of the debate over the level of political integration has centered around whether the Chumash were organized as a chiefdom (Arnold 1991, 2001; Gamble 1991; Gamble et al. 2001; King 1990). Indeed, certain villages were identified as capitals by early Spanish explorers—these were places where village chiefs from throughout the region met periodically (Johnson 1988). In 1542, Cabrillo encountered a female chief living in the village of Siujtu (on the mainland coast near Santa Barbara) who ruled all

78

CULTURAL CONTEXT

of the villages between Point Conception and Santa Barbara (Gamble 1991; Johnson 1988). In the early mission period, Chief Yanonali, also of Siujtu, reportedly had political influence with at least 13 villages along the coast (Brown 1967; Gamble 1991; Johnson 1988). The question is, Do these limited ethnohistoric accounts accurately depict Chumash sociopolitical organization? Johnson (2000) has argued that they may not have been accurate and that the Chumash were primarily organized at the village level. Most villages were politically autonomous, and the main function of chiefs was to lead war parties and organize intervillage fiestas. It appears that their power was somewhat limited beyond these functions (Geiger and Meighan 1976; Johnson 2000), but well-respected leaders were probably influential in many political situations. Chiefs living in larger villages on the mainland coast and islands likely had more power in economic decision making compared with chiefs living in smaller communities, but the degree of political cohesiveness between villages apparently was limited and depended upon the influence of individual men and women. Large political alliances were opportunistic, and, chiefly, rule on the regional level was highly unstable and not institutionalized.

specialization and exchange One of the hallmarks of Chumash society was intense economic interaction between individuals. Virtually all of the Spanish explorers who passed through the area were met by people in plank canoes eager to exchange food and goods. In the late 1700s, Longinos Martinez noted the following: All these Indians are fond of traffic and commerce. They trade frequently with those of the mountains, bringing them fish and beadwork, which they exchange for seeds and shawls (tapalos) of foxskin, and a kind of blanket made from the fibers of a plant which resembles cotton; and they prefer these to their own which they make from sea otter. When they trade for profit, beads circulate among them as if they were money, being strung on long threads, according to the greater or smaller wealth of each one. In their bargaining they use, as we use, weights, their poncos of strings of beads. This word ponco (Gabrieleno bead measure, Kroeber 1925, 565), is used for certain measure of these strings, two turns from the wrist to the extended middle finger. The value of the ponco depends on the esteem in which the beads are held, according to the difference in fineness and the colors that are common among them, ours being held in higher regard. The value depends

CULTURAL CONTEXT

79

upon the greater or smaller extent to which the beads have been circulated, the new values depending upon their abundance. The value which should be placed upon our beads is always estimated with respect to their own, and in everything they keep as much order as the most careful man who has accumulated some money. (Simpson 1939)

Individuals living in villages on the mainland coast traded food and manufactured items with people in the interior and on the northern Channel Islands. Shell beads were sometimes used as a medium of exchange for purchasing manufactured goods, services, and food (King 1976, 1990). In addition to subsistence-related activities, some people made canoes, mortars and pestles, bows and arrows, nets, and other utilitarian goods, whereas others produced craft items such as beads or pendants (Arnold 1987; Blackburn 1975; King 1976, 1990). Ethnohistoric accounts rarely describe the time individuals devoted to the production of utilitarian and craft goods, but these pursuits were probably not full-time occupations. Marriage alliances between island and mainland villages, often directly across the Santa Barbara Channel, facilitated intervillage commerce (Johnson 1988).

warfare and territoriality Lambert (1994, 1997) reviewed the available ethnohistoric evidence of violent conflict in the Santa Barbara Channel region. Although there are many accounts of “peaceful” Chumash people trading with one another, there are also many instances of interpersonal and group conflict. Many of these accounts indicate that violent interaction between villages in the Santa Barbara Channel region was relatively common (Johnson 2000). In 1542, Cabrillo wrote “From the pueblo de las Canoe (Magu?) to ‘Cabo de Galera’ (Point Conception) there is a well inhabited province called ‘Xexu.’ There are many different languages, and they carry on great wars with each other” (Wagner 1929, 87). As one would expect, there is evidence that individuals had complex relationships with people in other villages, trading in a friendly manner with some and engaging in warfare with others. Most of the Spanish that entered the region were well received, but virtually all of them noted intervillage conflicts. Fages wrote, “They receive the Spanish well, and make them welcome; but they are very warlike among themselves, living at almost incessant war, village against village” (1937, 31). There is also evidence that groups from the interior of California

80

CULTURAL CONTEXT

came to the coast and waged war against coastal populations. Raiding and more cold-blooded forms of warfare were also documented and apparently were relatively common (Lambert 1994).

ethnohistoric summary The highest population concentrations and largest villages occurred along the mainland coast, but large interior and island populations also existed. Marine resources (particularly fish) and acorns were of great importance to people living in coastal communities but terrestrial foods were more limited in island contexts. Fishing was important throughout the year, and acorns were collected during fall months and stored for later use. Marriage alliances connected people living on the northern Channel Islands with the mainland coast and interior (Johnson 1988). Cooperation between individuals in different villages throughout the region was manifested through exchange of both food and craft items (King 1976). Craft specialization was an integral part of these exchange relationships, particularly on the northern Channel Islands where terrestrial plant foods were limited. Competition and violent interaction between people occurred, and chiefs, some with inherited status, played an important role in warfare.

The Prehistoric Record The Santa Barbara Channel region has a long and rich cultural history beginning in the Terminal Pleistocene (~13,500 BP) (Erlandson 1993a, 1993b, 1994; Erlandson et al. 1996a; Johnson et al. 2000). Across the Holocene, there appears to be a general trend toward population increase, aggregation of populations into semisedentary villages, more intense use of marine resources, greater social differentiation, and political centralization (Glassow et al. 1988; King 1990; Lambert and Walker 1991). This culminated in the development of the complex huntergatherer societies evident at the time of European contact (Arnold 1991, 1992a, 1993, 2001; Johnson 1988, 2000).

chronology There are several chronological schemes for the Santa Barbara Channel region (e.g., Orr 1968; Rogers 1929; Wallace 1955), but the most comprehensive and widely used was developed by C. D. King (1982, 1990; figure 13). King (1982, 1990) established the current chronological sequence based on seriation of artifacts (beads, pendants, etc.) found in

King 1990

0

Late Period

Phase 3 Phase 2 Phase 1 Phase 5

Erlandson & Arnold 1992 Colten 1991

1950

Historic

Late Period Middle/Late Transition

Phase 4

1

1000

Phase 3

Middle

Phase 2

Period

Period

2

Middle

Late Holocene 0

Phase 1

Phase Ez

1000

3

2000

5

6

Phase Eya

Early

Early

Period

Period

Middle Holocene

3000

BC-AD

Kya

4

4000

Phase Eyb

5000

7 Phase Ex

6000

8

Early Holocene 9

7000

10

8000

figure 13. Primary chronological schemes used for northern Channel Islands prehistory (data from Arnold 1992a; Erlandson and Colten 1991; King 1990). All of these chronological schemes have been calibrated to calendar years (Kya  thousand years) (drafted by D. Kennett).

82

CULTURAL CONTEXT

burial lots throughout Chumash territory. Based on changes in artifact styles and limited radiocarbon dating, King (1982) originally defined three primary cultural horizons: Early Period (6000–1400 BC), Middle Period (1400 BC–AD 1150), and Late Period (AD 1150–1804), with a further reduction of each period into phases (e.g., Late Period, phase 1, 2, and 3). More recently, King (1990) refined the Santa Barbara sequence owing to changes in the seriation of bead types in California by Bennyhoff and Hughes (1987) and advancements in radiocarbon dating. The main change in the new chronology is a boundary shift between the Early and Middle Periods from 1400 BC to 600 BC along with phase shifts within the Early Period and the early Middle Period (phase M1–M3) (King 1990, 28). Arnold (1992a) has refined the chronology for the Late Period based on excavations of archaeological deposits on Santa Cruz Island and charcoal dates calibrated to calendar years (Stuiver and Braziunas 1993). She puts the start of the Late Period at 650 BP (AD 1350) and defines a new period known as the Middle to Late transition (800–650 BP; AD 1150–1350). Table 9 shows the relationship between the King (1990) and Arnold (1992a) chronologies. King’s original chronology was founded upon uncalibrated radiocarbon dates and could not be compared directly to Arnold’s chronology and the regional paleoenvironmental data. To facilitate direct comparison, King’s (1990) chronology is calibrated to calendar years using Stuiver and Braziunas’ (1993) Calib 3.0.3 program (see Erlandson and Colten 1991). A number of assumptions were made to calibrate the chronology. First, much of King’s (1990) chronology is based on radiocarbon dates of marine shell, so a reservoir age (R) of 233  35 was subtracted from each of the boundary and phase dates [in addition to the ~400 year global reservoir accounted for by Stuiver and Braziunas (1993)]. Second, King’s marine shell dates were run during the 1970s and were not corrected for 13 C/12C fractionation, a standard practice today. On average, the 13 C/12C fractionation factor for shell in the Santa Barbara Channel region is 430 (Erlandson 1988b). This was added to each of the period and phase boundary dates. The calibrated calendar ages were rounded to the closest century. Arnold’s age assignments for the Early to Middle Period boundary are based on King and were calibrated in the same way. When calibrated, King’s (1990) Middle Period, phase 5, correlates relatively well with Arnold’s (1992a) Middle to Late Period transition. Erlandson (1988b; Erlandson and Colten 1991) more recently divided the cultural sequence in the Santa Barbara Channel region into four

table 9. Two primary chronological schemes used for the Santa Barbara Channel region King 1990

Calibrated (BC–AD)

Calibrated (BP)

Arnold 1992a

Calibrated (BC–AD)

Calibrated (BP)

Early Holocene

n/a

n/a

n/a

BC 8000–5500

BC (8000)–6120

9,950–8,060

Early Period Phase Ex Phase Eyb Phase Eya Phase Ez

BC 5500–600 BC 5500–4000 BC 4000–3000 BC 3000–1000 BC 1000–600

BC 6120–490 BC 6120–4650 BC 4650–3590 BC 3590–970 BC 970–490

8,060–2,440 8,060–6,600 6,600–5,540 5,540–2,920 2,920–2,440

BC 5500–600

BC 6120–490

8,060–2,440

Middle Period

BC 600–AD 1150

BC 490–AD 1380

2,440–570

BC 600–AD 1150

BC 490–AD1150

2,440–800

Phase 1 Phase 2 Phase 3 Phase 4 Phase 5

BC 600–200 BC 200–AD 400 AD 400–700 AD 700–900 AD 900–1150

BC 490–AD 170 AD 170–660 AD 660–980 AD 980–1170 AD 1170–1380

2,440–1,790 1,790–1,290 1,290–970 970–780 780–570

Middle/Late Trans.

n/a

n/a

n/a

AD 1150–1300

AD 1150–1300

800–650

Late Period Phase L1 Phase L2 Historic (L3)

AD 1150–1782 AD 1150–1500 AD 1500–1782 AD 1782–1804

AD 1380–1782 AD 1380–1670 AD 1670–1782 AD 1782–1804

570–168 570–280 280–168 168–146

AD 1300–1782

AD 1300–1782

650–168

AD 1782–

AD 1782–

168–

sources: King 1990; Arnold 1992. Both schemes are calibrated to calendar years.

84

CULTURAL CONTEXT

broad periods: Terminal Pleistocene (before 10,000 BP), Early Holocene (10,000–7,000 BP), Middle Holocene (7,000–3,350 BP), and Late Holocene (3,350–200 BP). When King’s chronology is calibrated, the Early Holocene generally overlaps with part of the Early Period, phase Ex (8,060–6,600 BP) (see figure 13). The Middle Holocene corresponds with the Early Period, phase Eyb (6,600–5,540 BP) and phase Eya (5,540–2,920 BP). The Late Holocene encompasses the remainder of King’s chronology: Early Period, phase Ez (2,920–2,440 BP), Middle Period (phases M1–M5; 2,440–570 BP), and Late Period (phase L1–L3; 570–146 BP).

cultural overview The earliest evidence for occupation of the Santa Barbara Channel region are the Terminal Pleistocene skeletal remains from Arlington Springs (Santa Rosa Island; ~13,500 BP; Johnson et al. 2000) and midden deposits at Daisy Cave (San Miguel Island; ~12,300 BP; Erlandson 1994; Erlandson et al. 1996a, 1996b; Rick et al. 2001a). The only other evidence for occupation of the region during the Terminal Pleistocene is a fluted Clovis-like projectile point that was found on the mainland coast (Erlandson 1994; Erlandson et al. 1987). It is typologically similar to points found across western North America between 12,000 and 11,000 BP, but it was not found in a primary depositional context. Boats of some kind were needed to colonize the northern Channel Islands during the Terminal Pleistocene, and the deposits at Daisy Cave suggest that diverse foraging strategies existed in North America at this time (see Jones et al. 2002). Evidence for the occupation of the Santa Barbara Channel region during the Early Holocene is more substantial (10,000–7,000 BP). People living along the mainland coast during this period collected shellfish and complemented this rich meat source with a variety of terrestrial plant foods, primarily seed-bearing grasses (Erlandson 1991a, 1991b, 1994; Glassow 1996). Relatively large middens with diverse faunal and artifact assemblages suggest a certain degree of sedentism, but a variety of smaller site types also occur in the Santa Barbara Channel region. Many of the archaeological sites dating to this period are located some distance from the present-day coast and are dominated by groundstone tools (manos and metates), thus indicating an emphasis on terrestrial plant resources (Jones et al. 2002). Other sites of this age are found near these large coastal estuaries and the

CULTURAL CONTEXT

85

associated wetlands that once punctuated the Southern California coast (Erlandson 1994). Larger numbers of sites dating to the Middle Holocene (7,000–3,350 BP) suggest that population densities in the Santa Barbara Channel region were higher relative to the Early Holocene. Many of the estuaries that punctuated the mainland coast were filled with silt by 6,000 years ago, and there was an associated shift in human settlement toward other habitat types (Erlandson 1997a, 1997b; Glassow 1997). More diverse subsistence strategies are also evident regionally, including a gradual increase in the dietary importance of fish (Glassow 1997; Vellanoweth et al. 2000). New technologies were developed or introduced into the region during this period, including mortars and pestles (~6,000–5,000 BP) and contracting-stem dart points (Erlandson 1997a). Glassow et al. (1988) have argued that the increasing importance of mortars and pestles during the Middle Holocene indicates a growing dependence on acorns as a subsistence staple. Interestingly, there is evidence for a considerable amount of variation in subsistence strategies throughout the Santa Barbara Channel region during this time (Erlandson 1997a; Glassow 1997). The Late Holocene (~3,350–200 BP) is marked by a rise in importance of marine resources, particularly more intensive use of fish from a variety of coastal habitats. Specialized tools for exploiting the marine environment were developed, including the plank canoe (~1,500 BP; Arnold 1995; Gamble 2002; Hudson et al. 1978) and more specialized fishing equipment (Glassow 1977; Rick et al. 2002; Salls 1988). The advent of the plank canoe promoted cross-channel exchange and more intensive and extensive use of pelagic fish and provided the foundation for the possible control of exchange items by elites (Arnold 1991, 1995; Glassow 1992). Introduction of the bow and arrow (~1,500 BP) was associated with a marked increase in intervillage and intertribal warfare (Lambert 1994; Lambert and Walker 1991; Walker 1989), a pattern that continued to some extent until European contact. These cultural developments culminated after 650 BP with the emergence of more complex sociopolitical and economic structures (Arnold 1987). Coastal pit-house villages with large and dense residential middens suggest a higher degree of sedentism. Faunal assemblages at these sites, as well as osteological and isotopic evidence, indicate that subsistence pursuits were diverse, but included more intensive use of pelagic fish and marine mammals (Colten 1994; Glassow 1993a; Goldberg 1993; Walker 1986; Walker and DeNiro 1986; Walker and Erlandson 1986; Walker and Hollimon 1989). Intervillage exchange networks allowed for the

86

CULTURAL CONTEXT

distribution of resources from different environmental zones (Gamble 1991; King 1976). Craft specialization appeared (Arnold 1985, 1987, 1990a, 1991; Arnold and Munns 1994) and shell beads became a medium of exchange and a means for accumulating wealth. The unequal distribution of grave goods in large cemeteries dating to this period suggests that stratified social hierarchies were well established by this time (Gamble et al. 2001; King 1990; L. King 1969, 1982; Martz 1984, 1992).

development of sociopolitical complexity There is considerable disagreement about the timing and nature of the shift toward greater sociopolitical complexity in the Santa Barbara Channel region. This is due, in part, to the difficulties with identifying sociopolitical complexity in the archaeological record (Chapman et al. 1981). Much of our knowledge regarding prehistoric developments in sociopolitical complexity in the region is based on the analysis of burial associations in island and coastal cemeteries dating to various intervals during the Holocene. Based primarily on burial lots from different parts of the Santa Barbara Channel region, King (1990) interpreted Chumash prehistory as a gradual shift of social systems from egalitarian to nonegalitarian forms. Gradual intensification of the economic system resulted from the maximizing benefits of exchanging food and nonfood items between different environmental zones (King 1976). He suggested that the organizational change to nonegalitarian society occurred at the end of the Early Period (~2,550 BP) and that religious and hereditary political leaders emerged by the beginning of the Middle Period. King (1990) argued that by the end of the Middle Period (~800 BP), economic interrelationships between people living in Chumash territory existed with little influence by hereditary political leaders. This market economy was characterized by individual entrepreneurs and the use of shell beads as a medium of exchange. He pointed out that political leaders throughout California were responsible for community stores of food and redistribution during festivals and times of need. Citing archaeological evidence for ceremonial structures, mortuary practices, and artifact types, he suggested that these institutions date back to the early Middle Period but became separated from the market economy toward the end of the Middle Period (King 1990). Arnold (1987, 1991, 1992a, 1992b, 2001) has argued that the emergence of hereditary social ranking occurred at the beginning of the Late Period (~650–200 BP; AD 1300–1782), much later than suggested by

CULTURAL CONTEXT

87

King (1990). She views this as a punctuated event triggered by environmental deterioration, particularly ocean warming that resulted in decreased marine productivity between 800 and 650 BP (AD 1150 and 1300). Arnold’s work focuses on Santa Cruz Island and the role of craft specialization and labor control in the development of sociopolitical complexity (Arnold 1987). She argued that Islanders, in the only part of the Channel Islands and mainland region where the necessary chert and Olivella shell resources occurred in juxtaposition, became specialists in shell bead money production. Control over the bead money production system provided island canoe owners, traders and leaders with the leverage they needed to procure mainland food and goods. Individuals who commissioned or controlled the labor of money-making and canoe-building specialists possibly manipulated these valuable products in order to rise in power and wealth. Thus we may see the origins of chiefly positions, societal ranking, and new levels of organization in the Chumash case. (Arnold 1987, 11–12)

Much of our information regarding the emergence of social ranking in the Santa Barbara Channel region is based on burial data. King’s (1990) assessment of status positions developing near the end of the Early Period is based on differential distribution of grave goods in a limited number of burials from Santa Cruz Island dating to this time period. His interpretation that this represents the development of social inequality has been met with considerable skepticism and criticism (Arnold 1992a). Arnold sums up this criticism by writing, He bases the appearance of Chumash stratification entirely on one Ez site on Santa Cruz Island, in which “five . . . of the burials had far more artifacts indicating wealth than any of the other burials in the cemetery” (King 1990, 95). These five burials yielded more clam beads than others in the sample, but eight other burials that he does not discuss yield more beads than three of the five “wealthy” individuals. . . . King has based wide-ranging inferences about social inequality on an inadequate sample, which further, has never been dated. (1992a, 28)

Arnold (1987) emphasized interpretations by L. King (1969, 1982) that the Late Period economic system was controlled by political elites. This is based on analysis of mortuary data from Medea Creek located in the Santa Monica Mountains. C. D. King (1990) interpreted the same burial pattern as representing a secular economic system and a reduction in the power of elite individuals in economic affairs. In his own defense to L. King (1982), C. D. King wrote, I never meant to imply that the development of economic systems which operate relatively independently of political systems result in equal distributions

88

CULTURAL CONTEXT

of wealth. In our society, the presence of a well-developed economic system has increased the ability of individuals with wealth to realize greater economic benefits than individuals with no wealth. Because control of community stores was a function of chiefs, they were by definition usually wealthy. The distribution of wealth at the Medea Creek and other Chumash cemeteries indicates that more wealth was placed with individuals from chiefly families during every time period. The presence of several protohistoric bead types only with chiefly burials may indicate the presence of types of economic interactions restricted to chiefly families. (1990, xviii)

Based on the presence of beads with almost all of the burials at Medea Creek, King (1990) suggested that all individuals participated in a market economy during the Late Period. Indeed, work by Arnold and Munns (1994) suggests that bead production on the islands was the result of independent household production. This suggests heterarchical rather than hierarchical relationships in the historic period of Chumash society (Johnson 1988, 2000). It is clear that political leaders existed in Chumash society during the Late Period. It is the influence these political leaders exerted on the economic system that is in question. But what is the first evidence for hereditary political leadership in the Santa Barbara Channel region? Martz (1984) provides some insight into the development of the first hereditary political leaders with an analysis of mortuary data from five cemeteries in the Santa Monica Mountains. These burial data give us a glimpse of social relationships at different times during the past 2,000 years. Based on data from CA-VEN-26 (Simo’mo), near Point Mugu, Martz argued for lineage-based social ranking by ~1,300 BP (AD 700). Individuals in the status group at this site tended to be flexed, primary interments associated with a wide array of artifacts. They also appear to be spatially segregated from the rest of the burials at the site. The nonstatus group at the site was distinguished by the dearth of artifacts or preparation for burial (Martz 1984). Based on these data, Martz (1984) suggested “. . . that the wealth only group derived their wealth and status from economic activities. Wealth appears to be a product of, rather than a basis for, status (emphasis added).” Martz also argued for the decline of religious power relative to hereditary political leadership sometime after 950 BP and finds little evidence for a well-developed secular economy during the Late Period. More recent data from a cemetery in Malibu complement Martz’s findings and suggests that wealth-based differences in status were in place by the late Middle Period (Gamble et al. 2001).

CULTURAL CONTEXT

89

punctuated cultural development Added debate has focused on cultural changes that occurred in the Santa Barbara Channel region after 800 BP (Arnold 1992a; Colten 1995; Jones and Kennett 1999; Kennett and Kennett 2000; Raab and Larson 1997). Arnold (1992a, 1997, 2001) and Raab (1994; Raab and Larson 1997) argued for the emergence of a more complex sociopolitical organization after this time. Both propose punctuated cultural change in the Late Holocene but define the social and natural environments of these cultural developments differently. Arnold (1992a, 1993, 1995, 2001) argued that elites emerged during the Middle to Late Period transition (800–650 BP; AD 1150–1300) to manage economic activities in the region, particularly the production of bead money on the islands and the transport of food and nonfood items across the channel. Based on Pisias’ (1978) sea-surface temperature curve, and independent archaeological data (Arnold and Tissot 1993), Arnold suggested that decreases in subsistence resources associated with reduced marine productivity and drought provided the environmental context that allowed elites to control economic activities in the region. Arnold (1992a) argued that people living on Santa Cruz Island increased production of nonfood items, particularly beads, to exchange for food with people living on the mainland coast. Elites emerged and controlled a chert source on eastern Santa Cruz Island used to produce microdrills, an important component of bead manufacturing on the island. Arnold (1995) also argued that elites owned plank canoes and therefore controlled aspects of cross-channel exchange. In contrast, Raab (1994; also see Yatsko 2000) has argued that diminishing water supplies associated with widespread drought before and during the Middle to Late Period transition prompted: (1) the consolidation of Santa Barbara Channel populations around perennial water sources; (2) solidification of territorial boundaries; (3) increased violent competition for scarce resources; and (4) the emergence of a hierarchical social system after 650 BP. Citing evidence from SBA-1731 for a relatively productive marine system (Erlandson 1993c), Raab (1994; Raab and Larson 1997; also see Jones et al. 1999) argued that marine productivity did not decrease between 800 and 650 BP and that settlement shifts evident on Santa Cruz Island during this interval are better explained as responses to fluctuations in the terrestrial, rather than the marine, environment. Based on tree-ring analysis (Larson and Michaelson 1989), Raab and Larson (1997) emphasized the importance

90

CULTURAL CONTEXT

of widespread cultural change and drought conditions across western North America between 800 and 650 BP. Raab (1994) proposed environmental and social circumscription, citing settlement shifts to perennial sources of water in northern San Diego County between 800–650 BP (True 1990). Competition for perennial water solidified territorial boundaries, and trespassing was met with lethal or sublethal retaliation. Raab (1994) hypothesized that groups of people will break up if the costs of living together exceed the benefits. He argued that warfare restricts dissolution of settlements and that inter- and intragroup competition and alliance formation provides a form of mutualism where individuals and communities benefit from mutual military support and trade. The competitive nature of this context promotes the emergence of highly competitive and elite individuals, while people generally benefit from the protection of larger groups.

Summary Much of the debate regarding the Late Holocene emergence of more complex sociopolitical organization in the Santa Barbara Channel region has focused on the last 800 years of Chumash prehistory. Arnold (1991, 1992a, 1995, 1997, 2001) and Raab (1994; Raab and Larson 1997) both argued for punctuated shifts in sociopolitical organization associated with environmental changes. Arnold (1991, 1992a, 1995) highlighted evidence for lowered marine productivity and elite control of labor on Santa Cruz Island, whereas Raab (1994; Raab and Larson 1997) emphasized lowered terrestrial productivity and proposed that competition for limited water supplies on the mainland coast caused these developments. In the following chapters, the emergence of several hallmark behaviors related to Chumash sociopolitical complexity are explored on the northern Channel Islands—sedentism, intensive fishing, production of trade items, exchange, and the development of social hierarchies. The emergence of these behaviors is considered in light of new paleoclimatic records for the region (Kennett and Kennett 2000) and archaeological data from Santa Cruz, Santa Rosa, and San Miguel islands. The next chapter considers the archaeological and ethnohistoric record of these hallmarks on the islands during the historic period and thus provides a point of departure for exploring the emergence of these behaviors in the prehistoric past.

chapter 5

Historic Island Communities

Chumash consultants in the late nineteenth and early twentieth century named roughly 22 commun~ities on the northern Channel Islands that were occupied well into the historic period (Johnson 1982, 1993, 1999, 2001). Locational information for many of these villages is clear, and archaeological sites substantiate their existence (Arnold 1990b; Johnson 1982, 1999; Kennett et al. 2000; Kroeber 1925). Johnson (1982, 1993, 1999) has done the most comprehensive work on the historic geography of these islands, synthesizing information from ethnohistoric accounts, mission records, and published archaeological work. The earliest and best account of Chumash towns on the islands was provided by Juan Estevan Pico (Heizer 1955), a Chumash speaker who lived in the San Buenaventura Indian community during the late 1800s. His list of island Chumash villages, prepared and signed on November 21, 1884 (Johnson 1999), provides the locational information that is compared here against the archaeological record. Archaeological evidence for the presence and relative permanence of historic village locations on the northern Channel Islands comes from (1) the large size and depth of these deposits, (2) the presence of substantial domestic features (house depressions and berms), and (3) diverse faunal and artifact assemblages. The best archaeological indicators of historic Chumash occupation on the northern Channel Islands are glass trade beads and metal tools (e.g., scissors, needles, knives). All of these artifacts were introduced by the Spanish during the mission period (AD 1782–1825) and were important trade items throughout 91

92

HISTORIC ISLAND COMMUNITIES

Chumash territory. Needle-drilled shell beads (Olivella disks) were also produced on the islands during the mission period and are excellent indicators of historic period occupation. These materials are often found at sites with distinct late Middle and Late Period artifact assemblages, suggesting continuity at some residential locations starting as early as 1,300 years ago (Kennett 1998). Documenting the location and character of historically occupied sites provides a touchstone between ethnohistoric accounts of social, economic, and political complexity and the archaeological evidence for these patterns of behavior. This serves as a base of comparison for prehistoric patterns extending back into the Holocene. In addition, the historic geography of these islands has implications for our understanding of the character of social, political, and economic complexity at the time of European contact. The Chumash are often described as one of the most socially and politically complex hunting-gathering-fishing societies in North America (e.g., Arnold 1991, 1992a, 1993, 1996; Colten 1993; Erlandson et al. 1997; Gamble 1991; Kennett and Conlee 2002). As discussed in the previous chapter, the prominence of large coastal villages along the Santa Barbara mainland caught the attention of early Spanish explorers passing through the region in the sixteenth century. Individuals at these coastal villages were described as consummate fishermen and craftspeople eager to trade with the Spaniards. These ethnohistoric accounts have caught the imagination of archaeologists working in the region, with whom it has become popular to describe the Chumash as a chiefdom (Arnold 1992a; Colten 1993; Gamble 1991; Gamble et al. 2001; King 1990). There was certainly a high degree of economic interaction between individuals in the Santa Barbara Channel region at historic contact, and the archaeological record supports a certain degree of economic specialization (Arnold 1987, 1992a; Arnold and Munns 1994). But economic specialization need not equal sociopolitical complexity. The fundamental question is, Were Chumash people organized politically above the village level (Johnson 2000)? There are a limited number of ethnohistoric accounts of male and female chiefs having some measure of political influence over other villages (Johnson 1988). However, the vagaries of the ethnohistoric record warrant skepticism, and the proposition requires independent evaluation with archaeological data. The geographic placement of historic Chumash villages along the mainland coast and offshore islands provides one potential source of data for testing this idea. This will provide a starting point for exploring changes in

HISTORIC ISLAND COMMUNITIES

93

settlement and subsistence that are evident in the archaeological record on the islands throughout the Holocene.

Historic Island Villages Pico’s list consists of 20 place names on the northern Channel Islands (table 10); 11 on Santa Cruz, 7 on Santa Rosa, and 2 on San Miguel (Johnson 1982, 1993, 1999). Two additional village names on Santa Rosa are identified in mission records (Johnson 1999). Of these, 11 are positively identified archaeologically, 8 with a lesser degree of certainty, and 2 have remained undetected. There are no clear indications from Pico’s list that historic settlements were present on the small island of Anacapa. The entire island was systematically surveyed in the 1970s and no evidence for historic settlement was reported (Greenwood 1978). A total of 1,270 names appear in mission baptismal registers for the northern Channel Islands, a small percentage of the precontact and protohistoric population on the islands (Johnson 1999). Among the islanders baptized, 11 chiefs are named in mission records. Island people were missionized relatively late compared with people living along the mainland coast and interior. Islanders started moving to mainland missions in 1785, but a vast majority of people converted between 1814 and 1816 with the last recorded island baptism occurring in 1822 (Johnson 1982). The underlying causes of the rapid migration of islanders to mainland missions between 1814 and 1816 were certainly complex, but they were driven, at least in part, by: (1) economic and social instability related to depopulation caused by European disease; (2) collapse of cross-channel exchange; and (3) perturbations in the marine environment and an extended period of drought (Johnson 1982, 1993; Johnson and Walker 1999; Larson et al. 1994).

Santa Cruz Island Johnson (1982, 1993, 1999) and Arnold (1990b) provide the most recent and definitive statements regarding settlement locations on Santa Cruz during the historic period (table 10; map 6). Pico identified 11 village names on the island, and the existence of each community is corroborated by mission records. Six of these communities are identified archaeologically with a great degree of certainty (Arnold 1990b;

table 10. Ethnohistorical and archaeological information for historic period communities on the northern Channel Islands Ethnohistoric Record Historic Village Santa Cruz Island Kaxas Mashchal L’alale L’akayamu Ch’oloshush Shawa Liyam Nanawani Swaxil Lu’upsh Santa Rosa Island He’lewaskuy Qshiwqshiw Hichimin Silimihi Niaqla Nimkilkil Nawani Nilal’uy San Miguel Island Tuqan Niwoyomi

Archaeological Record Size (m2)

House Pits

No No No No No No No No Yes No

6,000 9,000 ? 8,600 5,400 5,600 6,580 3,800 6,300 3,000

? 11 8 (?) 19–20 15 11 ? 15 ? 6

No Yes Yes Yes Yes No No Yes

No No No Yes Yes Yes Yes No

6,600 ? 11,200 12,000 50,000 28,000 9,100 3,016

12 8 12 24 70 25 19 12

No No

Yes Yes

4,000 3,200

6 4

Pico

Baptisms

No. of Chiefs

Site No.

Artifacts

El Puerto Principal El Dirección al Oeste Punta del Diablo Mas al Oeste En Dirección Sudoeste En Dirección al Sur En Dirección al Este Mas al Este A la Punta del Este En Dirección al Norte

129 69 5 50 28 9 117 61 205 63

1 1 0 0 0 0 1 0 1 0

SCrI-240 SCrI-434 Unknown SCrI-328 (330) SCrI-236 SCrI-192 SCrI-1 SCrI-504(506) SCrI-423 SCrI-306*

Yes Yes No Yes Yes Yes Yes Yes Yes Yes

Santa Rosa Island Rancho Viejo El Puerto En Dirección al Oeste Más al Oeste Más al Oeste En Dirección el Sur Más al Sur

37 119 71 53 10 51 2 48

0 2 1 0 0 0 0 1

SRI-436 SRI-87 SRI-60 SRI-40 (502) SRI-2 SRI-15 SRI-97 (98) SRI-62

34 3

1 0

SMI-161 SMI-470

San Miguel Island ?

sources: Arnold 1990b; Johnson 1982, 1993; Kennett 1998; Kennett et al. 2000. *Plus five sites in the vicinity of SCrI-306 (see Arnold 1990b)

14

C

map 6. Map of the northern Channel Islands showing known historic period Chumash communities (Arnold 1990b; Johnson 1982, 1993, Kennett et al. 2000).

96

HISTORIC ISLAND COMMUNITIES

Johnson 1982, 1993, 1999). The true locations of L’alale (Lalale), Shawa (Chahua), Swaxil (Yshguagel), and Nanawani (Nanaguani) are less certain. Johnson (1982, 1993) placed the small village of L’alale in the vicinity of East Diablo Point on the north shore of the island, possibly SCrI-436 (Fry’s Harbor), but no mission period artifacts are known from the site. Glass trade beads and needle-drilled Olivella wall beads were unearthed during recent excavations at SCrI-192 at Morse Point, and this is generally considered to be the community of Shawa (Arnold 1990b; Johnson 1993). One of the most perplexing problems in identifying historic villages on Santa Cruz is locating the important community of Swaxil on the eastern end of the island. Pico associated two historic village names with this part of the island: Swaxil and Nanawani. The geographic position of these villages is confirmed with marriage register data (Johnson 1982, 1993). Pico placed Swaxil at the eastern end of the island (a la punta del este), and Fernando Librado, another Chumash consultant whose mother was from this village, placed it more specifically at Scorpion Anchorage. Fernando also visited Scorpion Anchorage when he worked as a sheepshearer for the ranching operation on the island (Johnson 1982). Pico placed the village of Nanawani between Liyam (Liam) and Swaxil on the eastern point of the island. Two hundred and five baptisms are reported from Swaxil, the largest number of baptisms from a single village on the northern Channel Islands (Johnson 1999). A mission census of island village populations in 1804 listed 145 adults at Swaxil, and, based on a formula for reconstructed population decline, Johnson (1982) estimated that 262 people were living at this location in 1782. One of the four named chiefs from Santa Cruz was also from Swaxil, suggesting that this community also had a significant amount of political influence. Nanawani was smaller than Swaxil and moderate in size by island standards. Johnson (1982, 1993) has documented 60 people from Nanawani in baptismal records at Mission San Buenaventura and one from Mission Santa Barbara. Johnson (1982) originally associated Swaxil with Scorpion Anchorage. This was based on ethnohistoric accounts and R. L. Olson’s description of a remnant midden deposit at the mouth of the drainage (SCrI-423 and 507; Olson, 1928). Olson does not mention historic artifacts at this location (designated at the time as SCrI-B-141), but described late prehistoric artifacts that are often found at historic locations. Johnson (1982) hypothesized that a majority of the site was destroyed by coastal erosion, flooding and historic disturbances associated with the ranch complex.

HISTORIC ISLAND COMMUNITIES

97

Johnson (1982, 1993) placed Nanawani in the Smugglers Cove area, associated with two archaeological sites showing clear evidence for historic occupation (SCrI-504 and SCrI-506; Hoover 1971; Olson, 1928; Schumacher 1875). In 1990, Jeanne Arnold visited eastern Santa Cruz Island to document the locations of Nanawani and Swaxil (Arnold 1990b). Her investigation was limited because of a complicated joint ownership agreement between the National Park Service and Francis Gherini, the former part-owner. Nonetheless, Arnold (1990b) visited both Smugglers Cove and Scorpion Anchorage, documenting historic period artifacts at Smugglers Point (CA-SCrI-506). She also identified substantial Middle and Late Period components at Scorpion Anchorage, but found no historic material at this location. Based on the absence of historic artifacts at Scorpion, Olson’s account of house depressions and glass trade beads at Smugglers Cove (CA-SCrI-504), and a clear historic period component at Smugglers Point, Arnold (1990) placed Swaxil at Smugglers Cove (CA-SCrI-504 and 505) and Nanawani at Smugglers Point (CA-SCrI-506). Glass trade beads and needle-drilled wall beads, however, were later discovered at Scorpion Anchorage (SCrI-423; Kennett et al. 2000). Thus, two primary loci of historic settlement are clearly delineated archaeologically on eastern Santa Cruz Island, one at Scorpion Anchorage and the other at Smugglers Cove. The presence of a historic period component at the mouth of Scorpion Anchorage substantiates Fernando Librado’s claim for Swaxil’s existence at this location. However, the extent of this component is difficult to assess due to natural disturbances and destruction related to ranch activities (Kennett et al. 2000). Historic Chumash occupation is best preserved on the north side of the creek (CA-SCrI-423), but it is likely that it stretched across the mouth of the canyon.

Santa Rosa Island Pico identified seven villages on Santa Rosa, anchored by two large, politically important, villages on the eastern end of the island: Qshiwqshiw and Hichimin (table 10, map 6). Two additional village names associated with Santa Rosa, He’lewashkuy and Xonashup, are identified in mission records but were not included in Pico’s original list (Johnson 1999). Solid archaeological evidence for five of these villages exists and

98

HISTORIC ISLAND COMMUNITIES

the other three locations are tentative designations based on late prehistoric components. Little is known about Xonashup, and it is not included in this discussion. A total of six chiefs are named in mission records for Santa Rosa Island, four at Qshiwqshiw, one at Hichimin, and one at Nilal’uy (Johnson 1982). Two additional chiefs were named in association with Hichimin but are not confirmed in mission records (Brown 1967; Johnson 1982).

hichimin (cheumen) A total of 71 baptisms are recorded for the community of Hichimin, including one person, Angel Alaya, identified as a village chief (Johnson 1982, 1999). Marriage ties for this village include four with other communities on Santa Rosa; and two with people from L’akayamu, located on the west end of Santa Cruz. Pico associated the island Chumash village of Hichimin with the port at Bechers Bay, making it the most securely placed village on the island. A large archaeological site (SRI-60) exists at the mouth of Ranch House Canyon, just to the north of the historic pier on the island. Johnson (1982, 1993) confidently placed Hichimin at this location based on material recovered during excavations of SRI-60 by Steven Bowers in 1876 (Benson 1982), Philip Mills Jones in 1901 (Heizer and Elsasser 1956), and David Banks Rogers in 1920 (Rogers 1929). Glass beads, broken glass, and brass and copper artifacts were found associated with burials at the site. Late prehistoric materials, including Olivella callus-cup beads and triangular microblades, are present across the surface of SRI-60. Cobalt blue glass trade beads, confirming a historic period occupation, were recovered from the upper levels of a small excavation unit taken from the seacliff at SRI-60 (Kennett 1998). Significant amounts of site disturbance (ranch activity, looters’ pits, and old excavation units) are evident, and large portions of the site continue to erode into the sea. Several house depressions are visible at the site, but most are obscured by looters’ pits and backdirt piles. Nevertheless, the deposits are impressive and are undoubtedly the remnants of Hichimin.

qshiwqshiw (siucsiu) One hundred and nineteen baptisms appear in mission records for Qshiwqshiw (Johnson 1982). Pico placed this large village at the mouth of Old Ranch Canyon, the largest drainage on the island and the only

HISTORIC ISLAND COMMUNITIES

99

one with a substantial estuary at its mouth. Mission records indicate that four chiefs lived in this community: Apolonio Ssetey, Marin Geele, Mitrio Ayuuanatset, and Damian Yaquinunaitset; and marriage linkages include: one with Tuqan on San Miguel, four with villages on Santa Rosa, three with villages on the south side of Santa Cruz Island, and two with the mainland coast (Johnson 1982, 151–152). Johnson (1982) suggested that Pico’s locational information for this village was incorrect based on Orr’s description of CA-SRI-85, a Middle and Late Period site at the mouth of Old Ranch Canyon. Instead, Johnson (1982) suggested that the village was located northwest of Old Ranch Canyon at Southeast Anchorage, within the broader limits of Bechers Bay. Orr (1968) describes a relatively large Late Period village (SRI-77) with 10 house depressions at this locality and implied that a historic period component existed. Work on the eastern end of Santa Rosa Island in the vicinity of Old Ranch Canyon suggests that Pico’s original village location may be correct (Kennett 1998). During the National Park’s coastal survey, glass trade beads (N  4) were found on the surface of CA-SRI-87, a site at the mouth of Old Ranch Canyon (Don Morris, unpublished site record). Additional glass beads and several needle-drilled wall beads were recovered at the site during subsequent visits (Kennett 1998). CASRI-87 is a large site with eight visible house depressions and the possible remnants of many others. When combined with SRI-88, SRI-85, and SRI-84, other Late Period sites on the margin of the estuary, it is the largest late prehistoric site complex on Santa Rosa Island.

silimihi (silimi) This site is described by Pico as the first village west of Hichimin on the north shore of Santa Rosa Island. Although Silimihi has a comparable number of baptisms (N  53) to Hichimin, no chiefs were identified among the individuals listed. Marriage linkages to other villages include: six to Santa Rosa Island, one to Santa Cruz (Ch’oloshush), and one to the mainland coast (Johnson 1982). Johnson (1982, 1993) and others (Brown 1967; Kroeber 1925; Orr 1968) place this village at the mouth of Cañada Verde, where a large village with house depressions exists (SRI-40). This site was excavated by Steven Bowers in 1876 and Philip Mills Jones in 1901 (Benson 1982; Heizer and Elsasser 1956). Jones focused his attention on the cemetery at the site and recorded the presence of glass trade beads.

100

HISTORIC ISLAND COMMUNITIES

The coastline between Bechers Bay and Cañada Verde was recently surveyed and no other candidates for this village were located (Don Morris, unpublished site records). CA-SRI-40, and a newly designated portion of the site (CA-SRI-502; Don Morris, unpublished site record), is a relatively large site complex with 24 visible house depressions. Orr (1968, 177) reported the presence of 66 house depressions but included many features that are clearly cow wallows on the eastern side of the site. Two cobalt blue glass beads were found on the surface of the eastern portion of the site (CA-SRI-502) during systematic surface collections, along with large quantities of Middle and Late, as well as protohistoric, Period material. Two additional glass beads were found during limited excavations in the western part of the site (upper 10 cm).

niaqla (niacla) According to Pico this community was on the north coast of the island, west of Silimihi (Heizer 1955). Only 10 baptisms are recorded for this village and no chiefs are identified in the existing records. Marriage linkages are also minimal; two with Silimihi and one with Ch’oloshush on Santa Cruz. Johnson (1982, 1993) associates Niaqla with SRI-2, a large pit-house-village site located at the mouth of Skull Gulch. This site was excavated by Phil Orr in the 1940s and 1950s (Orr 1968), and his notes indicate the presence of glass beads. Other candidates for this village, between Cañada Verde and Skull Gulch, were not identified in an intensive coastal survey (Don Morris, unpublished site records). Orr (1968, 189) describes CA-SRI-2 as a large village with 70 house depressions. Today, 15 large house depressions are visible on the eastern side of the gulch and many of the house depressions on the western side of the site are obscured by Orr’s extensive archaeological excavations. Review of Orr’s collection at the Santa Barbara Museum of Natural History revealed nine glass beads and eight needle-drilled Olivella shell wall beads from excavations in whale house, located in the western portion of the site (Kennett 1998). Although extensive excavations occurred across the site, the only other historic artifact in the collection is a perforated metal disk collected from somewhere within the site boundary. Orr (1968) also published a radiocarbon date of 330  50 (UCLA-0134) on charcoal collected from a hearth located 56 cm below the surface and a date of 400  80 on wood collected from turtle house (House 1) in the eastern portion of

HISTORIC ISLAND COMMUNITIES

101

the site. When calibrated, both radiocarbon dates fall within the early part of the historic period (see Kennett 1998).

nimkilkil (nimquelquel) According to Pico, Nimkilkil was the westernmost community on Santa Rosa Island. There were no chiefs recorded in residence at this village, and marriage linkages were isolated to villages on San Miguel [Tuqan (Toan)] and Santa Rosa, with the exception of two reported marriages to L’akayamu on Santa Cruz (Johnson 1982). Johnson (1982, 1993) suggests that SRI-15 on the west end of the island at Abalone Point is the most likely candidate for Nimkilkil. The stretch of coast between Skull Gulch and SRI-15 was systematically surveyed, and other large late prehistoric sites were not identified (Don Morris, unpublished site records). SRI-15 is a large, complex site that was occupied, at least intermittently, for the past 5,000 years. Orr’s field notes describe 25 house depressions on the eastern side of Abalone Point, but far fewer are visible today. There are also features that look suspiciously like domestic features on the western flank of the point that date to the late Middle Period (Kennett 1998). Surface collection on the eastern side of the site produced a limited array of late prehistoric material that did not produce definitive evidence for occupation historically. A 25  25 cm column was also excavated in this area and two radiocarbon dates place this locus firmly in the Late Period, but the historic period deposits remain elusive (Kennett 1998).

nawani (nahuani) South of Nimkilkil, on the southwestern coast of the island, Pico placed the settlement of Nawani. Only two baptisms are documented from this community and no marriage linkages are reported. Kroeber (1925) put this village at Bee Rock, probably based on excavations undertaken at CA-SRI-31 by Phillip Mills Jones. Orr (1968) argued that the probable location of this village was at China Point where two large village sites, with well-defined pit houses, occur in close proximity (SRI-97 and -98). Orr (1968) excavated a trench in SRI-97 but does not provide information about the artifact assemblage. The stretch of coast between Abalone Point and South Point is now surveyed (Don Morris, unpublished site reports), and three potential candidates for Nawani exist: China Point (SRI-97, -98), Bee

102

HISTORIC ISLAND COMMUNITIES

Rock (SRI-31), and a Late Period site to the northwest of Bee Rock (SRI-339). Surface collections were conducted at all three sites and small-scale testing was done at CA-SRI-31 and CA-SRI-97 (Kennett 1998). The Late Period site northwest of Bee Rock (SRI-339) is situated within an extensive dune field. Late Period triangular microblades, callus-cup beads, and small, concave base arrow points were found interspersed between the shifting dunes, but no historic material was found. Remnants of domestic features (house depressions), common at historic sites, were also not identified, and it seems unlikely that this is the site of Nawani. Farther south along this stretch of coast, near Bee Rock, a small number of subtle house depressions are visible on the surface of CASRI-31, the site of Nawani proposed by Kroeber (1925). These depressions are partially obscured by sand and surface collections were only possible near the cliff face. Two trapezoidal microblades, two triangular microblades, and a number of Olivella shell wall beads were found along the cliff edge, suggesting a late Middle and Late Period occupation. Two small column samples (25  25 cm) were excavated at the seacliff without encountering historic materials, and two radiocarbon dates at the site place the visible deposits firmly in the late Middle Period (Kennett 1998). Review of the Jones collection from CA-SRI-31 also suggests a late Middle Period occupation (Heizer and Elsasser 1956, plate 19). The complex of sites at China Point, to the southeast of Bee Rock, are impressive and the most probable location of Nawani (SRI-97, -98). Surface collections at this site complex revealed large amounts of Late Period material (Kennett 1998). Historic artifacts were not discovered during the excavation of a small column sample on the western edge of the site, but two radiocarbon dates suggest persistent use of this location into the historic period (Kennett 1998).

nilal’uy (nilalui) Nilal’uy is described by Pico as a village located on the south side of Santa Rosa. Johnson (1982, 1993, 1999) reports 48 baptisms from this community and one person, Santos Chacú, as a village chief. Marriage alliances were recorded with one village on San Miguel [Niwoyomi (Niuoiomi)], three villages on Santa Rosa, two villages on Santa Cruz and one on the mainland coast. Johnson (1982, 1993, 1999) associated this community with SRI-62, a site located east of South Point in the

HISTORIC ISLAND COMMUNITIES

103

shelter of Johnson’s Lee. This was based on early reports of glass beads being discovered at the site in the 1920s (Rogers 1929). The U.S. Air Force bulldozed the site in the 1940s and salvage operations conducted by Phil Orr failed to produce historic material (Orr 1968). Very little of the site is intact and no diagnostic artifacts were identified during recent surface collections. Although Rogers reported 12 house rings at this location, they are not evident today. A column sample excavated from the seacliff at this location revealed deposits that date much earlier in time (Kennett 1998; ~2,400 BP). Therefore, the only evidence for historic period occupation comes from Rogers’ mention of glass beads.

he’lewashkuy (elehuascui) Thirty-seven baptisms are recorded from the village of He’lewashkuy. This community is noteworthy owing to its high incidence of endogamous marriages (N  8), uncommon elsewhere in Chumash territory (Johnson 1982). Only three exogamous marriages are recorded: two with people from the west end of Santa Cruz and one to an unspecified village on Santa Rosa. Johnson (1982) tentatively suggested that He’lewashkuy was located on the east end of Santa Rosa, based on intervillage relationships recorded in mission records. Fernando Librado told Harrington that he did not recognize the village name, and it is absent from Juan Estevan Pico’s list of Santa Rosa Island villages. However, it is described in mission records as an island village (Johnson 1993). Johnson (1982, 1993) originally placed He’lewashkuy south of Skunk Point at the mouth of Old Ranch Canyon. However, the complex of sites at this location is most likely the large community of Qshiwqshiw. Three candidates exist for this village on the southeast shore of Santa Rosa: SRI-130, at the mouth of Jolla Vieja Canyon; SRI-432, just north of Ford Point; and SRI-436, at the mouth of San Augustine Canyon. All of these sites have visible house depressions and Late Period artifact assemblages, but historic artifacts were not found. A column sample at SRI-130 revealed trapezoidal microblades that suggest a late Middle Period occupation, and radiocarbon dates substantiate this interpretation (Kennett 1998). Work at SRI-432 and SRI-436 has revealed clear Late Period components at both sites. The most impressive late prehistoric material comes from SRI-436, and two radiocarbon dates place it in the Late Period (Kennett 1998). This site appears to be the most likely candidate for He’lewashkuy.

104

HISTORIC ISLAND COMMUNITIES

San Miguel Island Two village names are associated with San Miguel: Tuqan (Toan) and Niwoyomi (Niuoiomi). The village name Tuqan is linked with Cuyler Harbor on the eastern end of the island and 34 baptisms are recorded for this community (Johnson 1982, 1993, 1999). One chief, Cristoval Mascál, was identified from the village along with extensive marriage linkages with Santa Rosa, a single marriage to the village of L’alale on Santa Cruz, and three marriages to villages on the mainland (Johnson 1982). Johnson (1982) suggests that Niwoyomi (three baptisms) was a satellite community allied with the larger village of Tuqan. Johnson (1993) placed Tuqan at Cuyler Harbor based on early excavations in the vicinity (Benson 1982; Dall 1874). The best candidate for this site is currently SMI-163, a pit-house-village site located on the eastern edge of Cuyler Harbor (Kennett 1998). Six house depressions are evident at this location and Late Period beads and microblades are present on its surface. A radiocarbon date from the upper 10 cm of the deposit place it in the protohistoric period (Kennett 1998). The best candidate for Niwoyomi is currently SMI-470, a small village site located on the western side of the island next to Otter Creek (Kennett 1998). Much of the deposit is covered with dune sand, but several house depressions are visible on the surface of the site. Shell midden deposits exposed in the seacliff and creek bed contain a rich array of faunal material, and the artifacts clearly place it in the late prehistoric period. Two radiocarbon samples, one from the cliff face and one from the creek bank, both yielded calibrated radiocarbon dates in the early 1800s (Kennett 1998). This is currently the most likely candidate for Niwoyomi, but work on the western tip of the island near Point Bennett has revealed another protohistoric site (Walker et al. 2000). This raises the possibility of a third, previously unknown, center of historic occupation and another potential candidate for Niwoyomi.

Geographic Analysis The ethnohistoric record provides a synchronic view of Chumash society and the implications of these accounts for sociopolitical organization must be tested, not assumed. It is possible that the settlement geography of island villages provides insight into the degree of political cohesiveness between and among villages. The historic village locations described by Juan Estevan Pico and subsequent field truthing by

HISTORIC ISLAND COMMUNITIES

105

archaeologists provide an ideal situation for scrutinizing ethnohistoric accounts of sociopolitical integration. Below, Johnson’s (1993) use of Measured Geographic Centrality is discussed as a means of exploring the connectivity between villages on the northern Channel Islands. Johnson’s study is a point of departure for discussing the intervisibility of Chumash villages (viewshed analysis) and how the distribution of Chumash village sizes compares with the rank-size rule, a simple test of political integration. Johnson (1993) has analyzed the geographic position of historic villages on the northern Channel Islands based on ethnohistoric accounts of village locations coupled with known archaeological evidence for these locations. Johnson gauged village importance by the total village population, based on baptismal records, and the number of named chiefs in residence at each village. Using the size and position of historic Chumash villages, he used Measured Geographic Centrality (ALLOC program, see Johnson 1993) at different geographic scales to determine natural political and economic centers. The results of this study suggest that the village of Kaxas (Cajats) on Santa Cruz is the natural “optimal” center for the entire island chain. Kaxas was the second largest village on Santa Cruz and had one named chief in residence. Johnson (1993, 29) also divided the northern Channel Islands into two “natural” regions—one region consisting of all the historic villages on San Miguel and Santa Rosa islands, but including the two westernmost villages on Santa Cruz (L’akayamu and Ch’oloshush), and the other region comprising the remaining villages on Santa Cruz. The optimal center for the San Miguel and Santa Rosa group was Qshiwqshiw, and that for the Santa Cruz group remained Xaxas. Qshiwqshiw was the largest village on Santa Rosa and had at least four chiefs in residence during the mission period. When other measures of social network centrality were incorporated into the analysis, the village of Liyam emerged as the principle center on Santa Cruz. Johnson convincingly argued that the results of this social gravity model accurately depict the location of politically and economically important villages.

Viewshed Analysis Johnson focused on how historic period villages were connected politically and economically. There were certainly island villages that were larger and more economically important than others, but this does not

106

HISTORIC ISLAND COMMUNITIES

necessarily imply hierarchical control of the entire system by the chiefs living at these villages. Johnson (2000) has more recently argued that villages in the Santa Barbara Channel region were relatively autonomous political units that were more heterarchically, rather than hierarchically, organized. Viewshed analysis of historic island communities supports this idea of village autonomy on the northern Channel Islands (map 7). Three primary patterns emerge when the viewsheds of each historic period community are evaluated. First, the people that lived in these coastal communities had clear views along the coast and out to sea. Second, these same coastal locations had limited views of interior parts of each island. Third, villages on the same island are generally positioned out of view from one another. The geographic position of historic villages on the coast was certainly related to the importance of the maritime economy. However, the even distribution of settlements around the islands, each with a clear view of the surrounding ocean suggests that people were leery of attack from unrelated people on other islands or the mainland. Indeed, most island villages dating to this time were positioned on high seacliffs or on headlands, the most defensible locations possible along the coast. Limited views across the interior portions of these islands suggest that people were less concerned with aggressive action from people living in other villages on the same island. Baptismal and marriage register data indicate that people had familial relations with those on other islands and on the mainland but were more likely to have kin in villages on the same island (Johnson 1982). Territorial boundaries on each of the islands undoubtedly were well understood or maintained without continual visual monitoring from coastal locations. The lack of intervisibility between villages on the same island during this period is particularly striking. Historic villages on Santa Rosa were not intervisible across land or water, but the village of Tuqan on San Miguel was visible from Nimkilkil and Niaqla (western Santa Rosa Island). Several villages on the eastern end of Santa Rosa and the western end of Santa Cruz were also intervisible. Even villages in close geographic proximity were not necessarily intervisible. For instance, Nimkilkil and Niaqla were located close to one another on the north coast of Santa Rosa Island, but, because of the convoluted nature of the shoreline, they were not intervisible. The only exceptions to this rule were the communities of L’akayamu and Ch’oloshush, both positioned on the western end of Santa Cruz. The people living in these two villages were within each other’s line of sight.

map 7. Viewshed analysis of historic period coastal communities. Analysis done in Arc/View 3.0.

108

HISTORIC ISLAND COMMUNITIES

Mission records indicate that individuals in different island villages were interrelated, but the lack of intervisibility between historic communities suggests that they were also relatively autonomous units. Lock and Harris (1996) note a similar pattern in Neolithic Europe. They suggest that the pattern represents a certain degree of territorial behavior and village autonomy. Historic island communities were generally positioned on headlands close to perennial water. Beach access was also of great importance for landing plank canoes (Arnold 1995). Many optimal locations exist on the islands that fulfill most of the economic parameters used by Chumash people to establish settlements. People could have lived in villages that were intervisible, and the fact that most do not appears to be an expression of village sovereignty.

Rank-Size Analysis The autonomy of island communities suggested by viewshed analysis can be explored further with rank-size analysis of island village populations estimated from baptismal records. Rank-size analysis is based on the empirical observation that the largest modern cities are usually twice the size of the second largest cities, three times larger than the third largest cities, etc. (Haggett 1971). When the ranks and sizes of modern cities are plotted logarithmically, they exhibit a lognormal distribution (figure 14). Two opposite economic forces are thought to explain this rank-size rule: diversification and unification (Zipf 1965). Diversification promotes the distribution of numerous autonomous communities near raw material sources, and the force of unification moves these raw materials to centers of production and consumption. Politically and economically integrated urbanized countries generally exhibit a lognormal distribution when the populations of cities are ranked (Adams 1981; Berry 1961; G. A. Johnson 1977). Archaeologists rarely deal with highly integrated state-level societies, therefore the logarithmic distribution of site sizes are usually compared to the lognormal distribution. Primate, convex, and primo-convex distributions are well described and provide insight into the level of political and economic integration within a society (Falconer and Savage 1995; G. A. Johnson 1977). Primate distributions occur when a smaller number than expected of intermediate and large communities are present in a region, based on the rank-size rule (Falconer and Savage 1995) (figure 14). This is the case when the biggest population center is much larger than the

HISTORIC ISLAND COMMUNITIES

109

1000 Primate observed Expected Convex observed Expected

Size

100

10

1 1

10

100

Rank figure 14. Examples of Primate and Convex ranksize distributions for comparison to curves generated for the Chumash area (data from Falconer and Savage 1995; drafted by D. Kennett).

other communities in a region, suggesting a highly centralized political and economic system (G. A. Johnson 1977; Kowalewski 1982). Artificial primate distributions are caused either by incomplete delineation of a settlement system (G. A. Johnson 1977) or by systems on the peripheries of expansive colonial empires ( Blanton 1976; Smith 1976). Convex logarithmic distributions exhibit more intermediate and large places than expected by the rank-size rule. A convex distribution generally indicates little integration between villages within a region. Classification of the Chumash as a chiefdom suggests that a certain degree of political integration existed, at least at historic contact. If a number of simple chiefdoms existed in the Santa Barbara Channel region during the historic period then something between a concave and loglinear distribution would be expected. If a paramount chiefdom, with a large power center, existed, then something close to a primate distribution would be expected. The rank-size distributions for different segments of Chumash society are presented in figure 15. Chumash villages on the northern Channel Islands, the mainland coast, and the interior were ranked according to population size based on baptismal tabulations by Johnson (1982, 1988). The Chumash people baptized at the five missions in the region were only a small fraction of the total population, but the number of baptisms provides the only reliable proxy for village sizes in the region. Rank-size distribution for all of the island villages is a clear convex curve (figure 15a). This indicates that there are more large and intermediate

110

HISTORIC ISLAND COMMUNITIES

1000

1000

Santa Cruz Is. Size (Baptisms)

Size (Baptisms)

All Islands 100

Lo

g

N

or

m

10

al

a

100

1

Lo

g

rm

al

b 1

1

1

100

10

10

Rank

Rank 1000

1000

Total Population

Santa Rosa & San Miguel Is.

Size (Baptisms)

Size (Baptisms)

No

10

100

Lo

g

10

No

rm

al

100

Lo

g

No

rm

al

10

c

d 1

1 10

1

Rank

1

10

100

Rank

figure 15. Rank-size analysis of: (a) all historic villages on the northern Channel Islands, (b) historic villages on Santa Cruz Island, (c) historic villages on Santa Rosa and San Miguel islands, and (d) historic villages throughout Chumash territory. Village sizes are based on estimates from baptismal records (Johnson 1982, 1988; drafted by D. Kennett).

communities than predicted by the rank-size rule and suggests little political integration between communities. Convexity also occurs when the historic villages on Santa Cruz Island (figure 15b) are plotted separately from those of Santa Rosa and San Miguel (figure 15c). As noted above, convex curves can result from incompletely defining a political and economic system. For this reason, the rank-size distribution for the entire Chumash region, including the islands, mainland coast, and interior, was determined. Again, the distribution is clearly convex (figure 15d). These rank-size distributions suggest that the historic period villages on the northern Channel Islands were relatively autonomous politically. These data also suggest that the economic force of diversification dominated the economic and political system. That is, relatively small autonomous villages are scattered across the landscape, maximizing access to resource locations and raw material sources. Political autonomy neither rules out economic interdependence between individuals

HISTORIC ISLAND COMMUNITIES

111

living in different villages in Chumash territory nor the periodic integration of villages into larger political units as suggested by the ethnohistoric record.

Summary Historic Chumash communities were distributed relatively evenly along the coastline of San Miguel, Santa Rosa, and Santa Cruz islands. Ethnohistoric accounts of island village locations combined with archaeological ground truthing provide great insight into these distributions. Historic island communities varied in size along with economic and political importance (Johnson 1982, 1993, 1999). Larger, more centrally located communities tended to have more chiefs in residence (Johnson 1982, 1993), but a certain degree of village autonomy is suggested by the lack of intervisibility between communities. The rank and size of communities indicate that the economic forces of diversification rather than unification dominated the regional economic and political system, but ethnohistoric accounts indicate that integration of communities periodically occurred. The historic geography of the northern Channel Islands provides a point of departure for exploring changes in settlement, subsistence, and social/ political organization during the Holocene epoch.

chapter 6

Terminal Pleistocene to Middle Holocene Records

In this chapter, I summarize the Terminal Pleistocene through Middle Holocene archaeological records from the northern Channel Islands. These records provide some of the best evidence for human occupation along the west coast of North America during the Terminal Pleistocene and Early Holocene (~13,000–7,500 BP; Erlandson 1994; Erlandson et al. 1996a, 1996b; Johnson et al. 2000; also see Porcasi et al. 1999 and Raab and Yatsko 1992 for comparable work on the Early Holocene occupation of the southern Channel Islands). Compared with mainland sites of similar age, these deposits are often stratigraphically intact and better preserved. The mere presence of people on these islands between ~13,000 and 11,500 BP provides some of the earliest evidence for boat use in the New World (Erlandson 2001, 2002b; Jones et al. 2002). These sites have also produced some of the oldest hook-and-line fishing technology (bone gorges, ~10,000 BP) and eel grass cordage/basketry on the west coast of North America (Connolly et al. 1995; Erlandson 1994, 2001; Rick et al. 2001a). In comparison to the Middle and Late Holocene, only a small number of archaeological sites date to before 8,000 BP. These sites are often ephemeral and island population levels are inferred to be relatively low during this early time. Many of the earliest sites are located on the outer islands (San Miguel and Santa Rosa), and it is possible that they were only occupied sporadically, rather than permanently, at least during the Terminal Pleistocene and into the earliest Holocene. Human settlement proliferated on the northern Channel Islands after 7,500 BP as population levels increased and people established 112

TERMINAL PLEISTOCENE TO MIDDLE HOLOCENE

113

more permanent villages in the most productive habitats on the larger islands of Santa Rosa and Santa Cruz. Primary village sites were anchored to highly productive coastal locations for much of the year, usually in close proximity to perennial drinking water (springs or streams). Small prehistoric settlements in the interiors of the larger islands, often positioned on ridges and small hills, suggest the complementary use of marine resources with interior plant foods. This dual subsistence–settlement strategy dominated and persisted on the northern Channel Islands throughout the Middle Holocene in the face of relatively large-scale climatic variations and in the context of relatively low population density. If burial data reflect the status and wealth achieved during an individual’s lifetime, then it appears that people were starting to differentiate themselves from one another during the Middle Holocene and that some individuals had achieved greater status and wealth than others. The details of these economic and social changes are presented below and will be considered from the perspective of HBE in chapter 8.

Terminal Pleistocene Record Settlement of the northern Channel Islands occurred by at least 12,300 BP (10,390  130 radiocarbon years BP; Erlandson et al. 1996a) and possibly as early as 13,500 BP (11,490  70 radiocarbon years BP; Johnson et al. 2000). The best evidence for Pleistocene age settlement comes from Daisy Cave (CA-SMI-261), a rockshelter/cave complex positioned on the northeastern shore of San Miguel Island that has basal deposits dating to between 12,300 and 11,120 BP (Stratum G; map 8, table 11). Perched on a cliff face overlooking the ocean, the early occupants of this cave had a clear view across the Santa Barbara Channel to the mainland. When the cave complex was first occupied, sea level was ~62 m below the current stand and a land bridge connected San Miguel and Santa Rosa islands (Porcasi et al. 1999). At this time, the overall size of the island was much larger than its current extent (see figure 10), particularly along the north coast of the island where the seafloor is relatively shallow. A single pine needle (P. muricata) was found in the earliest level at the site and, along with pine pollen found in the same context, suggests that the coastal plain in the vicinity of the cave was covered with forest rather than the low sage scrub and grass that predominate across the island today (West and Erlandson 1994).

map 8. Terminal Pleistocene and Early/Middle Holocene sites on the northern Channel Islands discussed in text (map produced by D. Kennett with the assistance of J. Bartruff).

table 11. Known Terminal Pleistocene and Early Holocene archaeological sites on the northern Channel Islands Site No.

Location

Description

CA-SCrI-109

Punta Arena

Shell Midden Coastal

CA-SCrI-681

Scorpion Drainage Hilltop Scorpion Drainage Hilltop Daisy Cave Bay Point

Midden Interior Midden Interior Rockshelter/Cave Shell Midden

CA-SMI-350

Simonton Cove

CA-SMI-433

Simonton Cove

CA-SMI-438

Simonton Cove

CA-SMI-603

Cave of the Chimneys Bay Point

Noncultural? Rocky Coast Shell Midden Rocky Coast Shell Midden Rocky Coast Rockshelter Shell Midden Rocky Coast

CA-SRI-1

Garanon Canyon

CA-SCrI-691 CA-SMI-261

Shell Midden Rocky Coast

Calendar Yr BP (1 sigma)

Primary References

8920(8840)8670 8590(8480)8400 8110(7990)7930 8000(7790)7590 7670(7610)7560 8340(8329)8200

Glassow 1993a, 2000

7830(7750)7675

Clifford 2001

12780(12330)11950 10290(10120)10290 10150(9860)9700 9030(9000)8790 8640(8590)8480* 11260(11180)10810

Erlandson et al. 1996a

8340(8180)8050

Johnson 1972 Erlandson 1994 Johnson 1972

10800(10350)10050 8410(8350)8280 8170(8110)8010 8030(7950)7870 7950(7870)7790 7610(7560)7480 9260(8960)8890

Clifford 2001

Greenwood 1978

Rick et al. 2001a

Morris and Erlandson 1993 Erlandson 1994

table 11. (continued) Site No.

Location

Description

CA-SRI-3

Tecolote Point

Shell Midden Cemetery Rocky Coast

CA-SRI-5

Survey Point

CA-SRI-6

Arlington Point

Shell Midden Rocky Coast Shell Midden Rocky Coast

CA-SRI-26

Radio Point

CA-SRI-81

Old Ranch Canyon

CA-SRI-84

Old Ranch Canyon

CA-SRI-116

Lobo Canyon

CA-SRI-173

Arlington Springs

CA-SRI-246

Bechers Bay

CA-SRI-342

Southwest Coast

CA-SRI-666

Old Ranch Canyon

Shell Midden Rocky Coast Shell Midden Coastal Estuary Shell Midden Coastal Estuary Human Burial Coastal Skeletal material Shell Midden Pericoastal Cave Shell Midden Pericoastal Shell Midden Coastal Shell Midden Coastal

Calendar Yr BP (1 sigma)

Primary References

8300(7860)7500 8000(7910)7750 7920(7770)7660 7970(7720)7510 7800(7690)7620 8360(7970)7640 7760(7660)7580 9280(8990)8930 8290(8030)7850 7840(7740)7660 7650(7510)7390 7940(7840)7750

Orr 1968 Erlandson 1994

7430(7400)7300

Kennett 1998

8170(8140)8040

Rick et al. n.d.

10170(9890)9560 10110(9650)9540 9700(9550)9500 13780(13450)13330 13130(12980)12890 8360(7520)7660 7600(7560)7480

Morris and Erlandson 1993

Johnson et al. 2000

7650(7570)7510

Kennett 1998

8100(7990)7930

Erlandson 1994

Orr 1968 Erlandson 1994 Erlandson et al. 1999

Erlandson 1994

Erlandson 1994 Kennett 1998

note: 1 sigma range with mean intercept in parentheses. Calculated by Calib 3.0.3 (Stuiver and Reimer 1993; see Kennett 1998). *Additional radiocarbon dates also available.

TERMINAL PLEISTOCENE TO MIDDLE HOLOCENE

117

Although sea level was in flux for much of the Late Pleistocene and Early Holocene, Daisy Cave was always within a few hundred meters of the coast because the seafloor directly offshore drops off sharply. Faunal assemblages at the site suggest that, from its earliest occupation, rocky intertidal and kelp bed habitats were present in the vicinity (Rick et al. 2001a). Daisy Cave consists of a large exterior rockshelter (4  5 m wide), and a small cavelike fissure (1.5  3 m wide, 11 m deep) extending back into the cliff face (Erlandson et al. 1996a; figure 16). An apron of shell midden covers the slope immediately in front of the rockshelter, and finely stratified, dense shell midden deposits span much of the Holocene. Most of the Late Holocene deposits were removed during the original excavation of the rockshelter by Charles Rozaire (Rozaire 1978), and Jon Erlandson (Erlandson 1994; Erlandson et al. 1996a, 1996b; Kennett et al. 1997; Rick et al. 2001a) has led a team working on the remaining Terminal Pleistocene and Early Holocene deposits since the early 1990s. The basal deposits at the site contain small amounts of marine shell, bone, and charcoal and a small handful of stone flakes (Erlandson et al. 1996b). These data suggest an ephemeral occupation of the cave at this time, one in a series of short-term occupations of the site during the Terminal Pleistocene and Early Holocene (Rick et al. 2001a). Work at Arlington Springs (CA-SRI-173), located on the northwest coast of Santa Rosa Island, may push the initial occupation of the northern Channel Islands back to ~13,500 BP (Johnson et al. 2000). The site consists of a partial human skeleton excavated by Phil Orr in 1959 (Orr 1962, 1968). Orr discovered two human femurs and a humerus exposed in a deeply stratified alluvial deposit 500 m from the mouth of Arlington Canyon. Buried 11 m below the surface in finely stratified sediments, the depth of the material, and its stratigraphic position below a shell midden dating to ~8,000 BP, suggested that the material was of great antiquity. Orr (1968) determined that the skeletal material was not from a formal burial but rather the bones of an individual that had washed into an arroyo and had been buried. Carbonized plant material and the bones of an extinct mouse (Peromyscus nesodytes) were present in the soil surrounding the human remains. Radiocarbon dates on charcoal and collagen extracted from the bones indicated, at the time, an approximate age of 11,600 BP (10,080  800 radiocarbon years BP; Berger and Protsch 1989; Olson and Broecker 1961; Orr 1960), but the bones were poorly preserved and questions have lingered regarding the accuracy of the date. Subsequent work on

figure 16. Photograph of Daisy Cave (CA-SMI-261), San Miguel Island (photo by J. Erlandson).

TERMINAL PLEISTOCENE TO MIDDLE HOLOCENE

119

the stratigraphic section at Arlington Springs confirms the context of the skeletal material, and its age has been reevaluated using a combination of several modern collagen extraction techniques and AMS radiocarbon dating (Johnson et al. 2000). These new age determinations range between 7,500 and 13,500 BP (6610  60, 10,960  80, and 11,490  70 radiocarbon years BP), the oldest date being considered a reliable maximum age and the most recent argued to be too young (Johnson et al. 2000, 544). Pleistocene occupation of the northern Channel Islands was highlighted by Orr (1951, 1967, 1968), who argued for the coeval existence of pygmy mammoths (M. exilis) and humans on Santa Rosa Island as early as 40,000 years ago. In a series of controversial articles, Orr and Berger argued that the close association of mammoth bones, crude stone tools, and “hearths” indicated a clear Pleistocene occupation of the island (Berger 1980, 1982; Berger and Orr 1966; Orr and Berger 1966). Subsequent work suggested that the associations of flaked tools and bones could have easily resulted from natural processes, and some of the fire areas (hearths) may have been a product of chemical weathering (Johnson 1972). Orr’s (1968) Pleistocene shell deposits are equally questionable and can be explained by natural processes (Erlandson 1994). This leaves Daisy Cave and Arlington Springs as the only known sites with deposits clearly dating to the Terminal Pleistocene. Temporal overlap between humans and pygmy mammoths on the northern Channel Islands is still an open question. The presence of mammoths on these islands has elicited scientific interest since the bones of these animals were first reported in the proceedings of the California Academy of Sciences during the late 1800s (Agenbroad 2000b). Hundreds, if not thousands, of mammoth bones have been collected over the past century and curated at museums in Southern California (Orr 1956a, 1956b, 1960, 1967, 1968; Roth 1993). A nearly complete skeleton was discovered in 1994 and at least 140 new fossil mammoth locations have been recorded on San Miguel, Santa Rosa, and Santa Cruz since that time (Agenbroad 2000b). Morphological studies of long bones and teeth suggest that these dwarf mammoths were the size of a large steer (Agenbroad 2000b, 522), but that full-size animals were also present on the islands at some point in the past. The most likely scenario is that larger animals arrived on Santarosae, the Pleistocene island created by low–sea level stands (see chapter 3), but a more gracile mammoth species was favored evolutionarily because its smaller size and lower center of gravity allowed more efficient feeding

120

TERMINAL PLEISTOCENE TO MIDDLE HOLOCENE

in the rugged terrain of this offshore island (Agenbroad 2000b). Selective pressures for small body size would have intensified as the rising sea level significantly reduced the size of Santarosae and split the northern Channel Islands into their current configuration. Overall population sizes also would have been reduced as the sea level rose at the end of the Pleistocene. Preliminary chronological work suggests that the first mammoths arrived on Santarosae before 41,000 years ago and the population persisted on the island until at least ~15,500 years ago (12,840  410; Agenbroad 1998, 2000b). Direct evidence for mammoth hunting (e.g., cut marks or embedded projectile points in bone) has not been discovered on the northern Channel Islands. The new evidence for the early age of Arlington Springs, if correct, along with the most recent suite of dates from pygmy mammoth bones, put the first human occupation of these islands closer in time to mammoth extinction, now only separated by approximately 2,000 years (Agenbroad 2000b; Johnson et al. 2000). Mammoth population levels were surely low at the end of the Pleistocene, and human predation would have had an immediate and devastating effect. The extirpation of the largest ecologically naive animals in similar insular settings is well documented (Anderson 1983, 1989; Kennett et al. n.d.a; Steadman and Kirch 1990; Steadman and Rolett 1996; Steadman et al. 1990), with the bones of these animals, usually found in the earliest settlements, providing the only testament to their existence. If mammoths and humans coexisted on the northern Channel Islands, their relationship would have been fleeting and the evidence minimal, particularly given changes in sea level since that time and radical changes to the landscape that may have destroyed the primary context of most specimens. The close correlation between the appearance of the first humans and the ultimate demise of mammoth populations therefore provides tantalizing evidence for their extirpation by human hunters, but confirmation will require closing the gap between early colonization and mammoth extinction, and future discoveries of more direct evidence for hunting. Regardless of the early role of mammoth hunting, these animals would have been killed off rapidly, and it appears, based on the sparse evidence from Daisy Cave, that people were foraging for maritime resources by at least 12,300 BP (Rick et al. 2001a). Larger settlements, if they existed, were probably positioned along the well-watered northwestern shore of Santarosae. If large sea mammal rookeries were present on the western tip of this large island, as they are today on

TERMINAL PLEISTOCENE TO MIDDLE HOLOCENE

121

San Miguel, they were probably targeted by early hunters. Unfortunately, sea level rise has covered, and most probably erased, any evidence that may have existed for the earliest settlement along the north coast of Santarosae. The near absence of sea mammal bone in many island deposits through much of the Holocene (until ~1500 BP; see chapter 7) suggests that breeding populations were either absent or relegated to offshore rocks that were relatively inaccessible and, therefore, costly to access. Because oceanographic data suggest that conditions were somewhat favorable for breeding colonies it seems likely that human hunters played a role in the reduction of these populations. Within this context, shellfish and fish, visible in the archaeological record, provided the highest rates of return. Owing to the Terminal Pleistocene age of Daisy Cave and Arlington Springs, the northern Channel Islands have become embroiled in a debate regarding the route early migrants followed into the New World (Erlandson 1994, 2002b; Fitzgerald and Jones 2003; Jones et al. 2002; Turner 2003). Early occupation (14,000 BP) at Monte Verde, located in southern Chile, suggests that Paleo-Indian populations entered the New World at least 15,000 years ago (Dillehay 1997, 2000; Meltzer et al. 1997). If the early dates at Monte Verde are accepted, then people colonized the Americas before the ice-free corridor between the Laurentide and Cordilleran ice sheets opened around 13,000 BP (Meltzer et al. 1997), the route generally accepted by archaeologists promoting the idea that big-game hunters (Clovis) were the first people to colonize the New World (Fiedel 1999, 2000). If the ice-free corridor was not available, then the most likely route was along the west coast (Dixon 2001; Mandryk et al. 2001). Early sites in coastal Peru provide some evidence for this proposition (~12,000 BP; Keefer et al. 1998; Sandweiss et al. 1998). Although a coastal migration route is certainly plausible given the prevailing environmental conditions along the Northwest coast (Mandryk et al. 2001), the data from the northern Channel Islands, in and of themselves, are not early enough to support the coastal migration hypothesis (Erlandson 1994). If the early dates from Arlington Springs are upheld, then this skeletal material is coeval with Clovis sites found across western North America (Fiedel 2000), whereas the earliest occupation of Daisy Cave falls sometime between Clovis and Folsom. What these early sites on the northern Channel Islands do suggest is that Paleo-Indian foraging strategies were diverse and a broad range of resources were being used by the Terminal Pleistocene, a pattern that is becoming increasingly clear throughout

122

TERMINAL PLEISTOCENE TO MIDDLE HOLOCENE

the Americas (Anderson and Gillam 2000; Jones et al. 2002; Keefer et al. 1998; Rick et al. 2001a; Roosevelt et al. 1996; Sandweiss et al. 1989, 1998; Yesner 2001).

Early Holocene Record Daisy Cave (CA-SMI-261) also provides one of the earliest glimpses of post-Pleistocene subsistence and settlement on the northern Channel Islands. Two Early Holocene occupational strata have been identified at the site; one dating to between 10,290 and 8,980 BP and the other dating to between 9,270 and 8,480 BP (Erlandson et al. 1996a; Rick et al. 2001a). Similar to the Terminal Pleistocene stratum at the site, these strata suggest sequential short-term use of the rockshelter/cave complex and surrounding marine habitats (Rick et al. 2001a). These deposits are well preserved and contain eel grass cordage and basketry, including a possible sandal fragment (Connolly et al. 1995), and the oldest hook-and-line technology in North America (bone gorges of various sizes; Rick et al. 2001a). Other tools found in these early levels included occasional bifaces; side-scrapers, modified flakes, cores, and Olivella shell beads (Snethkamp and Guthrie 1988). The faunal assemblage at Daisy Cave contains shellfish from rocky shore habitats (California mussels, abalone, tegula, sea urchins, etc.) and the bones of fish, birds, and marine mammals (Snethkamp and Guthrie 1988; Rick et al. 2001a). Fish bone densities are substantial (ranging from 440 to 9,142 g/m3), and the presence of bone gorges in these early strata suggest that fishing was an important subsistence pursuit (Rick et al. 2001a). Dietary reconstructions indicate that fish comprised 50%–60% of the edible meat represented in the assemblage, slightly higher than estimates for the contribution of shellfish (30%–45%). Multiple lines of evidence suggest that people using the cave targeted marine fauna from rocky nearshore environments and kelp beds and that these habitats were within the immediate vicinity of the site. Nonetheless, Walker and Snethkamp (1984) have argued that Daisy Cave, and possibly San Miguel Island, were only occupied seasonally by people living more permanently elsewhere on the mainland or on the larger islands of Santa Rosa or Santa Cruz. Daisy Cave is one of ~19 sites on the northern Channel Islands dating to the Early Holocene (map 8; table 11). Most of these sites are located along the coasts of the outer islands (San Miguel and Santa

TERMINAL PLEISTOCENE TO MIDDLE HOLOCENE

123

Rosa), but Punta Arena, a prominent headland on the south side of Santa Cruz Island, has a clear occupational horizon dating to between ~8,900 and 7,500 BP (CA-SCrI-109; Glassow 1993a, 2000), and new radiocarbon dates from two ephemeral interior sites on eastern Santa Cruz suggest some occupation farther to the east (SCrI-608 and 610; see map 8). Clear evidence for the occupation of Anacapa, the smallest and most eastern island, is absent for the Early Holocene. Early Holocene sites include an isolated human burial (CA-SRI-116) dating to ~9,700 BP (Morris and Erlandson 1993) and a cemetery (CA-SRI-3) at Tecolote Point dating to between 8,000 and 7,500 BP, but a majority of the sites dating to this time are coastal shell middens. Settlement was most persistent and focused along the well-watered coast of Santa Rosa Island (Erlandson 1994; Kennett 1998; Orr 1968), but smaller sites dating to the Early Holocene also occur: along the north coast of San Miguel, around a relict estuary on Santa Rosa Island (Rick et al. n.d.), at Punta Arena on southern Santa Cruz (Glassow 2000), and at interior locations on eastern Santa Cruz (Clifford 2001; see map 8). Most of these sites are thin middens composed of marine shell and the bones of fish, birds, and occasionally sea mammals. Tool assemblages are limited and composed primarily of expedient flakes and cores made from locally available volcanic rocks and island chert. The remains of houses are absent at these sites, and many appear to be campsites or special purpose locations for collecting shellfish or plant foods, rather than permanent villages. The most concentrated activities during the Early Holocene were centered on the north coast of Santa Rosa Island, near the mouth of Arlington Canyon. A complex sequence of Early and Middle Holocene dune activity, interspersed with extensive evidence for human occupation (CA-SRI-3, -6, -173), is visible along the coast between Tecolote and Arlington canyons. The most prominent Early Holocene site in this area occurs at Tecolote Point (CA-SRI-3), a site that consists of a shell midden dominated by red abalone (H. rufescens) shells dating to between 8,000 and 7,500 BP and an associated cemetery with some burials dating to the same time (Erlandson 1994; Orr 1968). The remains of houses are not evident at this site, but the presence of a cemetery suggests a certain degree of sedentism, or at least a central place that people returned to with relative frequency—perhaps from hunting and foraging expeditions on San Miguel and Santa Cruz islands. Just to the east of Tecolote Point, two Early Holocene sites have also been identified, one on each side of Arlington Canyon.

124

TERMINAL PLEISTOCENE TO MIDDLE HOLOCENE

Arlington Springs (CA-SRI-173), on the west side of the canyon, is best known for its deeply buried skeletal remains dating to the Terminal Pleistocene, but a dark soil lens 3.3 m below the surface contains large red abalone shells and dates to ~8,000 to 7,500 BP (Erlandson 1994). A laterally extensive series of four strata, spanning the Early Holocene, is evident on the east side of Arlington Canyon (CA-SRI-6) and suggests repeated use of this location throughout this time (9,300 to 7,500 BP; Erlandson et al. 1999). The importance of fish and fishing reflected in the faunal assemblage at Daisy Cave is unique when compared to other sites dating to the Early Holocene. Most of the known Early Holocene middens appear to be dominated by mollusk shells. Fish and sea mammal bones are also present in these deposits indicating that bone preserves in these contexts and that nearshore fishing in kelp beds was an important subsistence pursuit, as was taking the occasional sea otter, seal, or sea lion. However, dietary analyses at several Early Holocene sites suggests that shellfish provided the bulk of edible meat consumed. Meat yield estimates for the Early Holocene levels at Punta Arena on southern Santa Cruz suggest that 91% of the consumed meat weight came from shellfish, 5% from fish, and 4% from marine mammals (Glassow 1993a). Excavations at two sites at the mouth of Old Ranch Canyon produced similar results (CA-SRI-666, 98% shellfish; CA-SRI-84, 97% shellfish; Kennett 1998; Rick et al. n.d.). Work at CA-SRI-6, one of several important sites in the Arlington Springs area, has also shown that shellfish provided 85% of the edible meat with fish providing another 15% (Erlandson et al. 1999). The fish assemblage at this site was dominated by sheephead (Semicossyphus pulcher), a species that is commonly found in rocky nearshore and kelp bed habitats. Early Holocene shellfish assemblages vary and appear to be dependent upon when they accumulated and the habitats that were available within the vicinity of each site. Shellfish assemblages in the vicinity of Old Ranch Canyon dating to between ~8,200 and 7,400 BP are composed of rocky intertidal (M. californianus) and estuarine (e.g., Washington clams) species, suggesting that both habitats were present and that the estuary was larger, and more productive, when sea level stabilized between 9,000 and 7,000 BP. Shellfish assemblages found elsewhere on the islands are composed primarily of species from rocky intertidal habitats, with a preference for the largest, most prolific species (Haliotis spp., Mytilus spp.) available at the time. Red abalone (H. rufescens), a large, cooler-water species (see Middle Holocene

TERMINAL PLEISTOCENE TO MIDDLE HOLOCENE

125

section below) is common in middens within the Arlington Springs region between 8,500 and 8,000 BP. However, midden samples analyzed from SRI-6, a site located in the same general vicinity with deposits dating to 9,300 BP, are composed predominantly of black abalone, a smaller abalone species that prefers warmer water (73% of assemblage; Erlandson et al. 1999). These data are consistent with the earliest Holocene strata at Daisy Cave, and other early sites on San Miguel, where red abalone are also relatively rare (Erlandson and Rick 2002b; Erlandson et al. 2004). If the presence or absence of red and black abalone in Early Holocene shellfish assemblages provide a proxy for changing water temperatures, these data suggest that marine conditions were warmer between 11,000 and 9,000 BP and cooler between 8,500 and 8,000 BP. This is generally supported by the new marine climate record for the region indicating warmer sea-surface temperatures and low marine productivity between 11,000 and 9,500 BP, and cooler, more productive conditions between 9,500 and 8,000 BP (see figure 11). Warm sea-surface temperatures, lower marine productivity, and changing sea level may have created unfavorable conditions for many species of shellfish, and the earliest Holocene shellfish assemblages (11,000 to 9,000 BP) on the northern Channel Islands appear to be depauperate relative to later assemblages. This may explain the anomalous focus on marine fishing evident in the early record at Daisy Cave (Rick et al. 2001a). The near absence of marine mammal bone in most Early Holocene middens on the northern Channel Islands is surprising, particularly given their abundance in sites on the southern Channel Islands (Porcasi et al. 2000). Whaling never appears to have been an important subsistence pursuit in the Santa Barbara Channel region, and the hunting of cetaceans (dolphins) appears to have first developed during the Middle Holocene and clearly occurred during the Late Holocene. Seals and sea lions (pinnipeds) frequent the waters surrounding the northern Channel Islands today and are particularly common around San Miguel Island where large numbers aggregate in rookeries seasonally (see chapter 3). One possible explanation for the dearth of marine mammal bone, particularly of pinnipeds, is that marine conditions during the Early Holocene did not favor the formation of coastal rookeries and, therefore, seals and sea lions were only sporadically available in the waters surrounding the islands. This may have been the case between 11,000 and 9,000 BP, and again between 8,000 and 7,000 BP, when sea-surface temperatures were warm and marine productivity was relatively low

126

TERMINAL PLEISTOCENE TO MIDDLE HOLOCENE

(see figure 11). However, marine conditions during the intervening interval (9,500–8,000 BP) were cool and productive. Three additional factors may account for the low frequency of marine mammal bone in Early Holocene sites. First, marine mammals may have been concentrated on the western end of San Miguel, as they are today, but evidence for hunting may be submerged in sites closer to the Terminal Pleistocene/Early Holocene shorelines. Second, human hunting during the Terminal Pleistocene and earliest Holocene reduced marine mammal populations and rookeries were restricted to offshore rocks where the costs of exploitation exceeded the benefits of doing so. Third, human populations were centered on the northern coast of Santa Rosa, far removed from San Miguel Island rookeries, and central place foraging returns were low because of travel distance and limitations imposed by the available maritime technology needed to transport meat back to more centralized residential bases. Future work will be required to determine why marine mammal bones are so rare in Early Holocene middens, but a combination of these factors is probably responsible. Dietary reconstructions from the northern Channel Islands are generally consistent with the economic patterns at Early Holocene sites along the Southern California coast (Erlandson 1994; Jones et al. 2002). On the coastal mainland, milling-stone sites provide clear evidence that shellfish from coastal estuaries were complemented with plant foods, primarily small seeds, from more interior locations (Erlandson 1994). Milling stones are rare in Early Holocene island sites and the stratigraphic integrity of the known examples are questionable at this point (Orr 1968; see Erlandson 1991a, 106). Plant foods surely played an important subsistence role on the islands, but clear evidence for their use is currently lacking. The presence of digging-stick weights in some Early Holocene burials (similar to the example in figure 17; page 135) is at least suggestive that plant foods were important. This idea is also supported by the position of some Early Holocene sites (CA-SRI-3, -4, -5, -6, -173) on the landward edge of the coastal plain, which extended 0.5 km seaward during the Early Holocene (see figure 10). Two new sites in the interior of eastern Santa Cruz (CA-SCrI-681, -691) may provide additional evidence for Early Holocene plant use. These sites both date to between 8,500 and 7,500 BP and are located on small hills along the ridge system surrounding Scorpion drainage. These sites are composed of dark midden soil mixed with marine shell and are similar in character to interior sites dating through the Middle Holocene (7,500 to 3,000 BP; Clifford 2000), inferred to represent relatively

TERMINAL PLEISTOCENE TO MIDDLE HOLOCENE

127

intensive prehistoric plant-harvesting practices. This suggests that these Middle Holocene strategies have an origin in the Early Holocene, at least on the eastern end of Santa Cruz Island. The idea that the Arlington Springs area was an important central place for populations on the islands is supported by the presence of a cemetery at Tecolote Point (CA-SRI-3). Burials at the site occur in a stable dune capped by a shell midden dating to the Middle Holocene (Erlandson 1994), and erosional cuts through the site expose the complexities of its stratigraphy. Phil Orr excavated 79 burials at this location in 1947 (Cemetery A) and argued, based on a limited number of radiocarbon dates, that it was occupied prior to 7,000 BP (Early Dune; Orr 1968, 115–135). Large red abalone shells were common burial goods, and red ocher stained the bones of many individuals. Much of the cemetery appears to be associated with the red abalone midden dating to between 8,000 and 7,500 BP. Subsequent work on the chronology of the cemetery indicates that some burials probably do date to this time but that others date to the Middle Holocene (~5,000 BP; Erlandson 1991a, 1994), suggesting settlement continuity at this location for several thousand years. An analysis of grave goods suggests that many of the burials date to the Early Holocene (Phase X; King 1990), but additional work will be required to determine the true chronological composition of the cemetery. Regardless, the presence of a formal burial ground on the northern Channel Islands suggests a permanent population living on the northern Channel Islands by between 8,000 and 7,500 BP and highlights the Arlington Springs area as a focal point of settlement at this time. The cemetery at Tecolote Point (CA-SRI-3) also provides a unique glimpse into the social organization of early island populations. Orr (1968) noted that a majority of the burials were flexed, a pattern that is typical for the Early Holocene elsewhere in Southern California. Red abalone shells were frequently interred with individuals, and their bodies appear to have been ceremonially covered with red ocher. A total of 334 grave goods were identified with 57 of the 79 burials (King 1990), a small number when compared to later time periods. These included Olivella shell beads (N  300), bone artifacts (abalone pry bars, awls, bipoints), groundstone implements (digging-stick weights), and bifacially flaked tools (leaf-shaped bifaces). Many of the burials were simple, containing one or two items added to the graves by loved ones. Most of the remaining grave goods were concentrated in a few burials, a pattern often interpreted as status differentiation (acquired or inherited). However, one of

128

TERMINAL PLEISTOCENE TO MIDDLE HOLOCENE

the burials that contained large numbers of items clearly dates to the Middle Holocene (Erlandson 1994), and the artifact types in the other two suggest that they do as well. The remaining burials are quite comparable and suggest that an egalitarian social form existed, perhaps with certain individuals gaining status and wealth through their lifetimes—a small-scale tribal society without clear evidence for ranking and hereditary status positions. The available data indicate that populations on the northern Channel Islands were relatively low throughout the Early Holocene. In fact, it is possible that a single group of between 30 and 50 people centered on the Arlington Springs area, foraged widely over the Channel Islands, and created the sparse archaeological record visible today. The cemetery at Tecolote Point, along with several substantial shell middens in the Arlington Springs region, suggests that this was an important settlement locus and that it provided a central place for a range of economic and ceremonial activities. Dietary reconstructions indicate that shellfishing in rocky intertidal habitats was the primary subsistence pursuit, but that fishing and marine mammal hunting were also important at certain times and places. It is also possible that much of the evidence for marine mammal hunting was covered by the final stages of the marine transgression between 9,000 and 7,000 BP. Protein-rich coastal resources were probably complemented with seeds and bulbs collected from interior island settings, but future research will be required to determine the extent and nature of these activities. Burial data suggest that islanders’ social differentiation was limited and an egalitarian social form was present at this early date. Early Holocene subsistence and settlement provided the foundation for further developments during the Middle Holocene, including the replication and expansion of economic and ceremonial activities to other large, well-watered drainages that occurred in the Middle Holocene.

Middle Holocene Record The economic and social patterns evident on the northern Channel Islands at the tail end of the Early Holocene solidified and persisted throughout the Middle Holocene (7,500–3,000 BP). These records are richer than the Terminal Pleistocene and Early Holocene records combined, and it is clear that population levels on these islands had increased substantially by 7,500 BP. Middle Holocene sites are still

TERMINAL PLEISTOCENE TO MIDDLE HOLOCENE

129

limited to the larger three islands (San Miguel, Santa Rosa, and Santa Cruz) but are more widely distributed in both coastal and interior settings. Compared with sites dating to the Late Holocene (see chapter 7), most sites dating to the Middle Holocene have a low density of artifacts and are sometimes difficult to place chronologically without radiocarbon dating. Faunal assemblages (e.g., red abalone) and site location (interior sites) provide clues that settlements are of Middle Holocene age, but radiocarbon dating is often necessary for age verification. Fifty-eight sites on the northern Channel Islands have verified Middle Holocene deposits (table 12). These sites vary substantially owing to differences in function, intensity of use, duration of settlement, and postdepositional alteration. Sites are grouped provisionally into four categories for heuristic purposes: primary villages, secondary villages, interior residences, and logistical encampments. Primary villages were those locations clearly occupied for extended periods of time and served as central places for a variety of economic and social activities. These sites are large, contain deeply stratified or laterally extensive midden deposits, and have substantial cemeteries associated with them. The accumulation of domestic debris (middens) at secondary villages is similar to primary villages, but cemeteries have not been discovered at these locations. In some cases this may be the result of insufficient archaeological work, in other instances the absence of cemeteries may be the product of differing prehistoric site use (e.g., smaller numbers of people living at the site for shorter amounts of time). Interior residences are often smaller than primary and secondary villages and positioned on ridges or small knolls away from the coast (Clifford 2001; Kennett 1998; Kennett and Clifford 2004). These sites range from small, ephemeral deposits, suggesting a single usage, to large, deeply stratified sites representing repeated use or more permanent settlement. Logistical encampments are those sites that suggest special-purpose activities, usually the collection and processing of a resource that is located far away from primary villages. The best known logistical encampments on the islands are thin pavements of marine shell (usually abalone or California mussel) positioned near rocky promontories or along highly productive stretches of rocky, but accessible, coastline. Many of the ephemeral “interior residences” may be logistical encampments that were used periodically to collect and process plant foods (seeds and bulbs) from surrounding environs. Map 9 shows the distribution of known Middle Holocene sites. The chronological placement of some of these sites is verified with

table 12. Known Middle Holocene archaeological sites on the northern Channel Islands Site No.

Location

Description

Calender (BP) 1 sigma 7260(7180)7020 4440(4290)4110 4520(4420)4350 6880(6790)6720 7420(7240)6990 7410(7270)7170 4770(4570)4490 7230(7070)6890 3580(3469)3387 5403(5295)5247 3200(3030)2880 3830(3710)3630 4870(4830)4650* 5450(5306)5241 6161(5933)5763 5894(5770)5688 4490(4395)4281 4818(4708)4552 4790(4611)4512 5896(5844)5725 6296(6239)6159 5336(5289)5238 5121(4986)4864 4408(4293)4157 3370(3330)3260

CA-SRI-1 CA-SRI-3

Garanon Tecolote Canyon

Logistical Encampment Primary Village

CA-SRI-4

Arlington Canyon

Primary Village

CA-SRI-5

Arlington Canyon

Primary Village

CA-SRI-22 CA-SRI-27 CA-SRI-41

Arlington Canyon Arlington Canyon Cañada Verde

Interior Residence Interior Residence Primary Village

CA-SRI-43

Cañada Verde

Interior Residence

CA-SRI-46 CA-SRI-50

Cañada Verde Cañada Verde

Interior Residence Interior Residence

CA-SRI-51 CA-SRI-54 CA-SRI-55 CA-SRI-70 CA-SRI-73 CA-SRI-84

Cañada Verde Cañada Verde Cañada Verde Arlington Canyon Cañada Verde Old Ranch Canyon

Interior Residence Interior Residence Interior Residence Interior Residence Interior Residence Secondary Village

References Erlandson 1994 Orr 1968 Erlandson 1994 Orr 1968 Kennett 1998 Orr 1968 Kennett 1998 Kennett 1998 Orr 1968 Kennett 1998 Orr 1968 Kennett 1998 Kennett 1998

Kennett 1998 Kennett 1998 Kennett 1998 Kennett 1998 Kennett 1998 Kennett 1998

CA-SRI-109 CA-SRI-109 CA-SRI-116

Carrington Point

Logistical Encampment

Lobo Canyon

Secondary Village

CA-SRI-147

Jolla Vieja

Interior Residence Secondary Village?

CA-SRI-246 CA-SRI-270 CA-SRI-287 CA-SRI-342

Bechers Bay Southwest Coast Southwest Coast Southwest Coast

Logistical Encampment Logistical Encampment Logistical Encampment Logistical Encampment

CA-SRI-344 CA-SRI-364 CA-SRI-462 CA-SRI-589 CA-SRI-590 CA-SRI-591 CA-SRI-626 CA-SRI-627 CA-SRI-628 CA-SCrI-9 CA-SCrI-34 CA-SCrI-36

Southwest Coast Southwest Coast South Coast Arlington Canyon Arlington Canyon Arlington Canyon Arlington Canyon Arlington Canyon Cañada Verde Coches Christy Beach Coches

Logistical Encampment Logistical Encampment Logistical Encampment Interior Residence Interior Residence Interior Residence Interior Residence Interior Residence Interior Residence Interior Residence Logistical Encampment Interior Residence

5312(5263)5123 5770(5699)5612 5300(5250)5040 6110(5990)5930 6300(6260)6180 4250(4120)3980 5700(5620)5580 7400(7300)7230 7550(7508)7436 5756(5685)5602 5699(5611)5564 6090(5967)5902 7594(7531)7464 6054(5953)5901 5685(5600)5558 7007(6906)6839 2748(2704)2554 4339(4236)4140 4279(4153)4064 4708(4562)4462 5284(5205)5002 6556(6457)6387 5652(5591)5475 6038(5869)5638 4997(4862)4815

Kennett 1998 Kennett 1998 Erlandson 1994 York 1996 Kennett 1998 Kennett 1998 Kennett 1998 Kennett 1998 Kennett 1998 Kennett 1998 Kennett 1998 Kennett 1998 Kennett 1998 Kennett 1998 Kennett 1998 Kennett 1998 Kennett 1998 Kennett 1998 Kennett 1998 Peterson 1994 Glassow 1997 Peterson 1994

table 12. (continued) Site No.

Location

Description

CA-SCrI-109

Punta Arena

Secondary Village?

CA-SCrI-236 CA-SCrI-333

Christy Beach Forneys Cove

Logistical Encampment Primary Village

CA-SCrI-424 CA-SCrI-426 CA-SCrI-427 CA-SCrI-428 CA-SCrI-429 CA-SCrI-430 CA-SCrI-608 CA-SCrI-608 CA-SCrI-608 CA-SCrI-608 CA-SCrI-610 CA-SCrI-669 CA-SCrI-670 CA-SCrI-673 CA-SCrI-678 CA-SMI-1

Forneys Cove Forneys Cove Forneys Cove Forneys Cove Forneys Cove Forneys Cove Scorpion Drainage

Logistical Encampment Logistical Encampment Logistical Encampment Logistical Encampment Logistical Encampment Logistical Encampment Interior Residence

Scorpion Drainage Scorpion Drainage Scorpion Drainage Scorpion Drainage Scorpion Drainage Nidever Canyon

Interior Residence Interior Residence Interior Residence Interior Residence Interior Residence Interior Residence

Calender (BP) 1 sigma 5300(5050)4850 5670(5600)5570 6400(6300)6210* 4969(4835)4783 3640(3510)3470 3970(3870)3770 4780(4480)4410 6170(5930)5750* 5854(5670)5568 5305(5227)5006 5647(5567)5447 4134(3977)3842 7537(7421)7319 6039(5903)5744 3686(3595)3478 4432(4359)4238 3718(3663)3602 3715(3634)3559 6909(6853)6775 4389(4258)4145 3245(3144)3016 3715(3623)3499 4379(4258)4151 3440(3340)3250 7079(6982)6879

References Glassow 1997 Glassow 2000 Glassow 1997 Wilcoxon 1993

Glassow 1997 Glassow 1997 Glassow 1997 Glassow 1997 Glassow 1997 Glassow 1997 Kennett 1998

Clifford 2001 Clifford 2001 Clifford 2001 Clifford 2001 Clifford 2001 Erlandson 1991c

CA-SMI-261

Bay Point

Logistical Encampment Rockshelter/Cave

CA-SMI-350 CA-SMI-388 CA-SMI-470 CA-SMI-492 CA-SMI-528 CA-SMI-557 CA-SMI-603

Harris Point Simonton Cove Otter Creek/Point East of Pt. Bennett Point Bennett Southeast Coast Bay Point

Logistical Encampment Logistical Encampment Logistical Encampment Logistical Encampment Logistical Encampment Logistical Encampment Logistical Encampment Rockshelter/Cave

3240(3130)2980 3550(3450)3360 4540(4460)4400 6800(6720)6630* 6850(6678)6483 7259(7158)6995 4347(4215)4081 5577(5472)5337 5901(5844)5718 6276(6161)5954 4160(4070)3960 4440(4390)4280 6680(6610)6490*

Erlandson et al. 1996a

Greenwood 1978 Greenwood 1978 Kennett 1998 Walker and Snethkamp 1984 Walker et al. 2000 Glassow 1997 Vellanoweth et al. 2000

note: 1 sigma range with mean intercept in parentheses. Calculated by Calib 3.0.3 (Stuiver and Reimer 1993; see Kennett 1998). *Dates show the age span for this site, but additional radiocarbon dates are available.

134

TERMINAL PLEISTOCENE TO MIDDLE HOLOCENE

radiocarbon dating, but others are inferred to be of Middle Holocene age because of their location (interior residences) or faunal constituents (red abalone sites). The Middle Holocene age of the latter will need to be verified in the future. As in the Early Holocene, the north coast of Santa Rosa Island continued to be an important node of settlement, but the number of primary and secondary villages increased and expanded to several well-watered locations east of Arlington Canyon (Kennett 1998; Orr 1968). Primary and secondary villages are also present on the western and southern sides of Santa Cruz Island (Glassow 1993a, 2000; King 1990; Wilcoxon 1993). Although extensive surveys were conducted on eastern Santa Cruz, large coastal middens suggesting primary or secondary village locations have not been identified along the coast (Clifford 2001; Kennett 1998; Perry 2003), but one interior site may have served as a village of some kind (CA-SCrI-608). Interior residences are common on Santa Rosa and Santa Cruz (Kennett 1998), and one site on San Miguel Island (CA-SMI-1) resembles these settlements and may date primarily to the Middle Holocene (Erlandson 1994). Logistical encampments are the dominant site type on San Miguel Island, along with red abalone middens, and these sites are also common along the coasts of Santa Rosa and Santa Cruz. The current evidence suggests that settlement and subsistence activities intensified at the mouth of Arlington Canyon during the Middle Holocene. Three sites (CA-SRI-3, -4, -5) on the west side of the canyon contain substantial deposits dating to this time, including several cemeteries (Orr 1968). The dunes that extend east along the coast from Tecolote Point (CA-SRI-3) into the mouth of Arlington Canyon (CA-SRI-5) are capped with midden deposits dating to the Middle Holocene. Limited surface collections on the upper dune surfaces in the area have revealed a range of formal artifacts in residential midden contexts (figure 17), however, no clear evidence for house floors or other domestic structures are known. Radiocarbon dates from CA-SRI-5 confirm the Middle Holocene age of these deposits (Kennett 1998). Phil Orr also identified and excavated three cemeteries on the west side of Arlington Canyon (CA-SRI-3, -4, -5). Several of the burials in the predominantly Early Holocene cemetery at Tecolote Point (CA-SRI-3, Cemetery A) are now known to date to the Middle Holocene (Erlandson 1994). Cemetery B, at the same site, is also likely to date to the Middle Holocene based on associated artifacts (Orr 1968, 130). CA-SRI-4, and -5, to the east of Tecolote Point, are basically continuations of CA-SRI-3. Orr identified and excavated

map 9. Distribution of settlements and cemeteries on the northern Channel Islands dating to the Middle Holocene (7,500–3,000 BP; map produced by D. Kennett with the assistance of J. Bartruff).

figure 17. Selection of artifacts found at primary coastal village sites on the northern Channel Islands dating to the Middle Holocene: (a, b, c) bone gorges from CA-SRI-5; (d) steatite pendant from CA-SRI-5; (e) square Olivella-shell-wall bead from CA-SRI-5; (f) abalone (Haliotis rufescens) pendant from CA-SRI-5; (g) bifacially flaked cobble tool from CA-SRI-5; (h) incised bird-bone hairpin from CA-SRI-41; (i) grooved sandstone net weight from CA-SRI-3; (j) fragment of digging-stick weight from CA-SRI-5. (Illustrations and layout by R. van Rossman.)

TERMINAL PLEISTOCENE TO MIDDLE HOLOCENE

137

burials at both locations, and artifact assemblages place a majority of these burials in the Middle Holocene (Orr 1968, 139–143 and 147–149). These included bone points, square clamshell bead ornaments, and Megathura shell ornaments (Glassow 1997). Similar to the Arlington area, an extensive dune formation capped with Middle Holocene middens extends out of Cañada Verde toward Brockway Point to the west (CA-SRI-41). Cañada Verde is the thirdlargest drainage on the island and is a vigorous perennial water source—even in dry years. Excavations at the site indicate that deposits span the latter half of the Middle Holocene (4,340–3,020 BP) and into the Late Holocene (until 1,200 BP; Kennett 1998). Human bone is scattered across its surface, and some formal artifacts occur in the residential midden exposed in the seacliff. A stone pavement was revealed in a deflated dune surface at the site and is the only substantial architecture known; however, it is unclear whether it is part of an old house. Orr (1968, 149–177) excavated two cemeteries at CA-SRI-41: one dating to between 3,500 and 3,000 BP (Cemetery A) and the other dating to between 4,840 and 4,400 BP (Cemetery X). Burial associated artifacts included Olivella shell beads [spire ground, barrel, and square wall beads (the latter in Cemetery X only)], bone points, contracting-stem dart points, mortars and pestles (Cemetery A only), and digging-stick weights (King 1990; Orr 1968, 149–176). Substantial residential midden deposits are also known from several other locations on Santa Rosa Island. One of these occurs at the mouth of Lobo Canyon (CA-SRI-116), located on the northeast end of the island. Just east of the drainage, a deep midden (1.5 m), composed of dark soil and mollusk shells (primarily California mussel), sits stratigraphically above an old dune containing an Early Holocene burial (9,700 BP; see section “Early Holocene Record” above). Radiocarbon dates from the upper and lower part of these deposits suggest a substantial occupation at this location between 6,260 and 5,230 BP. Similarly, extensive midden deposits also cap a dune complex on the southern periphery of Old Ranch Canyon (CA-SRI-84). These deposits are stratigraphically above the Early Holocene deposits at this same location. Radiocarbon dates suggest that much of this midden accumulated between 5,000 and 4,000 BP, rather than in the Early Holocene. Neither CA-SRI-84 nor CA-SRI-116 contains definitive evidence for permanent occupation (e.g., house floors) and excavations at these sites have not been substantial enough to identify formal cemeteries. However, the size, depth, and character of these sites suggest a use beyond

138

TERMINAL PLEISTOCENE TO MIDDLE HOLOCENE

small logistical encampments, and they were likely primary or secondary village locations. Pedestrian surveys along the southern and southwestern sides of Santa Rosa revealed Middle Holocene age sites (logistical encampments), but nothing large enough to be considered a primary or secondary village (Kennett 1998; York 1996). Substantial village sites are also known from western and southern Santa Cruz Island (Glassow 1993a, 2000; King 1990; Wilcoxon 1993). The most significant one identified thus far is CA-SCrI-333, a deeply stratified midden deposit and cemetery located near Fraser Point (west end) dating to between 5,740 and 3,500 BP. Midden deposits also extend into the Late Holocene at this location (until ~1,300 BP; Wilcoxon 1993, 148). The lowest strata (below 1 m; 5,740–5,000 BP) at the site contain large red abalone and California mussel shells, along with fish and bird bone. Upper deposits (3,500–1,300 BP) are composed of black abalone, California mussel, and fish bone (Wilcoxon 1993, 147). The excavated residential midden at this location had a paucity of finished artifacts, and many of the tools found were broken or expended. Artifact types associated with the lower strata included bone fish gorges, clam disk beads, and large side-notched projectile points. Upper levels contained j-shaped Mytilus and Haliotis fishhooks. Many of the houses identified at this site also occur in the upper levels. These data, combined with earlier excavations of the formal cemetery by Olson and Van Valkenburgh, suggest that this was a primary village location for much of the Middle and the Late Holocene (King 1990). It is less clear whether the deposits at Punta Arena (CA-SCrI-109), a large site (11,200 m2) located on the south side of Santa Cruz, represent a village or a logistical encampment that was used repeatedly throughout the Middle Holocene. This site is positioned on a rocky promontory close to a spring and a perennial stream located just 250 m northeast of the point (Glassow 2000). The principal occupation of this location took place between 6,150 and 5,250 BP (Glassow 2000, 555), a distinctive stratum separated from the Early Holocene levels by a relatively sterile dune deposit. The Middle Holocene deposits are thick (~1.5 m) and contain large numbers of California mussel shells mixed with the shells of red and pink abalone and other shellfish species in much lower densities (Sharp 2000). The bones of dolphins, pinnipeds, sharks, and fish are also relatively common. Igneous flakes and cores are the most common artifacts, and fire-cracked stones occur through the midden in relatively low concentrations (Glassow 2000, 557). House floors are not visible in the vertical midden exposures, and a formal burial ground dating to the

TERMINAL PLEISTOCENE TO MIDDLE HOLOCENE

139

Middle Holocene has not been identified at the site. The size and depth of this site suggest that it may have been a secondary village or a logistical encampment that was visited frequently throughout the Middle Holocene. Interior sites, contemporary with primary and secondary villages along the coast, were commonly used on Santa Rosa and Santa Cruz islands through the Middle Holocene (map 9; table 12). At least one site (CA-SMI-1), similar in character to these interior settlements, is also known from San Miguel Island (Erlandson 1991c). The site is positioned on the northeast side of the island near the headwaters of Nidever Canyon and appears to date primarily to the Middle Holocene, although Early and Late Holocene components are also present (Erlandson 1994; Kennett 1998). The vast majority of these interior sites are positioned on hilltops or ridgelines, rather than in drainage bottoms closer to water (figure 18; Clifford 2001; Kennett 1998; Kennett and Clifford 2004). Pedestrian surveys of island interiors have been limited, but over 200 sites positioned on hilltops or ridgelines are now recorded on the larger islands of Santa Rosa and Santa Cruz (Kennett 1998). Radiocarbon dates confirm the Middle Holocene age of 24 of these sites, with two sites on Santa Cruz dating to the Early Holocene (discussed previously) and one firmly dating to the Late Holocene (Clifford 2001). Interior sites were commonly encountered during a pedestrian survey of Arlington Canyon and Cañada Verde, two large drainages on the north side of Santa Rosa Island (Kennett 1998; map 10). The northern side of Santa Rosa is composed of a series of Quaternary terraces that gradually step down from the central mountain range separating the two sides of the island. Many of the interior sites are situated on the leeward side of ridgelines, hills, and small knolls, slightly protected from the prevailing northwesterly wind that batters the island. These locations afford broad views across the landscape but are not positioned close to perennial water sources (Kennett 1998). The large villages at the mouths of Arlington Canyon (CA-SRI-3, -4, -5) and Cañada Verde (CA-SRI-41) were the primary settlement loci in this area during the Middle Holocene. Interior sites were all within a 2-hour walk of these coastal villages (0.7–4.85 km), and they exhibit a broad range of sizes (20–6,800 m3) and depths (30–100 cm). This suggests that some were short-term logistical encampments, whereas others were occupied either more permanently, as satellite communities, or repeatedly, perhaps on a seasonal basis. Some settlement permanence is suggested by a range of artifacts found in natural exposures or surfaces of these sites,

140

TERMINAL PLEISTOCENE TO MIDDLE HOLOCENE

figure 18. Photograph of northern Santa Rosa Island (Arlington Canyon) showing distribution of Middle Holocene interior residences located on hilltops or ridgelines. Notice slight vegetational differences across the surface of CA-SRI-647 (foreground) caused by anthropogenic soil conditions. Also notice coastal sage scrub plant communities along the wall of Arlington Canyon (photo by S. Spaulding).

including contracting-stem projectile points, digging-stick weights, and other groundstone tools (figure 19). Preliminary surveys of the hills and ridgelines surrounding Old Ranch Canyon suggest a similar pattern of interior settlement on the eastern side of Santa Rosa Island. Phil Orr recorded many of the larger interior sites on Santa Rosa and defined his “Highlander Phase” based on excavations at several of these locations (Orr 1968, 99–100). He argued that people occupied these settlements permanently between about 6,000 and 4,000 years ago, when climatic conditions were moister than today’s, and the island was covered with forest (oaks and island cherry). The idea that these sites

map 10. Middle Holocene distribution of primary/secondary villages and interior residences on Santa Rosa and San Miguel islands. Maximum extent of the grayscale rings shows 2-hour walking distances (cost surfaces) radiating out from the four known villages on Santa Rosa Island (map produced by D. Kennett with the assistance of J. Bartruff).

figure 19. Selection of artifacts from interior hilltop/ridgeline sites collected during surveys on the north coast of Santa Rosa Island: (a, b, c, e) large bifaces; (d) bifacially flaked scraper; (f) digging-stick weight fragment. (Illustrations and layout by R. van Rossman.)

TERMINAL PLEISTOCENE TO MIDDLE HOLOCENE

143

were permanently occupied was based on the size and thickness of the deposits he excavated and the presence of faint surface depressions at these locations (2.5–3 m in diameter), interpreted by Orr as the remnants of houses. With regard to these surface depressions, Orr wrote No complete survey of the number of house pits has been made (at interior sites). The difficulties of counting these on more than 100 sites, many of them difficult to reach, has not appeared worth the effort, but a spot check of our locality records of 25 sites picked at random show 472 pits recorded. The largest number counted was 49, but the typical site contains about 15 to 20 house pits. (Orr 1968, 179)

Subsequent work at these sites suggests that these surface depressions were created by cows, present on the island from 1902 to 1998, wallowing in these midden deposits (Kennett 1998). It is not surprising then that Orr failed to find postholes, floors, or central hearth features during “numerous” excavations of pit features at CA-SRI-19, -24, -43, -44, -66 (Orr 1968, 179–188). Limited work at CA-SRI-50, a site located on the west side of Cañada Verde, provides some information about the character of larger interior settlements (Kennett 1998). The site is 3,500 m2 in size and positioned on a hilltop about 2 km south of CA-SRI-41, the closest primary village in the vicinity. Geophysical work at the site helped identify several subsurface burned deposits (probable hearths) and one possible circular house feature. A small test trench (50  100 cm) was excavated near the edge of the circular feature, and a flat, compact surface, with California mussels embedded in it, was encountered ~40 cm below the surface. Auger tests verified the extent of the feature, but no obvious postholes were encountered in our small test excavation. Radiocarbon dates from the site suggest that these deposits are partly contemporary with those at the primary village on the coast (CA-SRI-41; 4,710–4,400 BP). A larger excavation at the site will be needed to determine if this feature is, in fact, a house floor. Regardless, the dark, anthropogenic soil at this and other sites, mixed with marine mollusk shells and bone (sea mammal, fish, and bird) and containing relatively diverse tool assemblages, is at least suggestive of longer-term or repeated use. A systematic pedestrian survey of Scorpion drainage on eastern Santa Cruz Island revealed a similar pattern of Middle Holocene hilltop and ridgeline settlement (Clifford 2001; Kennett 1998). Scorpion drainage flows from west to east out of the El Montañon range and along the northeast side of the island. The drainage is surrounded by

144

TERMINAL PLEISTOCENE TO MIDDLE HOLOCENE

rolling hills and accessible ridgelines where extensive hilltop and ridgeline settlement started to develop during the Early Holocene. Radiocarbon dates place much of this settlement in the Middle Holocene (7,000–3,000 BP), but one site clearly dates to the Late Holocene (CA-SCrI-672). Midden constituents resemble the elevated sites on the north side of Santa Rosa (Clifford 2001). One of these interior sites is massive (40,000 m2) and occupies a flat ridgetop running from east to west along one of the northernmost branches of the Scorpion drainage. Excavations at two locations revealed densely packed shell midden up to 2 m deep and comparable in many ways to coastal midden sites on the islands. Four radiocarbon dates (upper and lower parts of the deposit) range from 4,360 to 3,600 BP. In the absence of a large Middle Holocene village along the coast of eastern Santa Cruz, it is at least plausible that this interior site served as a primary residential base and that marine resources, primarily California mussels and large barnacles, were harvested along the coast at logistical encampments and carried back to this location. The size and extent of this site supports this idea, but house floors were not encountered during our smallscale excavations and seasonality data suggest periodic, rather than permanent, occupation (see below). More substantial excavations will be required to determine if this site was a short-term residence used frequently or a longer-term residential base that approached a primary or secondary village in stature. On Santa Rosa, the largest population centers were located on the north coast, but smaller numbers of people also occupied the southern, more rugged side of the island during the Middle Holocene. Our best view of settlement on the south side comes from a systematic survey of Jolla Vieja Canyon and adjacent ridge systems (York 1996). The main part of this canyon extends north–northwest from the south coast of the island to a large confluence (~4 km inland), where a series of smaller drainages flowing out of the hills from the north and west join together. The canyon walls are steeply sloped at this juncture, rising quickly (50%–60% grade) to an elevation of 1,600 ft. Most of the sites in this drainage were relatively ephemeral and appear to date to the Late Holocene (York 1996). However, it seems that the main confluence of this drainage was an important loci of more permanent Middle Holocene settlement. The archaeological site at this location, first described by Jones in 1901 (Heizer and Elsasser 1956) and later by Orr (1968), is 200,000 m2 in size and consists of a series of rockshelters on both sides of the drainage that spill middens from their mouths to

TERMINAL PLEISTOCENE TO MIDDLE HOLOCENE

145

create a complex array of shell lenses exposed in the rivercut below. Several rock art panels are present in one of the rockshelters (Heizer and Elsasser 1956; Orr 1968; York 1996) and a bedrock mortar occurs next to the most central cave in the complex. A prominent feature at the site is a deeply stratified (3 m) deposit exposed in the northern tributary of the drainage (figure 20A). Large red abalone shells are mixed with California mussel shells at the base of the deposit, and the upper deposits contain small black abalones and a more diverse range of shellfish species. Radiocarbon dates indicate that the site was used from 7,250 to 420 BP, with the lowest 2 m dating to the Middle Holocene. Unlike the north coast, it is possible that this lowland settlement served as a primary base, perhaps a secondary village, for logistical forays to the coast. Of the Middle Holocene site types on the northern Channel Islands, the most common are thin lenses of shell along the coast, interpreted here as logistical encampments used for extracting and processing shellfish. Most of these are positioned on rocky stretches of coast or on headlands overlooking rocky intertidal habitats. Many of these sites consist of laterally extensive, but thin, pavements of abalone and California mussel shells (figure 20B). Fish and sea mammal remains are rare at these locations and tool assemblages are simple, usually consisting of expedient flakes manufactured from locally available volcanic rocks and chert. Red abalone middens are the most visible logistical encampments on the northern Channel Islands dating to between 7,500 and 4,500 BP (Glassow 1993b). The shells of this species are large (10–15 cm wide) and far less common in middens dating to after 4,500 BP. Most of these sites are composed of a single component and represent short-term, but intensive, collecting and processing events. When red abalone shells are present within multicomponent sites (e.g., CA-SCrI-333, CA-SRI-147, CA-SMI-528), they are often found in basal levels and are replaced in upper, more recent, strata by the smaller black abalone species or more diverse shellfish assemblages that include a range of smaller species (H. cracherodii; Kennett 1998; Walker et al. 2000; Wilcoxon 1993). The cool waters surrounding San Miguel Island provided more favorable conditions for red abalone later in time and small shells of this species are found in the Late Holocene, sometimes along with evidence for the manufacture of abalone shell fishhooks and beads (Walker et al. 2000). Unlike today, it is possible that red abalone was readily available in some intertidal habitats on the northern Channel Islands between 7,500

figure 20. Photographs of Middle Holocene sites: (A) River-cut exposure showing deeply stratified deposits (3 m) at CA-SRI-147, a large site at the main confluence of Jolla Vieja canyon on southern Santa Rosa Island. Lowest 2 m of site date to the Middle Holocene (photo by D. Kennett); (B) red abalone midden (CA-SRI-528) located on the western end of San Miguel Island near Point Bennett (photo by S. Spaulding).

TERMINAL PLEISTOCENE TO MIDDLE HOLOCENE

147

and 4,500 BP (Glassow 1993b). The primary distribution of this species extends north into the cooler waters along the northern California coast, where it is found in the intertidal zone today. Glassow (1993b) suggested that the presence of red abalone in Middle Holocene middens was indicative of cooler sea-surface temperatures during this interval. Oxygen isotopic analysis of California mussel shells (M. californianus) associated with one red abalone midden (CA-SCrI-333) indicated that water temperatures were 2°C cooler when the midden formed (~5,700 BP; Glassow et al. 1994). These data are somewhat consistent with the new paleowater-temperature data for the region showing cooler than average seasurface temperatures and high marine productivity between 6,400 and 5,800 BP (see figure 11). However, this cool-water interval appears to be unique in the Middle Holocene, which generally exhibits warmer-thanaverage sea-surface temperatures. Periodically, warm sea-surface temperatures are also suggested by diverse shellfish assemblages at Punta Arena (southern Santa Cruz; Sharp 2000). One possible explanation for the persistence of red abalone in intertidal habitats between 7,500 and 4,500 BP is the evidence for infrequent El Niños (short-term warm-water events not visible in lower-resolution climate records), as suggested by some paleoclimatic records from the eastern Pacific (Sandweiss et al. 2001). Another possible explanation is that this large abalone species was procured subtidally through diving and was never available in the intertidal zone (Sharp 2000). The disappearance of red abalone from middens on the northern Channel Islands was likely a product of changing environmental conditions (perhaps increases in El Niño frequency), coupled with intensive predation on this larger abalone species. The geographic position of middle Holocene settlements on the coast and in the interior suggests the complementary use of marine resources and plant foods. Fish, sea mammal, and bird bones are all found in residential middens dating to the Middle Holocene, but abalone and California mussel shells are the most numerous midden constituents. Quantitative midden constituent data from coastal residences, temporary encampments, and interior settlements on San Miguel and Santa Rosa islands also indicate that shellfish provided the primary meat source (table 13; Kennett 1998; Vellanoweth et al. 2000). The dietary importance of shellfish during the Middle Holocene is also evident in coastal residential middens and interior sites on Santa Cruz Island (Clifford 2001; Glassow 1980, 1993a). The strategic position of interior settlements dating to the Middle Holocene suggests the complementary use of coastal sagebrush and grassland communities

148

TERMINAL PLEISTOCENE TO MIDDLE HOLOCENE

(map 11). It is also possible that acorns were harvested from small stands of island oaks on Santa Cruz and Santa Rosa. Although substantial changes in the distribution of plant communities have occurred on the northern Channel Islands in the past 200 years, the current distribution provides some insight into the importance of different types of plants. On the northern Channel Islands, wind is one of the primary forces shaping plant geography. Coastal sagebrush occurs along drainage walls and behind hills, slightly sheltered from the wind, and grasslands predominate in more exposed portions of the island. On Santa Rosa Island grasslands dominate on the northwest side of the island where the winds are stronger and the topography is more gentle. Small stands of oaks are only available in sheltered canyons in the interior. Middle Holocene hilltop/ridgeline sites in the Arlington Canyon area are distributed evenly along the exterior edge of Arlington Canyon, apparently associated with grassland communities. Topographic relief in Cañada Verde is more variable, and interior sites cluster on hills near coastal sagebrush communities. Large lithic scatters occur in the lee of these hills where coastal sagebrush communities persist today. Grasslands and coastal sagebrush both contain plants that produce edible seeds and tubers. Some of these plants were most productive during spring (blue dick bulbs and seeds), and others were more productive in fall (acorns). The importance of grasses and sagebrush during the Middle Holocene is supported by the presence of milling equipment and digging-stick weights in burials and at sites in the interior (Kennett 1998; Orr 1968). Osteological data from island burial populations also support the hypothesis that various plant foods were an important part of the Middle Holocene diet (6,000–3,000 BP; Walker 1978; Walker and Erlandson 1986). The presence or absence of dental caries in prehistoric skeletons provides a proxy for the relative dietary contribution of carbohydrates (plants) and protein (meat). In fact, ethnographic and clinical studies indicate a direct link between high caries rates and carbohydrate-rich diets (Pederson 1938). Middle Holocene populations living on the Channel Islands had a high frequency of dental caries (80%) compared with later populations (Walker and Erlandson 1986), suggesting that plant foods (e.g., roots and tubers) played an important dietary role during the Middle Holocene. Stable nitrogen and carbon isotopic analysis of burial populations also provide information about the Middle Holocene diet on the northern Channel Islands (Goldberg 1993; Walker and DeNiro 1986). Nitrogen and carbon isotopic analysis

Table 13. Selection of faunal data (fish and shellfish) from deposits dating to the Middle Holocene on the northern Channel Islands Fish Site No. CA-SRI-147 CA-SRI-116 CA-SRI-116 CA-SCrI-277 CA-SCrI-146 CA-SMI-528 CA-SRI-109 CA-SRI-147 CA-SCrI-109 CA-SCrI-369 CA-SRI-116 CA-SCrI-109 CA-SCrI-292 CA-SMI-528 CA-SCrI-363 CA-SCrI-236 CA-SRI-50 CA-SCrI-363 CA-SRI-50 CA-SRI-5 CA-SRI-50 CA-SRI-147 CA-SCrI-292 CA-SCrI-277 CA-SCrI-369

Shellfish

Provenience

Age (1 sigma BP)

Weight (g)

Meat

%

Weight (g)

Meat

%

References

Unit 1, 290–300 cm Unit 1, 60–65 cm Unit 1, 65–70 cm Col. 1, 152–163 cm Col. 1, 0–19 cm Stratum 3, 660–690 cm Unit 1, surface, 10 cm Unit 1, 230–240 cm Col. 2, 104–119 cm Col. 1, 100–116 cm Unit 1, 10–20 cm Col. 1, 121–132 cm Col. 1, 50–57 cm Stratum 2, 530–551 cm Col. 1, 0–16 cm Col. 1, 235–250 cm Unit 2, 10–20 cm Col. 1, 20–29 cm Unit 2, 20–30 cm Unit 1, 50–70 cm Unit 1, 30–40 cm Unit 1, 100–110 cm Col. 1, 38–50 cm Col. 1, 0–20 cm Col. 1, 12–27 cm

7350(7252)7185 6303(6259)6175 6303(6259)6175 6730 (uncalib) 6028(5883)5693 5901(5844)5718 5770(5699)5612 5700(5621)5572 5655(5511)5320 5478(5312)5221 5300(5227)5023 5287(5037)4841 5286(4870)4653 4942(4837)4802 4991(4809)4524 4969(4835)4783 4818(4708)4552 4846(4618)4393 4790(4611)4512 4769(4574)4493 4490(4395)4281 4247(4098)3967 4082(3833)3627 3445(3278)3071 2752(2668)2355

0.70 2.40 0.50 1.89 0.08 1.40 2.00 0.40 0.89 0.28 12.80 1.75 11.20 0.20 0.21 0.00 2.00 0.86 5.20 2.50 0.90 8.20 7.87 0.32 1.93

19.39 66.48 13.85 52.35 2.22 38.78 55.40 11.08 24.65 7.76 354.56 48.48 310.24 5.54 5.82 0.00 55.40 23.82 144.04 69.25 24.93 227.14 218.00 8.86 53.46

2.66 14.50 8.81 2.62 0.30 6.63 6.36 2.97 2.78 0.92 49.52 3.19 17.59 1.99 0.52 0.00 8.38 1.60 13.04 8.14 1.83 29.88 10.59 0.62 4.86

2134.40 1180.90 432.00 5861.82 2197.89 1644.20 2455.80 1088.90 2600.00 2527.78 1088.60 4436.36 4377.14 821.80 3375.00 1333.33 1824.00 4407.47 2893.10 2353.50 4024.40 1605.60 5546.67 4312.00 3155.56

708.62 392.06 143.42 1946.12 729.70 545.87 815.33 361.51 863.20 839.22 361.42 1472.87 1453.21 272.84 1120.50 442.67 605.57 1463.28 960.51 781.36 1336.10 533.06 1841.49 1431.58 1047.65

97.34 85.50 91.19 97.38 99.70 93.37 93.64 97.03 97.22 99.08 50.48 96.81 82.41 98.01 99.48 100.00 91.62 98.40 86.96 91.86 98.17 70.12 89.41 99.38 95.14

Kennett 1998 Kennett 1998 Kennett, 1998 Glassow 1993a Glassow 1993a Kennett 1998 Kennett 1998 Kennett 1998 Glassow 1993a Glassow 1993a Kennett 1998 Kennett 1998 Kennett 1998 Kennett 1998 Glassow 1993a Glassow 1993a Kennett 1998 Glassow 1993a Kennett 1998 Kennett 1998 Kennett 1998 Kennett 1998 Glassow 1993a Glassow 1993a Glassow 1993a

note: 1 sigma range with mean intercept in parentheses. Calculated by Calib 3.0.3 (Stuiver and Reimer 1993; see Kennett 1998).

map 11. Distribution of coastal villages, interior ridgeline/hilltop settlements and lithic scatters on the north coast of Santa Rosa Island relative to the modern distribution of grassland and coastal sage scrub (map produced by D. Kennett with the assistance of J. Bartruff).

TERMINAL PLEISTOCENE TO MIDDLE HOLOCENE

151

of human bone is a proven method for determining the relative proportion of marine and terrestrial food in the diets of prehistoric individuals (DeNiro 1985). In general, individuals from the islands have enriched 15N and 13C isotopic values compared to people who lived on the mainland, indicating a greater dependence upon marine foods (Walker and DeNiro 1986). The importance of marine foods also increased through time on the islands (Goldberg 1993; Walker and DeNiro 1986). Goldberg (1993) analyzed human bones from two cemeteries on the northern Channel Islands dating to the Middle Holocene: CA-SRI-41 and CA-SCrI-333. Both studies also found that statistically significant dietary differences existed between men and woman during the Middle Holocene and suggested that women had greater access to plant food than men. The spatial patterning of Middle Holocene archaeological sites on the northern Channel Islands suggests that people established villages on sections of rocky coastline near springs or well-watered drainages on Santa Rosa and Santa Cruz. Cemeteries at some of these villages suggest that they were anchored to these productive locations, but interior residences and logistical encampments indicate that people foraged widely from these centralized locations (see map 10). In some instances this appears to have required the temporary relocation of people, perhaps smaller family units, to residences in the interiors of the larger islands, probably to exploit seasonally available plant foods. Shellfish harvesting profiles (18O) at a limited number of coastal villages, interior residences, and logistical encampments provide additional evidence for central place foraging during the Middle Holocene (figure 21; see Kennett 1998). Oxygen isotopic analysis of mollusk shells is a wellestablished approach for determining the season of shellfish harvest (see Kennett and Voorhies 1996; Kennett 1998). These data suggest that shellfish were harvested during most seasons at primary and secondary villages (CA-SRI-5, -41, -116), the expected pattern if people were living at these locations throughout the year. Compare these data to the highly seasonal harvesting profile (summer/fall) from CA-SRI-109, a red abalone midden located at Carrington Point on northeastern Santa Rosa (figure 21D), interpreted here as a logistical encampment used for harvesting and processing shellfish. Harvesting profiles from two interior sites, CA-SRI-50 (E) and CA-SCrI-608 (F), indicate that shellfish were transported to these locations during the winter/spring and summer/fall, with emphasized use during winter/spring. Evidence for site use during the winter/spring could be related to the exploitation of

CA-SRI-5, Primary Village 4769(4574)4493

CA-SRI-41, Primary Village 3826(3701)3615

A

B

Winter/ Spring

10

11

Winter/ Spring

Summer/ Fall

12

13

14

15

16

17

11

12

Summer/ Fall

13

14

16

17

18

CA-SRI-116, Secondary Village 6303(6259)6175

CA-SRI-109, Logistical Encampment 5770(5699)5612

C

D

Winter/ Spring

12

13

Summer/ Fall

14

15

Winter/ Spring

16

17

10

11

12

Summer/ Fall

13

CA-SRI-50, Interior Residence 4818(4708)4552

Winter/ Spring

Summer/ Fall

13

14

16

17

18

F

Winter/ Spring

12

15

C-SCrI-608, Interior Residence 3715(3634)3559

E

11

14

Water Temp (°C)

Water Temp (°C)

10

15

Water Temp (°C)

Water Temp (°C)

15

Water Temp (°C)

16

17

18

10

11

12

13

Summer/ Fall

14

15

16

17

18

19

20

Water Temp (°C)

figure 21. Seasonal shellfish harvesting profiles for six sites from the northern Channel Islands dating to the Middle Holocene. Each dot represents the water temperature of one mussel shell at the time it was collected prehistorically. This is based on an oxygen isotopic measurement of the final growth increment of each archaeological mussel shell. Note: Ages for each site are shown as 1 sigma ranges with the mean intercept in parentheses, calculated by Calib 3.0.3 (Stuiver and Reimer 1993). (Original data in Kennett 1998; drafted by D. Kennett.)

TERMINAL PLEISTOCENE TO MIDDLE HOLOCENE

153

plant foods (e.g., blue dick bulbs) during late winter and early spring. The transportation of shellfish to CA-SRI-50 throughout much of the year suggests the use of this location in spring and fall when seasonal plant foods were most available or the use of certain plants that were available throughout the year (e.g., Opuntia). Although the available data sets for Middle Holocene settlement and subsistence are incomplete, they generally support the hypothesis that small groups of people (perhaps 50 to 100) had established relatively permanent villages at optimal locations on the larger islands of Santa Rosa and Santa Cruz. The maximum island population probably did not exceed 400–600 people. One can only imagine the intricacies of ceremonial and social life at these villages, aspects of life that the archaeological record is virtually silent on. If burial data are a true reflection of social differences between individuals, then it appears that certain people in island Chumash society were distinguishing themselves during their lifetimes. Some burials are quite elaborate and contain hundreds of burial offerings including beads, spangles, and an array of shell and bone pendants (King 1990; see also Erlandson and Rick 2002a). The absence of elaborate infant burials on the islands suggests that leadership positions were not inherited, but achieved during one’s lifetime. Island villages were likely tied together through marriage alliances, and materials from the coastal mainland suggest that trade and marriage alliances extended to the coastal mainland (King 1990; Orr 1968; Rick et al. 2001b). What is clear from the record is that these people were skilled at crafting shell, stone, and bone items for ceremonial and subsistence purposes. They were also consummate fisherfolk, periodically fishing and hunting marine mammals, and using boats to forage widely through the island chain, intensively harvesting and processing shellfish for return back to more primary villages. These maritime traditions established the foundations for later developments in island Chumash society.

chapter 7

Late Holocene Record

Substantial changes in subsistence, demography, economy, and sociopolitical organization occurred on the northern Channel Islands during the Late Holocene (3,000–200 BP). Much of the archaeological work over the past century has focused on sites dating to between 650 and 200 years ago, the so-called Late or Canaliño Period (Arnold 2001; Erlandson and Rick 2002a; Heizer and Elsasser 1956; Kennett 1998; Orr 1968; Rogers 1929). It is during this interval that the economic, social, and political complexity associated with Chumash society is most evident archaeologically throughout the Santa Barbara Channel region. Indeed, many of the named historic villages on the northern Channel Islands were occupied through this interval, suggesting a certain degree of cultural continuity between the late prehistoric and early historic periods (see chapter 5). These deposits exhibit the best evidence for settled village life, intensive fishing, production of nonfood trade items, and intensive intervillage exchange. The aim of this chapter is to demonstrate that the cultural developments evident in the Late Holocene have roots in the Early and Middle Holocene, but that many of the characteristics associated with historic Chumash society emerged after 3,000 BP. The pace of cultural change also increased dramatically between 1,300 and 650 years ago, just prior to their solidification. Why these behaviors first developed and became dominant is a complex social and ecological problem, and this question is taken up in the final chapter.

154

L AT E H O L O C E N E

155

Population Growth and Demographic Expansion The demographic trends evident in the Early to Middle Holocene records continued into the Late Holocene, and there is some evidence for more rapid increases in island population after 1,300 BP. Population growth is suggested by larger numbers of coastal villages after 3,000 years ago (map 12 and figure 22C) and by increases in the total number of radiocarbon dated site components during the Late Holocene (figure 22D). Demographic expansion parallels evidence for decreases in body size in both men and women (stature; figure 22B) and more evidence for poor health, as indicated by increasing incidences of cribra orbitalia and periosteal lesions, in skeletal material through time (figure 22A; Lambert 1994, 1997). These demographic changes also parallel increases in diet breadth, resource intensification, heightened production of trade items, increases in interpersonal violence, and the emergence of social hierarchies (see below). A number of interpretive challenges are presented when using radiocarbon dates to estimate changes in population, particularly subtle fluctuations that push the limits of accuracy and precision inherent in radiocarbon chronologies (50–100 years; see Erlandson et al. 2001 for details). These data are used here only as a general proxy for demographic fluctuations through the Holocene, rather than for targeting specific intervals of population increase or decline. Five hundred and fifty-one radiocarbon dates are now available for the northern Channel Islands, and a large percentage of these dates come from the past 2,000 years (215). This suggests a threefold increase in population on the islands from the Middle to the Late Holocene, a time span when the number of primary villages on these islands expanded from ~6 to 22, the latter at the time of historic contact. Gradual increases in the number of radiocarbon dated components are evident after 1,800 years ago, and two peaks are evident in this record, one centered on 1,400 BP and the other on 600 BP. Decreases in the number of dates after 200 years ago are related to the collapse of island Chumash populations during the mission period, when the islands were finally, and totally, abandoned owing to the devastating effects of introduced diseases and the disintegration of the regional economy (ca. AD 1822; see chapter 5). More subtle decreases in the number of radiocarbon dated components also occurred during the protohistoric period (450 to 200 BP) and provide circumstantial evidence for a population decline related to the

map 12. Distribution of known Late Holocene village sites on the northern Channel Islands (map produced by D. Kennett with the assistance of J. Bartruff).

figure 22. General trends visible in the archaeological record on the northern Channel Islands during the past 3,000 years: (A) changes in health problems through time (Lambert 1994); (B) changes in body size (stature) through time based on femur length (Lambert 1994); (C) number of villages on the northern Channel Island per 100 years; (D) Frequency of radiocarbon-dated components on the northern Channel Islands as a proxy for demographic changes (Ka  thousand years) (drafted by D. Kennett).

158

L AT E H O L O C E N E

transmission of Old World diseases to island Chumash populations during the initial stages of European exploration, prior to the establishment of missions in the region (Erlandson et al. 2001, 11). Significant changes in the distribution of settlements and cemeteries also occurred on the northern Channel Islands after about 3,000 BP. The available chronological and typological information for Late Holocene archaeological sites on the northern Channel islands are presented in tables 14–17 (also see Kennett 1998; Kennett and Conlee 2002). Chronological information is based on radiocarbon dates and time-sensitive artifacts. All of these sites appear to be primary villages that were occupied long enough to develop substantial midden deposits. Sites have been divided into four primary chronological phases; primary villages occupied between 3,000 and 1,300 BP, 1,300 and 800 BP, 800 and 650 BP, and 650 and 200 BP. These chronological categories generally correspond to the late Early/early Middle Periods, the late Middle Period, the Middle to Late Period Transition, and the Late Period of Santa Barbara Channel prehistory (Arnold 1992a; Erlandson and Colten 1991; Kennett 1998; King 1990; see chapter 4). Population increase during the Late Holocene parallels the expansion of primary village locations around the coasts of Santa Cruz and Santa Rosa and along the north coast of San Miguel (Arnold 2001; Kennett 1998; Kennett and Conlee 2002; Munns and Arnold 2002; Peterson 1994, 2000; see map 12). By the latest Holocene, villages on Santa Cruz were positioned on many of the primary anchorages around the island (Arnold 1987, 2001; Munns and Arnold 2002), including those along the rugged northern side, where coastal access is limited and the countryside is relatively precipitous. Primary village locations along the well-watered north coast of Santa Rosa, that were occupied during the Middle Holocene, continued to be important settlement loci, but settlements expanded to the mouths of smaller drainages along the southern shore of the island (Kennett 1998). Three primary loci of Late Holocene settlement developed on San Miguel, one on the eastern side of Cuyler Harbor (CA-SMI-162/163), another in the vicinity of Otter Cove (northwest; CA-SMI-468, -470), and yet another on the western end of the island (CA -SMI-525, -528, -602). The settlements on western San Miguel represent the first primary villages in the vicinity of the modern-day sea mammal rookery located near Point Bennett (see chapter 3, figure 6E). Primary villages appear to be absent from the poorly watered southern coast of San Miguel, but smaller middens indicative of Late Holocene

table 14. Late Holocene sites dating between 3,000 and 1,300 BP Site No.

Location

CA-SRI-41

Cañada Verde

CA-SRI-62

Johnson’s Lee

CA-SRI-96 CA-SRI-1 CA-SRI-2 CA-SRI-3

China Camp Garanon Skull Gulch Tecolote/Arlington

CA-SRI-4

Tecolote/Arlington

CA-SRI-19 CA-SRI-28 CA-SRI-31 CA-SRI-173 CA-SRI-432 CA-SRI-587 CA-SCrI-1 CA-SCrI-127 CA-SCrI-145

Dry Canyon China Camp Bee Canyon Arlington Canyon Ford Point Cañada Verde Coche Prietos Anchorage Malva Real Anchorage N. or Kinton Point

CA-SCrI-191

Christy Beach/Ranch

CA-SCrI-195

Forney’s Cove

Radiocarbon (1 sigma) 3176(3013)2866 2284(2144)2076 2144(2056)1956 2670(2479)2357 1836(1769)1694 2308(2188)2103 1693(1561)1471 2872(2746)2357 2349(2317)2139 3105(2745)2203 2955(2862)2781 1483(1386)1313 3193(3051)2911 2006(1882)1775 1710(1603)1511 2748(2704)2554 2532(2334)2180 1858(1727)1601 1702(1527)1345 1618(1448)1289 1515(1386)1288 1810(1625)1475 1451(1386)1288 1694(1553)1511 2332(2144)1961

References Orr 1968 Kennett 1998 Kennett 1998 Kennett 1998 Erlandson 1994 Orr 1968 Orr 1968 Erlandson 1994 Orr 1968 Kennett 1998 Kennett 1998 Kennett 1998 Orr 1968 Kennett 1998 Kennett 1998 Glassow 1980, 1993a Glassow 1980, 1993a Glassow 1980, 1993a Glassow 1980, 1993a Glassow 1980, 1993a

table 14. (continued) Site No.

Location

CA-SCrI-236

Christy Beach/Ranch

CA-SCrI-240

Prisoners Harbor

CA-SCrI-333

Fraser Point

CA-SCrI-369 CA-SCrI-474

Twin Harbors Pozo Beach

CA-SCrI-615 CA-SMI-488 CA-SMI-492 CA-SMI-503

Little Scorpion NW San Miguel Island NW San Miguel Island NW San Miguel Island

CA-SMI-504 CA-SMI-525

NW San Miguel Island Point Bennett

CA-SMI-528

Point Bennett

Radiocarbon (1 sigma) 1557(1404)1292 1704(1557)1454 1711(1535)1354 2097(1942)1818 2345(2148)1986 3151(2910)2766* 1383(1299)1267 2319(2138)1995 2857(2776)2753* 2752(2668)2355 1527(1402)1309 1857(1785)1695 2705(2639)2457 1361(1292)1244 2423(2329)2285 2680(2493)2370 3036(2933)2846 3139(3013)2906 3210(3105)2981 1502(1395)2906 1558(1467)1342 1466(1361)1292

References Glassow 1980, 1993a Arnold 1987, 1990a, 1991, 1992a

Wilcoxon 1993

Glassow 1980, 1993a Arnold 1991, 1992a Arnold and Munns 1994 Kennett 1998 Walker and Snethkamp 1984 Walker and Snethkamp 1984 Walker and Snethcamp 1984 Kennett and Conlee 2002 Walker and Snethkamp 1984 Walker and Snethkamp 1984 Walker et al. 2000

note: 1 sigma range with mean intercept in parentheses. Calculated by Calib 3.0.3 (Stuiver and Reimer 1993; see Kennett 1998). *Dates show overall time span but do not include all radiocarbon dates for the site.

table 15. Late Holocene sites dating between 1,300 and 800 BP Domestic Features Site No. CA-SRI-2

2

Size (m ) 50,000

CA-SRI-6

Associated Artifacts

Pits

Berms

WB

TRP

JF

LP

Radiocarbon (1 sigma)

p

p

p

p

p

p

1240(1164)1064





p

p





1320(1260)1180 1068(984)927 1483(1386)1313 1414(1340)1286 1313(1264)1183

CA-SRI-15 CA-SRI-28 CA-SRI-31

28,000 9,200

— — p

p p p

p p p

p — p

— — —

p — —

CA-SRI-40

17,500

p

p

p

p

p

p

CA-SRI-41

105,000



p





p



12,000 10,000 6,612 14,177

p p — p

p p p p

p p p p

p — p p

p — — —

p — — —

5,250

— —

— p

p p

p p

— —

— —

CA-SRI-60 CA-SRI-77 CA-SRI-85 CA-SRI-130/131 CA-SRI-116 CA-SMI-468

1260(1197)1134

1281(1062)920 1161(918)662 1313(1264)1183 1276(1225)1146 909(797)707

References Orr 1968 Kennett 1998 Orr 1968 Kennett 1998 Kennett 1998 Kennett 1998 Kennett 1998 Orr 1968 Kennett 1998 Orr 1968; King 1990 Kennett 1998 Kennett 1998 Kennett 1998 Kennett 1998 Kennett 1998 Kennett 1998 Kennett 1998

table 15. (continued) Domestic Features

Associated Artifacts

Size (m2)

Pits

Berms

WB

TRP

JF

LP

Radiocarbon (1 sigma)

References

CA-SMI-503/504

20,425





p



p

p

1248(1147)1033

CA-SMI-510 CA-SMI-525 CA-SMI-528

65,000 54,000 55,000

— — —

— — —

— — p

— — —

p — p

— — p

CA-SCrI-127

14,700 Mound



p

p





1259(1182)1107 1280(1225)1137 1264(1182)1092 1466(1361)1292 1223(1036)925

Walker and Snethkamp 1984 Kennett and Conlee 2002 Walker and Snethkamp 1984 Walker and Snethkamp 1984 Kennett 1998

Site No.

CA-SCrI-191

6,072

?



p

p





CA-SCrI-195

8,536

p

p

p







CA-SCrI-240

14,154





p

p

p

p

CA-SCrI-257 CA-SCrI-333

? 27,000

— p

— p

p —

p —

— p

— —

1175(1110)1062 1275(1211)1164 1515(1386)1288 1057(953)922 1288(1218)1064 1416(1345)1293 1451(1326)1256 983(913)704 1168(1062)966 981(916)823 1279(1168)1057* 1383(1299)1267

Glassow 1980, 1993a Arnold 1991 Arnold 1991, 1992a Arnold and Munns 1994 Glassow 1980, 1993a Wicoxon 1993

Arnold 1990a, 1990b, 1991, 1992a

Arnold 1991, 1992a Wilcoxon 1993; King 1990

CA-SCrI-396 CA-SCrI-474

? ?

— ?

— ?

— p

p p

— —

— —

CA-SCrI-495 CA-SCrI-504

? ?

— —

— —

— p

p p

— —

— —

CA-SCrI-506

?

p

p

p

p





CA-SCrI-507

?





p

p





886(736)695 1161(1036)939 1323(1285)1178 1527(1402)1309*

910(834)742

Arnold 1990b Arnold 1991

Arnold 1991, 1992a Arnold 1990a Kennett 1998 Arnold 1990a Kennett 1998 Arnold 1990a Kennett et al. 2000

notes: WB, Olivella-wall bead; TRP, trapezoidal microblade; JF, j-shaped fishhook; LP, leaf-shaped projectile point; p, present. 1 sigma range with mean intercept in parentheses. Calculated by Calib 3.0.3 (Stuiver and Reimer 1993; see Kennett 1998).

table 16. Late Holocene sites dating between 800 and 650 BP Site No.

Location

CA-SRI-2

Skull Gulch

CA-SRI-15

Abalone Point

CA-SRI-97

China Camp

CA-SRI-85 CA-SCrI-191

Old Ranch Canyon (mouth) Christy Beach

CA-SCrI-192

Morse Point

CA-SCrI-240

Prisoners Harbor

CA-SCrI-330

Forney’s Cove

CA-SCrI-474

Pozo Beach

CA-SMI-468

Otter Canyon

Radiocarbon Dates 650(617)534 730(651)557 656(622)543 764(678)633 820(733)671 760(706)663 797(724)666 656(622)543 687(637)551 665(652)564 (equivocal) 782(666)552

Closest Perennial Stream Tecolote

References

Garanon Canyon

Orr 1968 Kennett 1998 Kennett 1998

Unnamed

Kennett 1998

Old Ranch Canyon

Kennett 1998

Cañada Christy

Arnold 1991

Johnson’s Canyon

Arnold 1991, 1992a

885(737)655 (equivocal) 757(670)577

Prisoners Drainage

Arnold 1987, 1990b, 1991, 1992a Arnold 1991, 1992a

650(610)542 712(672)663 886(736)695 939(922)791 647(603)532 883(766)686 909(797)707

Posa Canyon

Arnold 1991, 1992a Arnold and Munns 1994

Otter Canyon

Kennett 1998

Cañada Christy

note: 1 sigma range with mean intercept in parentheses. Calculated by Calib 3.0.3 (Stuiver and Reimer 1993; see Kennett 1998).

table 17. Late Holocene sites dating between 650 and 200 BP Features Site No.

2

Associated Artifacts

Size (m )

Pits

Berms

TM

CC

LB

CF

CB

CA-SRI-2 Skull Gulch

50,000

p

p

p

p

p

p

p

CA-SRI-15 Abalone Point

28,000

p

p

p

p



p

p

CA-SRI-40 Cañada Verde CA-SRI-60 Becher’s Bay CA-SRI-62 Johnson’s Lee CA-SRI-84 Old Ranch Canyon CA-SRI-85 Old Ranch Canyon

17,500

p

p

p

p

p

p

p

468(392)303 515(477)315 650(617)534 730(651)557 550(511)474 656(622)543 764(678)633 299(264)154

6,000

p

p

p

p

p

p

p

299(264)154

62,500

p



p

p

p

p

p

12,240

p

p

p

p

p

p

p

1,875



p

p

p



p



24,000





p

p

p

p

p

Kennett 1998

625





p

p







Kennett 1998

p

p

p

p

p





CA-SRI-87 Old Ranch Canyon CA-SRI-88 Old Ranch Canyon CA-SRI-97 China Camp

9,100

Radiocarbon

507(468)417 656(622)543 687(637)551

AD 1677(1716)1848

Primary References Orr 1968 King 1990 Kennett 1998 Kennett 1998

Orr 1968 Kennett 1998 Rogers 1929 Kennett 1998 Rogers 1929 Orr 1968 Kennett 1998 Orr 1968 Kennett 1998

Orr 1968 Kennett 1998

table 17. (continued) Features Site No. CA-SRI-98 China Camp CA-SRI-130/131 Jolla Vieja CA-SRI-427 San Augustine CA-SRI-432 Ford Point CA-SMI-161/163 Cuyler Harbor CA-SMI-470 Otter Point CA-SMI-602 Point Bennett

CA-SCrI-1 Coche Prietos Anchorage CA-SCrI-191 CA-SCrI-236

Associated Artifacts

Size (m2)

Pits

Berms

TM

CC

LB

CF

CB

8,840

p

p

p

p







1,200

p

p

p

p







Orr 1968 Kennett 1998 Kennett 1998

34,000

p

p

p

p

p





Kennett 1998

6,600

p

p

p

p







Kennett 1998

4,000

p

p

p

p







463(312)293

Kennett 1998

7,373

p

p

p

p

p

p



Kennett 1998

7,200

p



p

p

p

p

p

6,580





p

p







248(125)0 253(125)0 532(491)440 527(491)447 395(294)260 243(115)0 150

11,476

p



p

p







Radiocarbon

473(413)293 525(506)472

Primary References

Walker et al. 2000

Glassow 1980, 1993a Peterson 1994 Arnold 1990b, 1991 Arnold 1990b, 1991, 1992a Glassow 1980, 1993a

CA-SCrI-257 Christy Beach/ Ranch

530(512)496 521(514)508 537(524)516 550(535)514 664(597)534 (SCrI-236) 312(293)0 655(610)534 675(583)529 782(666)552 150 501(303)0 643(545)514 638(532)500

Arnold 1990b, 1991, 1992a

CA-SCrI-192 Morse Point

10,863

p



p

p





p

CA-SCrI-195 Forney's Cove

8,536

p

p

p

p

p

p

p

14,154





p

p

p

p

p

3,000

p



p

p







664(597)534 536(480)402

Arnold 1987, 1990a, 1990b 1991, 1992a

8,600

p



p

p





p

264(141)0 (SCrI-328) 309(148)0 (SCrI-328) 518(465)295 (SCrI-328) 551(510)326 (SCrI-328) 623(540)514 (SCrI-330) 769(690)577 (SCrI-330)

Glassow 1980, 1993a Arnold 1990b, 1991, 1992a

CA-SCrI-240 Prisoners Harbor CA-SCrI-306 CA-SCrI-392, -416 CA-SCrI-420, -421, -422 China Harbor CA-SCrI-328, -330 Forney’s Cove

Glassow 1980, 1993a Arnold 1990b, 1991, 1992a

Glassow 1980, 1993a

table 17. (continued) Features Site No.

Size (m2)

CA-SCrI-434 CA-SCrI-436 CA-SCrI-474 Pozo Beach CA-SCrI-475 CA-SCrI-423, -507, -615 Scorpion/Little Scorpion CA-SCrI-504, -505, -506 Smugglers Cove Smugglers Point

Associated Artifacts

Pits

Berms

TM

CC

LB

CF

CB

p p

— —

p p

p p

— —

— —

— —

Radiocarbon

650(610)542

10,100

? p

? p

p p

p p

— p

— p



p

p

p

p

p

p

p

254(132)0 (SCrI-507)

Primary References Arnold 1990b Arnold 1990b Arnold 1991, 1992a Arnold 1991 Arnold 1990b Johnson 1993 Kennett et al. 2000 Arnold 1990b Johnson 1982, 1993 Kennett et al. 2000

notes: p, present, TM, Triangular microblade (with retouch); CC, Callus-cup bead; LB, Olivella lipped bead; CF, Circular fishhook; CB, Concave base point arrow. 1 sigma range with mean intercept in parentheses. Calculated by Calib 3.0.3 (Stuiver and Reimer 1993; see Kennett 1998).

L AT E H O L O C E N E

169

settlement suggest logistical forays from primary villages positioned on the north coast. Small sites dating to the Late Holocene have also been identified on Anacapa Island (CA-ANI-8; Rozaire 1993), and probably represent logistical encampments used for exploiting the resources available on these islets (e.g., bird rookeries). The use of interior residences appears to be reduced with the expansion of primary villages around the exteriors of the larger islands after 3,000 BP. This suggests decreases in residential mobility, an overall reduction in foraging range, and a significant shift in the way plant foods were collected on the islands. There are very few substantial interior midden deposits on Santa Rosa and San Miguel dating to the Late Holocene (Kennett 1998; Kennett and Conlee 2002), and only one interior hilltop residence on eastern Santa Cruz has been radiocarbon dated to after 3,000 BP (Clifford 2001). Peterson (1994) reported Late Holocene age “residential bases” on the south side of Santa Cruz, but based on the descriptions of these sites they appear to have been occupied only for short periods of time. These types of sites (thin lenses of shell) occur elsewhere on the islands but are generally located in valley bottoms and are much more ephemeral than Middle Holocene residences on hilltops or ridgelines. These small sites probably accumulated during logistical forays into the interior to collect plants or were created along footpaths that connected Late Holocene island Chumash communities. The absence of substantial interior settlement after 3,000 BP suggests that people were becoming more tethered to coastal villages. Regional climate records indicate that these changes in settlement occurred when marine conditions were productive and cool (see figure 11) and terrestrial conditions were dry (see figure 12), the first in a series of millennial-scale climatic changes characterizing the Late Holocene. The shift also co-occurs with increasing evidence for intensified fishing and greater maritime trade relations with people living on the adjacent mainland coast (Arnold 2001; Erlandson and Rick 2002a; Kennett 1998; Kennett and Kennett 2000; Rick et al. 2001b). Short-term logistical forays (e.g., daily) to collect and process plants are suggested by thin lenses of shell exposed in the bottom of drainages (Clifford 2001; Kennett 1998; York 1996) and by the presence of globular mortars at certain interior locations. The strategic placement of cemeteries at interior locations also suggests a more rigid form of territoriality, and there is increasing evidence for the control of valuable resources by some communities (e.g., chert on eastern Santa Cruz; Arnold 1987, 2001; Kennett 1998; Kennett and Conlee 2002).

170

L AT E H O L O C E N E

Although communities expanded along the coasts of these islands, changes in the character of these settlements did not occur abruptly, but more subtle changes in site structure appear to have developed gradually through the Late Holocene. The character of coastal settlements on the islands between 3,000 and 1,300 BP does not appear to change substantially compared with the Middle Holocene (Kennett 1998). Domestic features (e.g., house depressions or floors) at sites dating to this interval are rare and artifact assemblages at these locations are no more diverse than in middens dating to the Middle Holocene. In fact, settlements along the coast dating to between 3,000 and 1,300 BP are difficult to identify because formal artifacts are rare and faunal assemblages do not contain large quantities of fish bone (table 14; Kennett 1998). The only way to securely identify sites dating to this interval is with radiocarbon dating, and it is likely that sites dating to this time are underrepresented in the record. On Santa Rosa Island, it appears that most of the coastal villages occupied during the Middle Holocene (SRI-5, SRI-41, SRI-116) continued to be used well into the Late Holocene (until ~1,300 BP). Continued use of cemeteries at some of these sites also indicates a certain degree of settlement continuity (Kennett 1998; King 1990). However, residential middens and cemeteries were also established at other locations, usually in association with smaller drainages and more varied coastal habitats. By the late Middle Period (1,300 BP), primary villages were evenly distributed around the perimeters of Santa Rosa and Santa Cruz, and along the north coast of San Miguel (Glassow 1993a; Kennett 1998; Kennett and Conlee 2002; Munns and Arnold 2002). There also appears to be a change in the character and structure of primary villages after this time that suggests even greater settlement stability. These changes are difficult to quantify because the large-scale excavations needed to better characterize them are sorely lacking. However, primary villages occupied during the latest Holocene (after 1,300 BP) are generally located on promontories or cliff faces close to springs or near the mouths of perennial streams. Compared with primary village sites dating to the Middle Holocene (CA-SRI-4, -5, -41), these villages tend to be more compact and not as laterally extensive (Kennett 1998). Evidence for relative stability after this time includes more substantial domestic features, larger and deeper midden deposits (figure 23A), and greater faunal and artifact diversity (table 15 and figure 24). Older excavations by Phil Orr at CA-SRI-2, a Late Holocene primary village site located at Skull Gulch on the north coast of Santa Rosa

figure 23. Photographs of Late Holocene archaeological sites on the northern Channel Islands. (A) Robert DeLong cleaning cliff face profile at CA-SMI-525, a well-stratified Late Holocene site on the west end of San Miguel Island near Point Bennett; (B) Photograph of CA-SRI-2, a Late Period village showing house depressions near promontory. (Photos by D. Kennett.)

figure 24. Selection of late Middle Period (1,300–650 BP) artifacts from the northern Channel Islands: (a, d) trapezoidal microblade cores; (b, c) trapezoidal microdrills; (e) triangular microblade without retouch; (f, g) j-shaped fishhooks; (h–m) Olivella-shell-wall beads; (n) Olivella barrel bead; (o–r) leaf-shaped arrow points. (Illustrations and layout by R. van Rossman.)

L AT E H O L O C E N E

173

(figure 23B), suggest that house pits on the surface of this site correspond to relatively formal house floors with postholes, and that at least one of these houses dates to ~1,000 BP (Orr 1968; House #3). More recent excavations at El Montón (CA-SCrI-333), a large village site near Frasier Point on western Santa Cruz, provide the most substantial look at site structure through this interval (Wilcoxon 1993). El Montón is a large (30,000 m2) shell mound that reaches 7 m in elevation above the marine terrace that it occupies. Circular-to-oval-house depressions (38 in total) are visible on the southern and eastern sides of the mound, and these structural features are surrounded by elevated refuse deposits. Excavation of four structures revealed house floors that were yellowish brown to light gray in color, and composed of ash, fragmented shells, and coarse beach sand. All of these features appear to occur in the upper meter of the deposit that has been dated to between 3,500 and 1,300 BP. Unfortunately, radiocarbon dates from individual house floors are not available and it is possible that all of the structural features tested are contemporary with the terminal age of the midden (1,400–1,250 BP). Regardless, these data suggest that substantial houses were in use by at least this time, a pattern that is consistent with other evidence for reduced settlement mobility and the consolidation of village locations by 1,300 BP. The consolidation of primary villages after 1,300 BP also corresponds to some notable changes in the location and structure of cemeteries. The large burial grounds located near primary villages on the north coast of Santa Rosa during the Early and Middle Holocene (CA-SRI-4, -5, -41) were not used extensively after this time (Orr 1968). Similar to the Middle Holocene, formal cemeteries were located next to, or within the confines of, primary coastal villages. The people buried in these newly established cemeteries were sometimes placed in discrete burial plots and other times jumbled together in burial pits (Orr 1968). Burials were more frequently placed at interior cave and hilltop locations away from the coast, a pattern that appears to develop sometime after 3,000 BP (Heizer and Elsasser 1956; Kennett 1998; Orr 1968). Chester King has also noted some interesting differences in burial patterns through the Holocene: The phases and subphases so far described (~8,000–1,300 BP) are mainly defined on the basis of cemeteries from which burial lots cannot be sorted into more than one time period. Beginning with phase M3 burials (after 1,300 BP), the burials found in island cemeteries in the sample I studied are from a number of phases. (Calendar years added; King 1990, 35)

174

L AT E H O L O C E N E

In other words, cemeteries used during much of the Middle Holocene were no longer used after 1,300 BP. New cemeteries were established at this time and were often used continuously into the historic period, supporting other data suggesting settlement continuity at certain locations between 1,300 and 200 BP. Although some coastal communities were relatively stable on the northern Channel Islands after 1,300 BP, there appears to have been a disruption in settlement between 800 and 650 BP at some locations (Arnold 1991, 2001; Kennett 1998; Munns and Arnold 2002). Only three sites on Santa Cruz (CA-SCrI-240, -191/-236, -474; table 16) show definitive evidence for occupation during this time (artifacts and radiocarbon dates; Munns and Arnold 2002, 136). There is also evidence for an occupational hiatus at several villages on Santa Cruz that were occupied before and after this time (CA-SCrI-330, -306, -191; Arnold 1991). A partial abandonment of the outer islands also appears to have occurred between 800 and 650 BP (Kennett 1998). Definitive evidence of the occupation of Santa Rosa during this interval comes from only four primary villages (CASRI-2, -15, -85, -97), and there is evidence from one of these (CA-SRI15; Abalone Point) of an occupational hiatus followed by the complete relocation of the residential compound from one side of the point to the other (Kennett 1998). On San Miguel Island, Late Holocene occupation persisted in the vicinity of Otter Harbor (CA-SMI-468), but there appears to be a lacuna in settlement between 1,000 and 650 BP at Point Bennett, where there is clear evidence for primary village settlement between 1,500 to 1,000 BP (CA-SMI-528) and again from 650 and 250 BP (CA-SMI-602; Kennett 1998; Walker et al. 2000). Paleoclimatic reconstructions for the region suggest that the years leading up to this interval were particularly dry and severe droughts may have been more frequent between 800 and 650 BP (Arnold 2001; Jones et al. 1999; Kennett and Kennett 2000; Munns and Arnold 2002; Raab and Larson 1997; Stine 1994). This could account for the hiatus in settlement at many of these locations. Indeed, the coastal communities that were most stable during this period were located near the largest drainages on each island (table 16; compare maps 12 and 3) and settlements that were abandoned during this interval (CA-SCrI-330; CA-SRI-15; CA-SMI-528/602) were positioned in more marginal locations with respect to freshwater availability (Arnold 2001; Kennett 1998; Kennett and Conlee 2002; Kennett and Kennett 2000; Munns and Arnold 2002). Late Period settlement stabilized on the islands after this disruption, and many of the villages occupied between 1,300 and 800 BP were

L AT E H O L O C E N E

175

reoccupied. Most villages were established near perennial streams, but this does not appear to be a necessary determinant of settlement location (Kennett 1998). Compared to the Middle Holocene, village locations varied greatly with respect to proximity to rocky coastlines (Arnold 1991; Kennett 1998), but access to at least a small section of beach was an important determinant of settlement location. Arnold (1991) has suggested that beaches were critical for landing plank canoes, the primary watercraft used in the region after 1,500 BP (Gamble 2002). Proximity to sandy beach may also have been important for acquiring Olivella biplicata, the primary species of shellfish used to make shell beads (see below). Shell lenses exposed in river cuts and in small rockshelters dating to this period suggest temporary rather than permanent occupation of the interior (Kennett 1998; York 1996). Sites occupied during the Late Period were generally located on promontories or cliff faces close to a spring or near the mouth of a perennial stream (figure 23B). House depressions are visible on the surfaces of most of these primary village locations, and the midden deposits are composed of thick, dark lenses of anthropogenically altered soil mixed with large quantities of marine shell and bone. Faunal assemblages at these village sites are overwhelmingly dominated by fish bone, and these deposits are often laden with temporally diagnostic artifacts (figure 25). In many instances, the surfaces of these sites are also covered with debris from the manufacture of Olivella-shell beads or chert debitage from the manufacture of microliths that were ultimately used for the production of these beads (Arnold et al. 2001; Kennett and Conlee 2002; Preziosi 2001). On Santa Rosa, a natural break in the landscape, usually an arroyo, gulch, or stream, separates these sites into sectors (e.g., CA-SRI-2, -40, -60, -97, -85, -87). In some instances, this included satellite communities that were positioned up to 500 m away from the coastal villages that they appear to be associated with (map 13; Kennett 1998; see below). Oxygen isotope seasonality data indicate that shellfish harvesting strategies changed and are consistent with the idea that settlement mobility decreased through the Late Holocene (3,000 to 250 BP; Kennett 1998). As discussed elsewhere (Glassow et al. 1994; Kennett 1998), the oxygen isotopic composition through the incremental growth of California mussel shells records seasonal changes in water temperature. The oxygen isotopic composition of the final growth increment represents the water temperature that a mollusk was living in just prior to harvest and, hence, is an approximate measure of when it was collected.

figure 25. Selection of Late Period (650–200 BP) artifacts from the northern Channel Islands: (a, b) large chert bifaces; (c–f) concave-base arrow points (Franciscan chert); (g, k) triangular microblade cores with retouch; (h–j) triangular microliths with retouch; (l) large chert drill; (m) shell-fishhook blank (M. californianus); (n–o) grooved-shank circular fishhooks (H. rufescens); (p) raptor talon pendant; (q–s) abalone-shell pendants; (t–z, bb) various Olivellashell beads; (aa, dd) mussel-shell beads (M. californianus); (cc, ee) abalone-shell (H. rufescens) beads. (Illustrations and layout by R. van Rossman.)

map 13. One-hour walking distances (cost surfaces) from coastal village sites occupied between 650 and 200 BP. Notice satellite communities associated with three of these coastal sites, and interior cemetery locations on Santa Rosa Island (map produced by D. Kennett with the assistance of J. Bartruff).

178

L AT E H O L O C E N E

Because mollusks are available and easy to harvest in the intertidal zone throughout the year, it is likely that they were always collected by some members of a group (children or elderly; Bird and Bliege Bird 2000). Therefore, these data can also provide information regarding settlement permanence or function within a broader settlement system (Deith 1983, 1985, 1988; Jones et al. 2002; Kennett and Voorhies 1996; Killingley 1981; Voorhies et al. 2002). A majority of the shellfish harvesting profiles from Late Holocene villages suggest the year-round exploitation of California mussels and a high degree of settlement stability (figure 26). The harvesting profiles dating to after 1,300 BP all show clear evidence for year-round collection (figure 26D–H). Yearround collection of mussels is less clear in the samples that were radiocarbon dated to before 1,300 BP. One shellfish harvesting profile from CA-SRI-1 (2,310–2,100 BP; see figure 26A), a site located on the north coast of Santa Rosa Island at Garanon Canyon, shows that California mussels were harvested throughout much of the year, but the profile appears to represent a bimodal, rather than continuous, collection strategy. The profile from another site (CA-SRI-62; 2,140–1,960 BP), located on the south side of Santa Rosa near Johnson’s Lee, indicates more focused harvesting during winter and springs months (figure 26B). Yet another site, positioned on the southwest coast of the island at Bee Rock (CA-SRI-31), suggests an emphasis on mussel collecting during winter and spring months, with minimal exploitation during the summer and fall (figure 26C). These data indicate that villages on the south side of Santa Rosa, dating prior to 1,300 BP, were less permanent and possibly shifted several times within a given annual cycle. These data, along with clear evidence for village stability on the south coast of Santa Rosa after 1,300 BP (figure 26D and G), appear to be consistent with the hypothesis that primary village locations were more stable after 1,300 BP. Oxygen isotope seasonality data are also in-line with osteological data that suggest both decreases in stature and increases in disease and nutritional stress through the Late Holocene (see figure 22A and B; Lambert 1994, 1997; Lambert and Walker 1991; Walker 1986). Cribra orbitalia and periosteal lesions provide two of the best skeletal indicators of malnutrition and disease in prehistoric populations and are used to gauge the nutritional stress associated with the emergence of more sedentary communities (Cohen and Armegellos 1984; Lambert and Walker 1991). Cribra orbitalia is a sievelike pitting in the upper portions of the eye orbits that has been linked to chronic iron-deficiency anemia

figure 26. Seasonal shellfish harvesting profiles for eight levels dating to the Late Holocene. Dots represent the water temperature when California mussels were harvested prehistorically. This is based on an oxygen isotopic measurement of the final growth increment of each shell. Note: Ages for each site are shown as 1 sigma ranges with the mean intercept in parentheses, calculated by Calib 3.0.3 (Stuiver and Reimer 1993). (Original data in Kennett 1998; drafted by D. Kennett.)

180

L AT E H O L O C E N E

in modern populations (White 1991). Walker (1986) argued that the high incidence of cribra orbitalia in the Santa Barbara Channel region prehistorically was probably associated with diarrhea caused by waterborne pathogens that became concentrated during dry periods. Periosteal lesions form on bones when the tissue surrounding them is traumatized or infected. Bacterial infection and traumatic injury are the primary causes of such injuries. The frequency of individuals with cribra orbitalia on the northern Channel Islands increased through time. Relative to the Middle Holocene, the incidence of cribra orbitalia in the Late Holocene is high, particularly after 1,300 BP when the overall frequency of infection was close to or greater than 50% in men. The frequency of periosteal lesions was also greater in the Late Holocene compared with the Early/Middle Holocene. The greatest frequencies occurred between 1,300 and 600 BP, suggesting a high degree of sedentism after this time (Lambert 1994, 1997).

Territoriality and Warfare Increases in island population levels, the establishment of new primary villages, and further reductions in settlement mobility appear to parallel increases in territoriality and more despotic forms of human behavior during the Late Holocene. The distribution of settlements on the northern Channel Islands during the Middle Holocene suggests that the foraging ranges, or territories, of each community were relatively large and included a wide range of island habitats (see map 10). This may have included shared foraging territories in island interiors and on the south side of Santa Rosa and San Miguel. Smaller, more rigid, territories appear to develop on these islands with increases in population and the expansion of primary villages during the Late Holocene. By the late Middle Period (~1,300 BP), primary villages were evenly distributed around the coasts of Santa Rosa and Santa Cruz, and along the north coast of San Miguel (Arnold 1987; Kennett 1998; map 12). As outlined above, many of these locations were occupied into the historic period and their names are known from mission records and informants familiar with the cultural geography of the island Chumash (chapter 5, map 6; Johnson 1982, 1993). The even distribution of these communities, along with the lack of intervisibility between them (see map 7), suggests that lineal-descent groups partitioned each island into smaller territories. Although mission records indicate that marriage alliances

L AT E H O L O C E N E

181

existed between these communities during the historic period, they were also in direct competition with one another for island resources through the Late Holocene. Primary villages on Santa Rosa and San Miguel dating after 1,300 BP were often strategically positioned on promontories near perennial water sources (see figure 23B). These positions afforded the best surrounding views along the coast and out to sea and suggest that the people living in these communities were concerned with possible outside threats, from people living either on the mainland or in other island villages. In some instances, these villages were composed of multiple residential loci, each commanding a different view of the landscape within the immediate vicinity of the settlement. For instance, all of the known primary villages occupied during the Late Holocene were located on the marine terrace that rings each island. Although these positions provided clear visibility along the coast, interior views were limited by a series of uplifted Quaternary terraces that rise up steeply behind many of these settlements. By the Late Period (~650 BP), several coastal communities on Santa Rosa Island had smaller residential compounds in elevated positions away from the coast (Kennett 1998). These residential compounds (satellite communities; map 13) were close to economically valuable plants but also fostered views of the interior that were not possible from coastal locations. The best example of this occurs at CASRI-40, a large Late Period (~650–200 BP) village positioned on a promontory near the mouth of Cañada Verde on the north coast of Santa Rosa Island (see map 12 for location). Two primary residential sectors, each with distinctive house depressions, are separated by a small arroyo at this location and a steep seacliff limits direct coastal access to these residential loci. The cemetery at the site was excavated by Philip Mills Jones in 1901 and is thought to be positioned somewhere in the middle of the flat, coastal plain (Heizer and Elsasser 1956). A third residential sector (CA-SRI-611) is positioned about 500 m farther into the interior, just on the edge of the upper marine terrace that overlooks the coastal plain and the large coastal village that occupies it. This residential compound has visible house depressions and, although it is covered with a thick blanket of grass, an impressive artifact collection that includes triangular microblades (with retouch) and callus-cup beads— artifact types that were clearly produced on the islands between 650 and 200 BP. A radiocarbon date confirms that the debris surrounding the structural depressions at the site dates to the Late Period (520–440 BP; Kennett 1998, 470). The site is strategically placed to have a clear view

182

L AT E H O L O C E N E

out across the mouth of Cañada Verde to the other residential sectors in the site complex, but it also allowed members of the community to monitor activities in the island’s interior. The pairing of primary coastal villages with satellite communities also occurs at several additional locations on Santa Rosa (e.g., CA-SRI-165 with CA-SRI-6; CA-SRI-166 with SRI-77). The strategic placement of primary villages is also evident on Santa Cruz Island during the Late Holocene, although the locations of these settlements appear to be more varied than on the outer islands (Arnold 2001; Munns and Arnold 2002). Some of these communities were located on large open bays (e.g., Prisoners and China harbor), whereas others were positioned on small anchorages (e.g., Scorpion Anchorage). These locations had either relatively clear views along the coast or strategically placed additional residential loci established by the inhabitants of these communities to serve this purpose. An example of the latter occurs at Smugglers Cove, an important area of Late Holocene settlement located in the lee of Santa Cruz Island south of East Point (see map 12). Three Late Holocene sites occur in the Smugglers Cove region, two on the beach near the middle of the cove (separated by the drainage; CA-SCrI-504, -505) and one at Smugglers Point (CA-SCrI-506). This complex of sites is the most probable location of Nanawani, the historic village name associated with the Smugglers region (chapter 5; Johnson 1982, 1993; Kennett et al. 2000), but settlement at this location extends back to at least 850 years ago (see tables 15 and 17). The largest residential sector of this community was established within the sheltered confines of the harbor (CA-SCrI-504, -505), but views along the coast were limited from this position. CA-SCrI-506 appears to be another residential compound associated with this community, but it is located on a headland near the southern periphery of the harbor in a location that is more strategically placed to monitor activities along the coast. A similar relationship may have existed between the villages at Scorpion Anchorage (CA-SCrI-423, -507), and a small Late Holocene settlement positioned on Little Scorpion (CA-SCrI-615) that has much clearer views along the coast (Arnold 1990b; Kennett 1998; Kennett et al. 2000; Perry 2003). The settlement data from the northern Channel Islands suggest that villages were strategically positioned to control vital island resources, particularly fresh drinking water, and that islanders developed varied and innovative ways of monitoring the region surrounding these primary villages.

L AT E H O L O C E N E

183

The distribution of cemeteries on Santa Rosa Island provides additional data to suggest that territorial boundaries on the islands became more rigid through the Late Holocene. Cemeteries are often used by descendants to show cultural affinity to the land that they occupy (Buikstra 1981). This is clearly evident with the prominent display of megalithic tombs across many parts of the European landscape, highly visible features that have been interpreted as territorial markers (Renfrew 1973, 1976; Sjögren 1986). Late Holocene cemeteries on Santa Rosa were positioned in a variety of coastal and interior settings and some may have served as territorial markers (Heizer and Elsasser 1956; King 1990; Orr 1968; see Kennett 1998; 549–552). Similar to the Middle Holocene, most cemeteries on Santa Rosa dating to the Late Holocene were associated with primary villages on the coast, a pattern that makes intuitive sense given reduced settlement mobility evident in the archaeological record and increases in the rigidity of territories centered on these communities as inferred from the strategic positioning of coastal villages. Unlike the Middle Holocene, however, a new tradition of placing cemeteries at strategic interior locations on Santa Rosa is also evident in the record. Most of the known cave complexes in the interior contain burials dating to the Late Holocene (e.g., CA-SRI-147; Orr 1968). Other cemeteries were established on hilltops, the same locations used as residential bases during the Middle Holocene (Heizer and Elsasser 1956; Kennett 1998, 549–552). Burials were carefully placed in the residential middens of these old communities and represent a conscious attempt to show continuity with previous island occupants. This interior burial tradition first started to emerge after 3,000 BP and was well established after 650 BP (Kennett 1998). The sacred reasons for why certain people were buried in the interior are lost, but the hallowed ground represented by these cemeteries probably served to mark the smaller, more rigid territories that appear to have developed on the islands by the Late Period (map 13). There is also ample evidence for violent interaction on the northern Channel Islands throughout the past 8,000 years, with the highest incidences of lethal and sublethal violence developing during the Late Holocene (Lambert 1994). Patricia Lambert and Phillip Walker have examined skeletal collections from island contexts dating to different periods of time (Lambert 1994; Lambert and Walker 1991; Walker 1989). Skeletal evidence for violence includes forearm parry fractures, nasal fractures, healed cranial depressions, and projectile point injuries. The best evidence for temporal trends in island violence comes from

Lambert (1994) Island Violence (%)

Marine Productivity Sea-Surface Temperature (°C) 9

10

11

12

13

14

15

Oxygen (pachyderma-bulloides) -0.4

-0.2

0

0.2

0.4

0.6

0.8

1

1.2

Precipitation (in.)

Water Temperature (°C) 10

12

14

16

18

20

22

0.0 0.2

California Droughts (Stine 1994)

A

B

15

16

17

18

19

20

C

0

21

10

20

30

40

1950

L

0.4 0.6 0.8

LM

1000

1.2

BC-AD

AGE (ka)

1.0

1.4 1.6 1.8

EM

0

2.0 2.2 2.4 2.6 2.8 3.0

(Kennett and Kennett 2000) δ18O (G. bulloides )

(Kennett and Kennett 2000) δ18O (M. californianus )

D

LE

E 1000

Projectile Wounds Cranial Injuries

L AT E H O L O C E N E

185

the frequency of cranial vault fractures and projectile point injuries (figure 27). Many of the cranial vault fractures show evidence of healing and suggest a form of sublethal violence akin to Yanomamo Club fighting (Chagnon 1997; Lambert 1994; Walker 1989). These fractures are generally small and circular (ellipsoidal) in shape, suggesting a blow with a relatively slender, but blunt, object. Cranial injuries were common during all phases of island prehistory, but were most frequent between ~2,500 and 1,300 BP. Temporal shifts in lethal violence are indicated by projectile points found embedded in skeletal material or wounds clearly made by projectiles (e.g., holes; Lambert 1994). Spear and dart point injuries occur throughout the Holocene and arrow wounds increased significantly after the introduction of the bow and arrow (~1,500 BP). Projectile wounds increased slightly between 2,500 and 1,300 BP but were most frequent on the islands between ~1,300 and 600 BP. The number of projectile wounds decreased significantly after 600 BP, which suggests decreases in interpersonal violence after that time. Osteological data indicate that increases in sublethal violence between ~2,500 and 1,300 BP co-occurred with decreases in the overall health and stature of island populations (Lambert 1994, 1997; Lambert and Walker 1991; Walker 1986, 1989). The exponential increases in lethal violence between 1,300 and 600 BP also correlate with osteological evidence for subsistence-related stress, and similar patterns are evident along the Southern California mainland (Lambert 1994; Raab and Larson 1997). The despotic behavior that these data represent occurred during a time of great climatic and social instability (figure 27; Kennett

figure 27. Late Holocene climate records for the Santa Barbara Channel region: (A) Sea-surface-temperature record based on oxygen isotopic analysis of G. bulloides (Kennett and Kennett 2000); (B) Upwelling record based on differing oxygen isotopic compositions of G. bulloides (surface-dwelling foraminifera) and N. pachyderma (deeper-dwelling foraminifera). This is thought to be a proxy of marine productivity (see Kennett and Kennett 2000). (C) Median, maximum, and minimum oxygen isotopic values from California mussel shells collected from radiocarbon-dated archaeological sites on Santa Rosa Island. (Original data in Kennett 1998.) (D) Precipitation record based on tree-ring records from the Southern California Bight (Larson and Michaelson 1989). Stine’s record for Late Holocene droughts is based on submerged tree stumps (Stine 1994). (E) Quantitative evidence for lethal and sublethal violence on the northern Channel Islands (Lambert 1994). (Drafted by D. Kennett.)

186

L AT E H O L O C E N E

and Kennett 2000). Although the waters surrounding the islands were generally cool and productive, dry conditions predominated across Southern California for much of this interval (1,500–600 BP; see bristlecone pine record in figure 12). Climatic conditions also appear to have been some of the most unstable in the Holocene (figure 11), and included at least two severe droughts; one centered on ~1,000 BP and the other on ~700 BP (Kennett and Kennett 2000; Stine 1994). Periodically, dry conditions throughout this interval likely contributed to the solidification of the evenly distributed pattern of settlement around the islands, each community strategically placed in defensible locations and in close proximity to perennial water sources of varying qualities. Severe drought conditions between 800 and 650 BP may also be related to the abandonment of marginally watered settlement locations around the northern Channel Islands evident in the settlement record (Arnold 2001; Kennett and Kennett 2000; Munns and Arnold 2002). Increases in lethal violence between 1,300 and 600 BP correlate more broadly with extended periods of dry weather across Southern California and a host of behavioral responses to these climatic conditions, including migration, technological innovation, and sociopolitical development (Jones et al. 1999; Raab and Larson 1997; Yatsko 2000). The frequency of radiocarbon dates on the mainland coast indicate rapid increases in population (Glassow 1996), and linguistic evidence suggests that coastal and interior Chumash language groups diverged at around this same time (Johnson 2000). The similarities of coastal and interior Chumash dialects suggest fissioning of coastal mainland communities and the establishment of interior satellite villages, a trend that was possibly stimulated by subsistence stress throughout Southern California (Johnson 2000). Island and coastal mainland languages were not mutually intelligible and diverged as early as 3,000 BP (Johnson 2000, 313). This suggests that mainland populations did not expand and replace people living on the northern Channel Islands between 1,300 and 600 BP, but that increases in lethal violence during this time may be related to attempts by mainlanders to do so. Adding to the unstable environmental and social conditions in Southern California through this interval, the bow and arrow was introduced to the area sometime after ~1,500 BP. This new technology spread rapidly across North America between 1,800 and 1,350 BP due to intergroup contact, conflict, and competition (Blitz 1988). Its introduction into California is suggested by the replacement of large atlatl dart points with small arrow points and the simultaneous appearance of

L AT E H O L O C E N E

187

arrow-shaft straighteners (Moratto 1984). In the Santa Barbara Channel region, the presence of the bow and arrow is marked by the appearance of small leaf-shaped projectile points between 1,500 and 1,150 BP (Glassow 1996; see figure 24o–r). The introduction of this technology parallels exponential increases in violence regionally. Many of the initial projectile point wounds between 1,500 and 1,150 BP resulted from atlatl darts or spear points (Lambert 1994), an expected pattern when new technology is introduced into a region. The bow and arrow was in full use by 1,150 BP and is evident in the record by small leaf-shaped and concave-base projectile points (figure 25c–f ) . Mission records indicate a high degree of matrilocality in island Chumash society, and its development may be linked to increases in warfare between 1,300 and 600 BP (Johnson 2000, 312). Matrilocality concentrates related women in villages, while splitting up blood-related males (Divale 1984). Patrilocality is often found in societies in which feuding is common between communities in close geographic proximity (e.g., villages on the same island), because related males band together for protection (Divale 1974; Ember and Ember 1971; Otterbein 1968). Matrilocality is favored in societies engaged in “external” wars because the alliances formed between multiple communities help create an integrated force against a common enemy. At the local level, matrilocal residence patterns favor the development of institutions for socializing unrelated males (e.g., men’s houses; Divale 1984, 20–26), a pattern clearly evident in greater Chumash society by the historic period (Johnson 2000). Linguistic distinctions between island and coastal peoples suggest two distinct populations, and the matrilocal residence pattern evident on the islands may be a product of a persistent threat from mainland communities that started during the early stages of the Late Holocene (~3,000 BP) and intensified between 1,300 and 650 BP. Matrilocality on the islands may also have favored greater political integration between island villages and the periodic development of small island chiefdoms (politically integrated villages heavily influenced by a small number of hereditary lineages) by the Late Period (after 650 BP; see emergence of sociopolitical complexity below).

Economic Intensification Foraging strategies intensified through the Late Holocene as people filled the coastal landscape on the northern Channel Islands, more rigid

188

L AT E H O L O C E N E

territories formed, and despotic behavior became more frequent. The intensive collection of shellfish evident in Middle Holocene records on the islands continued into the Late Holocene, but the species available changed through time. Changes evident in island mollusk assemblages are sometimes difficult to interpret, because they are a complex product of changing environmental conditions (e.g., sea-surface temperature, marine productivity; see figure 11), human selection, and long-term impacts to shellfish populations resulting from human predation. The expansion of primary villages along the coasts of the islands after 3,000 BP reduced the number of refugia for the larger, more sought after, mollusk species (e.g., abalone). A situation created by a shift from logistical to residential exploitation and the expansion of more intensive (local) harvesting practices. The availability of more resilient species [California mussels (M. californianus)] declined and rebounded through the Holocene (Erlandson et al. 2004). Beyond this there appear to be several general trends from the Middle to Late Holocene evident in the record. These include: (1) decreases in the largest species of shellfish (red abalone, H. rufescens), (2) size reduction in the larger and most prolific species (California mussel, M. californianus; black abalone, H. cracherodii), and (3) an overall expansion in the number of species consistently taken (Septifer bifurcates, and others). These changes are evident at CA-SCrI-333 (western Santa Cruz), where Middle Holocene deposits are composed of large red abalone shells and the Late Holocene deposits contain shells of the smaller black abalone (Wilcoxon 1993). This pattern is also evident on Santa Rosa (CA-SRI-147; Kennett 1998) and San Miguel (CA-SMI-528; Walker et al. 2000; see figure 20B). The effects of intensive exploitation on intertidal habitats from the Middle to Late Holocene are also evident at Daisy Cave (CA-SMI-261) and Cave of the Chimneys (CA-SMI-603), two sites positioned at Bay Point on San Miguel Island (Erlandson et al. 1996a; Vellanoweth et al. 2000). The shellfish assemblages from four strata dating between 4,300 and 2,500 BP [CA-SMI-603, 4,300 BP (Stratum 4); 4,000 BP (Stratum 3); 2,500 BP (Stratum 2); CA-SMI-261, 3,400 BP (Stratum A)] have been analyzed in great detail and show changes in the availability of shellfish in the intertidal zone within the vicinity of Bay Point (Vellanoweth et al. 2000). Red abalone, the largest species in the assemblage, provided the highest proportion of edible meat in the lower two strata at Cave of the Chimneys (Stratum 4, 2401.5 g; Stratum 3, 2222.6 g), but the contribution of this species is less significant in the more recent strata at the site (Stratum 2, 414.1 g). Similar trends are evident

L AT E H O L O C E N E

189

with black abalone and sea urchin (Vellanoweth et al. 2000). As discussed previously (see chapter 6), marine conditions (either cooler seasurface temperatures or the lower frequency of El Niños) may have favored red abalone in the intertidal zone during parts of the Middle Holocene. If environmental conditions influenced the composition of these mollusk assemblages, the age of the deposits at Cave of the Chimneys suggests that a lower frequency of El Niño was responsible (or people were diving; Sharp 2000), because marine conditions were relatively warm at this time (see figure 11). The reduction in both red abalone (cool-water species) and black abalone (warm-water species) suggest that predation pressure also played a role in the reduction of these larger species. These data are consistent with evidence for predation pressure effectively reducing the size and abundance of black abalone on the southern Channel Islands through time (Raab 1992). The general reduction in the larger, more prolific, species at Bay Point from the Middle to Late Holocene also parallels a general decrease in the meat contribution of shellfish and an overall increase in the contribution of fish (Vellanoweth et al. 2000). Increases in the meat contribution of fish, relative to shellfish, visible in the Late Holocene records at Daisy Cave and Cave of the Chimneys, provide a bellwether for a general increase in fishing and the dietary importance of fish during the Late Holocene on the northern Channel Islands, an economic pattern that intensified through the interval (Kennett and Kennett 2000). This economic shift started during a time when sea-surface temperatures were cool (but variable) and marine productivity was relatively low compared to the preceding thousand years (see figure 11). Compared with the Middle Holocene, terrestrial conditions were dry during the Late Holocene, with the exception of the past 500 years and a 500-year interval between 2,100 and 1,600 BP. During the Middle Holocene, the available data suggest that shellfish were the main source of meat for many people living on the islands (Glassow 1993a, 1993b; Kennett 1998; Vellanoweth et al. 2000; see chapter 6). Midden constituent data indicate that shellfish remained a prominent meat source during the Late Holocene (table 18), but its importance relative to fish decreased through time. These temporal data suggest a gradual increase in the importance of fishing between 3,000 and 1,300 BP. In fact, shellfish continued to be the primary meat source at some locations during this time. An exponential increase in fish-bone density occurs in some middens dating to between 1,300 and 800 BP and suggests intensified fishing at some locations during this

table 18. Faunal data (fish and shellfish) from deposits dating to the Late Holocene on the northern Channel Islands Fish Site No.

Provenience

CA-SRI-31 CA-SMI-525

Unit 2, 10–25 cm Prof. D, Strat. 27, 227–235 cm Prof. N, Strat. 12, 200–230 cm Unit 1, 10–20 cm Profile N, Strat. 4, 45–60 cm Profile N, Strat. 8, 30–52 cm Unit 1, 40–50 cm Unit 1, 20–30 cm Unit 1, 20–30 cm Unit 1, 0–10 cm Unit 1, 0–10 cm Unit 1, 28–40 cm Prof. N, Strat. 9, 48–64 cm Unit 1, 12–20 cm Unit 1, 7–20 cm Unit 1, 40–44 cm Prof. D, Strat. 9, 70–79 cm Unit 1, 0–10 cm Prof. N, Strat. 6, 89–97 cm

CA-SMI-504 CA-SRI-19 CA-SMI-488 CA-SMI-503 CA-SRI-62 CA-SRI-41 CA-SRI-62 CA-SRI-96 CA-SRI-432 CA-SRI-31 CA-SMI-492 CA-SRI-31 CA-SRI-130 CA-SRI-130 CA-SMI-525 CA-SRI-41 CA-SMI-510

Age

(1 sigma)

Weight

Meat

3,051 3,013

3,193–2,911 3,139–2,906

57.1 47

1582 1301.9

2,933

3,036–2,846

0

2,862 2,639

2,955–2,781 2,705–2,457

2,493

2,680–2,370

2,479 2,144 2,056 1,769 1,603 1,340 1,292

6.23 37

0

Shellfish % 77.82 86.59 0

Weight

Meat

1358 607.3

450.9 201.62

55.3

18.36

%

References

22.18 13.41

Kennett 1998 Walker and Snethkamp 1984 Walker and Snethkamp 1984 Kennett 1998 Walker and Snethkamp 1984 Walker and Snethkamp 1984 Kennett 1998 Kennett 1998 Kennett 1998 Kennett 1998 Kennett 1998 Kennett 1998 Walker and Snethkamp 1984 Kennett 1998 Kennett 1998 Kennett 1998 Walker and Snethkamp 1984 Kennett 1998 Walker and Snethkamp 1984

100

172.57 1024.9

54.82 93.57

428.41 212

142.23 70.38

45.18 6.426

125

3462.25

96.65

361.7

120.08

3.352

2,670–2,357 2,284–2,072 2,144–1,956 1,836–1,694 1,710–1,511 1,414–1,286 1,361–1,244

8.2 69.46 7.5 80.9 13.2 65.96 300

227.14 1924.04 181.5 2240.93 365.64 1827.09 8310

43.88 81.35 41.21 90.78 62.06 74.99 94.89

875.1 1329.3 779.99 685.94 673.23 1835 1349.3

290.53 441.23 258.96 227.73 223.51 609.22 447.97

56.12 18.65 58.79 9.225 37.94 25.01 5.115

1,264 1,264 1,225 1,225

1,313–1,183 1,313–1,183 1,276–1,146 1,280–1,137

27.72 11.6 6.25 196

755.92 321.32 173.12 5429.2

69.55 36.59 76.7 97.68

996.98 1677.6 158.45 389.8

331 556.96 52.6 129.11

30.45 63.41 23.3 2.323

1,197 1,182

1,260–1,134 1,259–1,107

18.79 38

520.48 1052.6

60.65 66.49

1017.2 1598

337.71 530.54

39.35 33.51

CA-SRI-15 CA-SRI-15 CA-SRI-15 CA-SRI-15 CA-SRI-15 CA-SRI-97 CA-SRI-97 CA-SRI-15 CA-SRI-85 CA-SRI-15 CA-SMI-525 CA-SMI-485

Unit 1, 126–135 cm Unit 1, 117–126 cm Unit 1, 80–95 cm Unit 1, 65–75 cm Unit 1, 19–30 cm Unit 1, 60–70 cm Unit 1, 40–50 cm Unit 1, 0–10 cm Unit 1, 70–80 cm Unit 2, 40–50 cm Prof. D, Strat. 3, 30–37 cm Prof. S, 10–20 cm

CA-SRI-15 CA-SMI-602 CA-SMI-602 CA-SRI-85 CA-SMI-602 CA-SRI-60 CA-SRI-40 CA-SRI-97 CA-SMI-602

Unit 2, 20–29 cm Unit 5, 40–50 cm Unit 5, 0–10 cm Unit 1, 0–10 cm Unit 2, Strat. B Unit 1, 40–50 cm Unit 1, 60–70 cm Unit 1, 10–26 cm Unit 2, Strat. A

1,067 984 984 984 733 706 724 678 622 622 544

1,178–958 1,075–922 1,075–922 1,068–927 820–671 760–663 797–666 764–633 656–543 656–543 616–510

21.23 74.4 124 165.1 80.7 78.4 88.6 75.86 115.92 81.8 31

588.07 2061.4 3443.5 4573.27 2234.7 2172.3 2454.2 2101.32 3210.98 2265 858.7

87.84 94.89 90.19 89.66 94.33 90.1 92.15 81.45 88.55 99.29 66.49

245.28 333.97 1127.6 1588.6 404.51 718.9 629.8 1414.2 1250.9 509 1303.3

81.43 110.9 374.4 527.41 134.3 238.7 209.1 478.47 415.3 16.29 432.69

12.16 5.105 9.806 10.34 5.669 9.9 7.851 18.55 11.45 0.714 33.51

516

574–469

103

2853.1

93.2

626.6

208.03

6.796

511 491 491 468 294 264 264 243 115

550–474 527–447 532–440 507–417 395–260 299–154 299–154 273–102 243–0

79.2 2194 137.29 3802.93 57.87 1603 40.84 1131.27 16.73 463.42 35.37 979.75 347.81 9634.34 414.4 11478 43.82 1213.81

682.9 21.85 667.18 221.5 164.12 54.49 551.2 183 163.2 54.18 1658.2 550.53 3099.4 1029 626.18 207.89 183.97 61.08

0.986 5.504 3.288 13.92 10.47 35.98 9.65 1.779 4.791

99.01 94.5 96.71 86.08 89.53 64.02 90.35 98.22 95.21

Kennett 1998 Kennett 1998 Kennett 1998 Kennett 1998 Kennett 1998 Kennett 1998 Kennett 1998 Kennett 1998 Kennett 1998 Kennett 1998 Walker and Snethkamp 1984 Walker and Snethkamp 1984 Kennett 1998 Walker et al. 2000 Walker et al. 2000 Kennett 1998 Walker et al. 2000 Kennett 1998 Kennett 1998 Kennett 1998 Walker et al. 2000

192

L AT E H O L O C E N E

interval (Kennett 1998; Kennett and Conlee 2002; Kennett and Kennett 2000). Large quantities of fish bone are present in middens throughout the northern Channel Islands by 650 BP (Colten 1993, 1994, 1995, 2001; Kennett and Conlee 2002; Kennett and Kennett 2000), and these data are consistent with ethnohistorical accounts of the importance of fish and fishing in island Chumash society (see chapter 4). Midden constituent data are also consistent with stable nitrogen and carbon isotopic measurements on skeletal material from island burial contexts. Relative to the Middle Holocene, the nitrogen and carbon isotopic measurements from the Late Holocene are enriched (more positive), indicating a greater dependence upon foods from marine sources (Goldberg 1993; Walker and DeNiro 1986). The sex differences in diet evident during the Middle Holocene also disappear during the Late Holocene (Goldberg 1993). The types of fish targeted on the northern Channel Islands varied across space and through time but generally provide additional evidence for economic intensification during the Late Holocene. Intersite variability in island fish assemblages is a product of local ecological conditions and variations in sea-surface temperature and marine productivity across space. Fish populations tend to be resilient to increasing predation pressure and are inherently intensifiable for this reason. The local effects of varied marine conditions around the northern Channel Islands is demonstrated by the higher proportion of warm-water fish species in assemblages on northern Santa Cruz (CA-SCrI-240 and -306), as opposed to the higher frequency of cooler-water species in sites on western Santa Cruz (CA-SCrI-191, -474; Pletka 2001b, 242), a pattern consistent with the modern distribution of fish around the islands (Engle 1993, 1994; see chapter 3). Similar to Early and Middle Holocene fish assemblages on the outer islands (Erlandson et al. 1999; Rick et al. 2001a), a high proportion of species in Late Holocene deposits are from nearshore kelp bed and rocky coast habitats. Fish from beyond nearshore habitats (midwater, deep ocean, open ocean) required the development of more sophisticated fishing technologies and appear to increase in sites on Santa Cruz Island dating to after 800 BP (Pletka 2001b; between the Middle and Late Periods), with people living at CA-SCrI-191 exploiting mid- and deep-water habitats most frequently between 800 and 650 BP (Pletka 2001b; Transitional Period). Increases in the types of fish taken suggest an expansion of diet breadth through the Late Holocene, a process that had already started during the later stages of the Middle Holocene.

L AT E H O L O C E N E

193

A number of new fishing technologies (e.g., single-piece fishhooks of different types and sizes, toggling harpoons) developed during the Late Holocene and expanded the fishing capabilities of islanders to new habitats and target species (Rick 2001; Rick et al. 2002). The development of the plank canoe between 2,000 and 1,500 BP also enhanced the ability of people living in the Santa Barbara Channel region to fish offshore for swordfish and marlin (Bernard 2001; Davenport et al. 1993; Gamble 2002), and the bones of these large fish in island midden deposits dating to after 650 BP indicates the use of these boats by at least this time (Pletka 2001b). Net fishing appears to have been less important on the islands when compared to the mainland, at least during the Late and Historic Periods (Pletka 2001b; but see Bowser 1993). The development of single-piece shell fishhooks was the primary technological change associated with the shift toward more intensive use of the fishery on the northern Channel Islands. The earliest directly radiocarbon-dated fishhooks on these islands date to between 2,500 and 2,100 BP (Rick et al. 2002; table 19; figure 28). These hooks were manufactured from abalone or mussel shell, and the first examples are relatively simple in form (j-shaped). This form persisted on the islands for several thousand years with only small changes in style (e.g., notching). Fishhooks appear to be more common in deposits dating to after 1,500 BP, and the greatest diversity in size and form occurs between 1,500 and 1,100 BP (see figure 28). Fishhooks became more standardized on the islands after 650 years ago (circular form with a grooved shank). The diversity evident in the fishhook form between 1,500 and 1,100 BP suggests that a range of species were targeted, a pattern consistent with overall increases in fish-bone density and the expansion of species diversity in northern Channel Island middens during this time. Island settlement data suggest that the use of interior residences became more restricted during the Late Holocene as people became more tethered to coastal locations. These data seem to suggest that plant foods were de-emphasized as maritime foraging strategies intensified. The macrobotanical data needed to evaluate changes in plant use from the Middle to Late Holocene are not currently available. However, a study of floral remains from several Late Holocene sites on Santa Cruz Island provides some insight into the economic use of plants after 1,500 BP (Martin and Popper 2001). Over 1,500 seeds were preserved in these deposits, representing 20 genera and 18 families. Most of these species are found today on the islands and suggest consistent use of a range of plants through the latest Holocene, including nuts (acorns),

table 19. Direct AMS radiocarbon dates on shell fishhooks from the northern Channel Islands Cat No.

Provenience

Lab No.

Material

300-902 301-500 301-40 301-41 301-39 301-18 300-160 301-501 300-901 301-502

CA-SRI-131, Unit 3, 0–10 cm CA-SMI-481, below biggest dune CA-SMI-525, Unit1D, Strat3, 110–120 cm CA-SMI-525, Unit1D, Strat3, 140–150 cm CA-SMI-525, Unit1D, Strat3, 160–165 cm CA-SMI-525, Unit 1c, 210–220 cm CA-SRI-62, Unit 1, 40–50 cm CA-SMI-87, East Dune, Surface CA-SRI-1, Surface CA-SMI-152, Surface

OS-38219 OS-38218 OS-34571 OS-34572 OS-34570 OS-30000 OS-34569 OS-26071 OS-38217 OS-37145

Haliotis sp. H. rufescens Mytilus californianus M. californianus M. californianus Haliotis sp. H. rufescens Haliotis sp. Haliotis sp. Haliotis sp.

note: 1 sigma range with mean intercept in parentheses. AMS 14C measurements were completed at the National Oceanic Sciences AMS facility (NOSAMS). All dates were calibrated using Calib 4.3 (Stuiver and Reimer 1993) using a reservoir age of 225  35.

14

C Age

1,880  55 2,150  50 2,380  30 2,450  30 2,590  30 2,730  35 2,970  30 2,980  35 3,020  45 3,060  80

1 Sigma (BP) 1273(1226)1158 1533(1483)1396 1808(1733)1690 1873(1822)1776 2038(1981)1924 2260(2150)2100 2499(2437)2348 2540(2450)2350 2676(2506)2424 2720(2650)2440

figure 28. Stylistic changes in shell fishhooks on the northern Channel Islands through time. A selection of these fishhooks have been directly AMS radiocarbon dated (see table 19), and the remainder come from well-dated contexts at CA-SMI-528 (1,500–1,100 BP) and CA-SMI-602 (550–250 BP). (Illustrations and layout by R. van Rossman.)

196

L AT E H O L O C E N E

fruit (prickley pear, manzanita, and lemonade berry), seeds (sage, barley, red maids, among others), and greens (e.g., goosefoot). Several plant foods also appear to have been imported from the mainland (e.g., black walnut, sea purslane, canary grass). This is consistent with a number of ethnohistorical accounts suggesting that islanders traded shell beads with people living in mainland communities in exchange for plant foods (e.g., acorns; Hudson and Blackburn 1982; King 1976). The exchange of nonfood items (beads, otter pelts, etc.) for plant foods was yet another way that people living in island villages could increase diet breadth, and these exchange relations would have been greatly facilitated by the development of the plank canoe between 2,000 and 1,500 BP (Davenport et al. 1993; Gamble 2002) and the greater use of these watercraft thereafter (Arnold 1995). Archaeological data indicate that six marine mammal species were hunted on the northern Channel Islands: northern elephant seal (Mirounga angustirostris), California sea lions (Zalophus californianus), southern fur seals (Arctocephalus townsendi), northern fur seals (Callorhinus ursinus), harbor seals (Phoca vitulina), and sea otter (Enhydra lutris). Cetaceans (e.g., dolphins) were also taken during the Middle Holocene (Glassow 2000), and hunting of other pelagic animals (e.g., swordfish) appears to increase after the development of the plank canoe in the Late Holocene (between 2,000 and 1,500 BP; Davenport et al. 1993). Whale bones are commonly found in a variety of archeological contexts on the islands (e.g., middens, houses, and burials), but they are attributed to scavenging of stranded animals, rather than hunting, because these large animals were not pursued during the historic period (Glassow 1980). Predation of pinniped populations in coastal rookeries (sea lions, fur seals, etc.) has received the greatest attention because these large animals need to come ashore to breed and, if present, would have been an easy source of prey for Native American populations living along the west coast of North America. Hildebrandt and Jones (1992) originally proposed that pinnipeds were preferentially targeted, which forced breeding populations offshore to islands and rocks (Hildebrandt and Jones 1992; Jones and Hildebrandt 1995; Lyman 1995; see chapter 2). They also argued that the overall reduction of the population of larger, more susceptible animals (e.g., sea lions and fur seals) stimulated the pursuit of smaller, more elusive animals (e.g., sea otters and harbor seals). This, in turn, prompted the development of sophisticated watercraft (e.g., oceangoing plank canoes) and the weaponry (composite

L AT E H O L O C E N E

197

harpoons) necessary to pursue these animals in offshore rookeries (Jones and Hildebrandt 1995). The northern Channel Islands support some of the largest breeding populations of pinnipeds in the eastern Pacific (Bartholomew 1967; Le Boeuf and Bonnel 1980) and are well positioned to provide information on the changing role of sea mammal hunting in coastal California. Harbor seals and sea otters are present around all of the islands, but the largest aggregations of pinnipeds (e.g., sea lions, fur seals, elephant seals) occur on the western end of San Miguel Island at Point Bennett (Walker et al. 2000). The importance of marine mammal hunting prehistorically varied spatially, but the evidence overwhelmingly indicates that marine mammals played a minor role in the diets of islanders during most of the Holocene (Colten and Arnold 1998, 2000). Shellfish (e.g., abalone and California mussels) provided the bulk of edible meat in the island diet during the Middle Holocene (Glassow 1993a, 2000; Kennett 1998; Vellanoweth et al. 2000), and shellfish and fish were the primary meat sources at most locations during the Late Holocene (Colten and Arnold 1998, 2000). This is not surprising on Santa Cruz and Santa Rosa, because rookeries were probably never present on these islands. The lack of evidence for marine mammal hunting on San Miguel Island is surprising because of the large aggregations of these animals today, at least near Point Bennett. Excavations at two sites near Point Bennett (CA-SMI-528, -602; map 12) provide some insight into the exploitation of pinnipeds at this location (Walker et al. 2000). CA-SMI-528 is a long, narrow shell midden that runs along the spine of a high dune on the eastern periphery of the modern-day sea mammal rookery. The dune is capped with a thick, dark brown, midden soil (Stratum I; ~70 cm thick) that dates to between 1,500 and 1,200 BP (late Middle Period), and two additional strata in the dune deposits below, one dating to ~4,800 BP (Stratum II) and the other to 5,850 BP (Stratum III, see figure 20). The other site (CA-SMI-602) is a shallow, stratified midden located in the middle of the modern-day pinniped breeding colony, 250 m north of Adams Cove. Multiple houses are evident on the surface of this site and scattered human remains suggest the presence of a cemetery somewhere in the vicinity. Radiocarbon dates and diagnostic artifacts indicate that this village was occupied during the Late Period (~650–200 BP). Small amounts of marine mammal bone are present in the Middle Holocene deposits at CA-SMI-528, but both strata are dominated by abalone and California mussel shells. In comparison to Late Holocene sites

198

L AT E H O L O C E N E

elsewhere on the northern Channel Islands, the late Middle and Late Period deposits at CA-SMI-528 and -602 contained large quantities of marine mammal bone, a pattern that makes sense given the geographic proximity of these sites to the modern-day rookery. At CA-SMI-528, an excavation unit (1  2 m) in Stratum I (1,500–1,200 BP) contained higher densities of marine mammal bone (1,042.1 kg meat/m3) compared with fish (233.6 kg meat/m3) and shellfish (146.6 kg meat/m3). CA-SMI-602 (650–200 BP) contained extremely high concentrations of fish bone (836.5 kg meat/m3), compared with marine mammal (262.7 kg meat/m3), shellfish (23.5 kg meat/m3), and bird (5.8 kg meat/m3). These data suggest that marine mammal hunting was not an important pursuit until a more permanent village was established at this location at ~1,500 BP. The relative absence of marine mammal bone in the Middle Holocene deposits at CA-SMI-528 suggest that rookeries may have been restricted to offshore rocks and were too costly to hunt given the available watercraft and hunting technology. High transportation costs back to primary village sites, possibly on Santa Rosa (see chapter 6), may have also contributed to the lack of interest in marine mammal hunting during the Middle Holocene. The absence of animals in San Miguel Island middens during much of the Holocene is baffling and could have resulted from early and persistent disruption of breeding populations on the island. Increased predation after 1,500 BP likely resulted from the pursuit of these animals to offshore rocks using newly available technologies (e.g., plank canoe and harpoon) and is partially supportive of the Hildebrandt and Jones (1992) predation pressure model.

Increases in Trade and Exchange The presence of tools or ornaments in island contexts that were manufactured from materials only found on the mainland (e.g., obsidian, chert, serpentine) show the persistence of trade and exchange across the Santa Barbara Channel during much of the Holocene (King 1990; Rick et al. 2001b; Orr 1968; Pletka 2001a). These interactions played a prominent role in the evolutionary history of these islanders and in Chumash society more generally (Arnold 1992a; King 1990). The diversity of formal artifacts increased markedly at primary village sites on the islands during the Late Holocene, particularly after 1,300 years ago. This pattern suggests further decreases in settlement mobility but is

L AT E H O L O C E N E

199

also indicative of more vigorous regional trade and exchange after this time. By the historic period, a wide range of items were traded between the islands and the mainland. Islanders provided fish, beads, otter pelts, and other craft items to mainlanders, and mainlanders furnished a variety of food items (e.g., acorns and seeds) and baskets (see chapter 4; King 1976, 1990). Olivella-shell-bead production, one of several important trade items manufactured by islanders, increased between 1,300 and 650 BP and has been linked to intensified trade after this time (Arnold and Munns 1994; Kennett 1998; Kennett and Conlee 2002; King 1990), and elite members of island Chumash society likely controlled some aspects of this production and distribution system (Arnold 1992a, 2001; see section below). A hallmark of trade and exchange across the Santa Barbara Channel throughout the Holocene is the low-level presence of obsidian in middens and burials on the northern Channel Islands (King 1990; Rick et al. 2001b). Obsidian, a volcanic glass used to manufacture stone tools, does not occur naturally in the Santa Barbara Channel region and is only found in several discrete interior locations in California, Nevada, and Oregon. The earliest securely dated obsidian sample on the islands dates to between 7,400 and 5,580 BP and was found at CA-SRI-147, a site located in the interior of Santa Rosa Island (see chapter 6; Rick et al. 2001b). Obsidian has also been recovered from a variety of other Middle Holocene island sites (CA-SRI-3, -4; CA-SMI-1, - 172) and from sites of Late Holocene age (CA-SCrI-191, -236, -240, -474; CA-SRI-2, -9, -19, -60; CA-SMI-528). Geochemical analysis of this material indicates that a majority comes from one of three sources in the Coso volcanic field region, located near the southern end of the Owens Valley in central California, the closest high-quality material available to island populations (Rick et al. 2001b). Trace amounts of obsidian from other more distant sources (e.g., Mt. Hicks, Casa Diablo, and Massacre Lake) were also present in the sample and all come from Late Holocene contexts. The presence of small amounts of obsidian throughout the Holocene suggests that islanders were linked with people in central and Southern California through a broad network of exchange that possibly became more extensive during the Late Holocene. Other material exotic to the islands also indicates cross-channel contact and intensified trade and exchange through the Late Holocene. As previously mentioned, the remains of plants exotic to the islands have been found in Late Holocene village sites on Santa Cruz and are suggestive of trade and exchange of carbohydrate-rich plant foods

200

L AT E H O L O C E N E

(Martin and Popper 2001). This is consistent with ethnohistorical accounts for the exchange of plant foods across the channel from the mainland, but many of the most important plants mentioned in these accounts were not present in the sample (e.g., acorns). Tools manufactured from deer bone, not available on the islands, were also identified in Late Holocene sites on Santa Cruz and indicate exchange with the adjacent mainland (Wake 2001). Spear and arrow points manufactured from mainland stone sources, other than obsidian (e.g., Monterey and Franciscan cherts, fused shale; Kennett 1998; Pletka 2001a), also suggest the existence of cross-channel exchange relationships. In a detailed study of stone tools from sites on Santa Cruz spanning much of the Late Holocene (2,600–250 BP), a high proportion of the raw material used in manufacture was from the mainland (Pletka 2001a, 133). Medium-quality Monterey chert is available on eastern Santa Cruz and islanders manufactured microblades and small bifaces from this material after 1,150 BP. In fact, most of the small arrow points made on Santa Cruz after 650 BP were manufactured with material from this local source (Pletka 2001a). Many of the spear and arrow points recovered thus far from the outer islands of Santa Rosa and San Miguel were also crafted from materials only available on the mainland, but tools made from these exotic sources appear to persist throughout the Late Holocene (Kennett 1998). A source of chalcedonic chert (Cico) was available on San Miguel Island (Erlandson et al. 1997), but local sources of chert on the outer islands are generally smaller and more restricted than on Santa Cruz and the mainland. The persistent use of bifaces manufactured from mainland sources suggests that these items served to maintain individual social relationships (Pletka 2001a, 148) or that these tools were packaged with other items that were not available on the islands (e.g., wooden spear or arrow shafts, deer sinew for hafting). Increased use of local chert on Santa Cruz after 650 BP may be indicative of significant changes in the production and distribution of trade items within the broader Chumash interaction sphere and the control of these relationships by elite individuals (Arnold 2001; Pletka 2001a). It could also be that island chert was more suitable for the manufacture of small arrow points and/or that the island economy was saturated with arrow shafts and deer sinew by the Late Period. The manufacture and use of groundstone tools on the northern Channel Islands provide additional data suggesting increases in trade and exchange during the Late Holocene. Intensive mortar and pestle

L AT E H O L O C E N E

201

manufacture is well known at a complex of sites on the northwestern coast of San Miguel Island (CA-SMI-503, -504; Conlee 2000; Kennett and Conlee 2002; Rozaire 1983; Walker and Snethkamp 1984). Radiocarbon dates on levels with small amounts of manufacturing debris at these locations suggest that the industry extends back to at least ~2,500 BP. Groundstone quarrying and production activities were centered at CA-SMI-503 and -504, two adjacent sites located directly above an Eocene conglomerate formation that is exposed in the seacliff just to the north (Bremner 1932; Weaver and Doerner 1969). Boulder-sized volcanic porphyries from this formation provided the raw material for mortar and pestle manufacturing. These porphyries grade from gray to red, and chipping waste from groundstone production virtually covers these sites; in addition, mortars and pestles in all stages of production litter their surfaces (Walker and Snethkamp 1984). The upper stratum at CA-SMI-503 dates to between ~1,250 and 1,000 BP and contains the highest densities of manufacturing debris. Four other sites in the vicinity of CA-SMI-503 also have levels with high concentrations of chipping debris that date to this same interval of time. Mortars and pestles in the Santa Barbara Channel region were used primarily to grind acorns (Glassow 1980), and the intense production of groundstone on San Miguel is somewhat paradoxical given the lack of oak trees and the limited availability of other plant foods on the island. Other California groups used groundstone to process rodents, fish, insects, and large mammals, but the most prudent explanation is that these tools were manufactured on San Miguel to process plant foods (Conlee 2000). This means that mortars and pestles were being exported or that acorns were being imported to the islands. Regardless, increased production of groundstone suggests heightened intervillage exchange and may represent an attempt by these islanders to expand their overall dietary breadth. Changes in the use of groundstone on Santa Cruz Island also signal significant changes in subsistence and exchange during the Late Holocene (Delaney-Rivera 2001). The density of groundstone in Santa Cruz Island assemblages remained relatively constant through the Middle Period (2,600–850 BP), but plant-processing tools were much less frequent in the subsequent Late Period (after 650 BP). This corresponds with a slight increase in groundstone density in some sites on the adjacent mainland coast. These data are consistent with the intensified use of marine resources, particularly fish, evident in island faunal assemblages (Colten 2001; Kennett and Kennett 2000; Pletka 2001b)

202

L AT E H O L O C E N E

and osteological data (Lambert and Walker 1991; Walker and DeNiro 1986). However, they may also indicate that processed plant foods (e.g., acorn mush) were exported from the mainland to the islands. Processing edible plants on the mainland would have reduced their bulk (acorn shells) and increased the net delivery rate of carbohydrate-rich plant foods to islanders. Olivella-shell beads were among the most important trade items in the Chumash interaction sphere and, because of their great value, were exchanged throughout western North America (Bennyhoff and Hughes 1987; King 1990). Various bead types were produced on the northern Channel Islands during the Holocene, and stylistic changes are temporally sensitive (King 1990). Shell bead production peaked on the islands between 650 and 200 years ago when cross-channel trade was at its height (Arnold 1987; King 1990). Beads manufactured during this interval were highly standardized and produced from the callus portion of the Olivella shell, a small gastropod found along sandy beaches on the islands. Strands of beads appear to have been a medium of exchange, and most of the beads found throughout the Santa Barbara Channel region were produced on the northern Channel Islands during this time. Many of the primary village sites dating to this time are covered with Olivella-bead-manufacturing detritus (map 14B), but beads were manufactured on all the islands, including Anacapa (Rozaire 1993). Arnold (1987; Arnold and Munns 1994) has documented the most intensive production of callus-cup beads at sites on the west end of Santa Cruz, but large-scale bead manufacturing also occurred on Santa Rosa Island (Kennett 1998; Kennett and Conlee 2002). Concentrations of bead-manufacturing detritus are particularly great at a series of sites near the mouth of Old Ranch Canyon on eastern Santa Rosa (CA-SRI-85, -87, -88, -187), all associated with long, sandy beaches that occur on this part of the island, the perfect habitat for Olivella biplicata. Although the most intensive bead manufacturing clearly occurred between 650 and 200 BP, Olivella-bead manufacturing started to intensify on the islands between 1,300 and 650 BP (Christy 2004; Kennett 1998; Kennett and Conlee 2002; King 1990, 154). Olivellabead-manufacturing density data from radiocarbon-dated column samples on San Miguel, Santa Rosa, and Santa Cruz islands serve as a proxy for the intensity of bead manufacturing on the islands between 3,000 and 200 BP (table 20; Arnold and Munns 1994; Christy 2004; Kennett 1998; Kennett and Conlee 2002; Walker and Snethkamp

map 14. Distribution and intensity of Olivella-shell-bead manufacturing at sites on the northern Channel Islands between (A) 1,300 and 650 BP and (B) 650 and 200 BP. White squares represent locations at which Olivella-bead manufacturing has been identified, and circles represent midden samples that have been analyzed quantitatively (map produced by D. Kennett with the assistance of J. Bartruff).

table 20. Olivella-shell-bead-manufacturing detritus densities in column samples from the northern Channel Islands

Site No. CA-SRI-83 CA-SRI-31 CA-SRI-19 CA-SMI-488 CA-SRI-62 CA-SMI-503 CA-SRI-41 CA-SRI-62 CA-SRI-187 CA-SRI-96 CA-SCrI-474 CA-SRI-6 CA-SRI-6 CA-SCrI-474 CA-SRI-31 CA-SRI-130 CA-SRI-31 CA-SRI-130 CA-SRI-41 CA-SRI-77 CA-SRI-77 CA-SCrI-191 CA-SRI-2 CA-SRI-2 CA-SRI-15

Provenience Unit 2, 40–48 cm Unit 2, 10–25 cm Unit 1, 10–20 cm Profile N, 0–10 cm Unit 1, 40–50 cm Profile N, 0–35 cm Unit 1, 20–30 cm Unit 1, 20–30 cm Unit 1, 10–20 cm Unit 1, 0–10 cm Unit 1S, 11W, 70–75 cm Unit 2, 20–30 cm Unit 2, 0–10 cm Unit 1S, 11W, 55–60 cm Unit 1, 28–40 cm Unit 1, 7–20 cm Unit 1, 12–20 cm Unit 1, 40–44 cm Unit 1, 0–10 cm Unit 1, 40–50 cm Unit 2, 50–60 cm Unit 35S, 3W, 75–80 cm Unit 2, 20–30 cm Seacliff, 50–60 cm Unit 1, 0–10 cm

Age/Period (1 sigma) 3,253 3,320–3,200 3,051 3,193–2,911 2,862 2,955–2,781 2,639 2,705–2,457 2,479 2,670–2,357 2,493 2,680–2,370 2,144 2,284–2,072 2,056 2,144–1,956 1,880 1,930–1,820 1,769 1,836–1,694 M2 1,790–1,290 1,520 1,580–1,480 1,512 1,548–1,456 M3 1,290–970 1,340 1,414–1,286 1,264 1,313–1,183 1,264 1,313–1,168 1,225 1,276–1,146 1,197 1,260–1,134 1,140 1,180–1,070 1,010 1,070–950 M4 970–780 790 880–740 730 780–670 678 764–633

Olivella Detritus Weight (g/cu m) 190 0 194 275 54 375 368 454 1,018 373 572 370 469 1,970 507 148 1,068 417 432 2,973 3,035 778 2,632 2,888 1,402

References Christy 2004 Kennett 1998 Kennett 1998 Walker and Snethkamp 1984 Kennett 1998 Walker and Snethkamp 1984 Kennett 1998 Kennett 1998 Christy 2004 Kennett 1998 Arnold and Munns 1994 Christy 2004 Kennett and Conlee 2002 Arnold and Munns 1994 Kennett 1998 Kennett 1998 Kennett 1998 Kennett 1998 Kennett 1998 Christy 2004 Christy 2004 Arnold and Munns 1994 Christy 2004 Christy 2004 Kennett 1998

CA-SCrI-474 CA-SCrI-474 CA-SCrI-191 CA-SRI-84 CA-SRI-241 CA-SRI-60 CA-SRI-85 CA-SCrI-330 CA-SCrI-191 CA-SCrI-330 CA-SCrI-330 CA-SCrI-192 CA-SCrI-192 CA-SMI-485 CA-SMI-602 CA-SMI-602 CA-SRI-85 CA-SRI-436 CA-SRI-436 CA-SMI-602 CA-SRI-40 CA-SRI-60 CA-SRI-40 CA-SRI-97 CA-SMI-602

Unit 1S, 11W, 10–15 cm Unit 1S, 11W, 35–40 cm Unit 35S, 3W Unit 4, 30–35 cm Unit 1, 30–40 cm Unit 1, 60–70 cm Unit 1, 70–80 cm Unit 3S, 28W, 105–110 cm Unit 35S, 3W, 35–40 cm Unit 3S, 28W, 65–70 cm Unit 3S, 28W, 40–45 cm Unit 2N, 23E, 65–70 cm Unit 2N, 23E, 40–45 cm Profile S, 0–40 cm Unit 5, 40–50 cm Unit 5, 0–10 cm Unit 1, 0–10 cm Unit 2, 50–60 cm Unit 1, 60–75 cm Unit 2, Strat. B Unit 2, 30–40 cm Unit 1, 40–50 cm Unit 1, 60–70 cm Unit 1, 10–26 cm Unit 2, Strat. A

MLT 800–650 MLT 800–650 MLT 800–650 650 680–630 640 660–620 630 650–560 622 656–543 L 650–200 L 650–200 L 650–200 L 650–200 L 650–200 L 650–200 516 574–469 491 527–447 491 532–440 468 507–417 440 470–400 420 460–360 294 395–260 290 330–270 264 299–154 264 299–154 243 273–102 115 243–0

6,176 6,176 3,000 10,053 6,245 9,733 6,907 8,428 11,836 23,036 13,702 15,794 14,942 5,100 3,576 3,741 44,118 8,942 3,894 10,425 10,022 33,088 1,701 48,700 11,875

Arnold and Munns 1994 Arnold and Munns 1994 Arnold and Munns 1994 Christy 2004 Christy 2004 Christy 2004 Kennett 1998 Arnold and Munns 1994 Arnold and Munns 1994 Arnold and Munns 1994 Arnold and Munns 1994 Arnold and Munns 1994 Arnold and Munns 1994 Walker and Snethkamp 1984 Walker and Snethkamp 1984 Walker and Snethkamp 1984 Kennett 1998 Christy 2004 Christy 2004 Walker and Snethkamp 1984 Christy 2004 Kennett 1998 Kennett 1998 Kennett 1998 Walker and Snethkamp 1984

note: M2, middle period, phase 2; M3, middle period, phase 3; M4, middle period, phase 4; MLT, middle to late transition; L, late period.

206

L AT E H O L O C E N E

1984). Little evidence for bead manufacturing exists in levels dating to the Middle Holocene, although finished beads found at sites and in burial lots indicate that they were being produced somewhere in the Chumash interaction sphere (King 1990). Trace amounts of Olivellabead-manufacturing detritus were found in column samples dating to between 3,000 and 1,300 BP, and densities increased significantly between 1,300 and 800 BP. The highest densities of Olivella detritus between 1,300 and 650 BP occurred at sites on western Santa Cruz and eastern Santa Rosa islands (map 14A). The increase in bead manufacturing between 1,300 and 800 BP cooccurs with the first evidence for microlith production on eastern Santa Cruz Island (Arnold 1985, 1990a; Kennett 1998). Small microdrills, fashioned from these microliths, are frequently found in direct association with Olivella-shell-bead manufacturing at sites on all of the islands (Arnold 1987; Kennett 1998; Rozaire 1993), and shell polish found on these drill bits indicates that they were used to perforate shell beads (Preziosi 2001). Three types of microliths were produced on the islands: trapezoidal, triangular, and triangular with retouch (see figure 24b, c, and e; figure 25h–j). Trapezoidal microblades emerged as the dominant microlith type between 1,100 and 800 BP. Triangular microblades with retouch were the dominant form after 800 BP (Arnold 1987, 1990a), and triangular microliths (no retouch) are found associated with artifact collections dominated by trapezoidal and triangular (with retouch) forms and are considered to be temporally undiagnostic (Arnold 1987). Significant reduction in the standard deviations of microdrill-shaft width and thickness occurred from the Middle (1,100–800 BP) to Late Periods (650–200 BP) and suggest that the production of microliths became more standardized through time (Arnold et al. 2001; Preziosi 2001, 162). Microlith production always appears to have been focused on the eastern end of Santa Cruz Island, where quality chert outcrops and subsurface chert beds occur naturally (Arnold 1987; Arnold et al. 2001; Kennett 1998; Preziosi 2001). Smaller concentrations of chert are present elsewhere on the islands, and there is some evidence that these sources were used to produce microblades during this period (Erlandson et al. 1997). However, it is unlikely that the production of microliths on other parts of the northern Channel Islands ever reached the scale visible on eastern Santa Cruz. The character of chert mining and microlith production changed significantly on eastern Santa Cruz between 1,100 and 200 BP (map 15). Trapezoidal microblades were produced by people at coastal

L AT E H O L O C E N E

207

villages and small interior workshops on eastern Santa Cruz Island during the late Middle Period (1,050–800 BP) (Arnold 1987, 1990a; Kennett 1998; Perry 2003). Chert for trapezoidal-microblade production was also mined from ridgetop outcrops and boulders west of El Montañon on a geologic contact between the Monterey and volcanic formations (Arnold 1987). Ten outcrop/boulder quarries have been documented along this contact, each associated with a small trapezoidal-microblade-production locus (see figure 1A). Subsurface chert deposits were mined at a small pit quarry on a ridgetop in upper Scorpion Canyon for the purpose of producing trapezoidal microblades (Kennett 1998). Trapezoidal-microblade workshops have also been identified at a series of caves and ridgetop sites in Scorpion drainage and at coastal village sites at Smugglers Cove, Scorpion Anchorage, and Prisoners Harbor (Clifford 2001). Much of the microlith production during this time occurred in the interior of the island, near the easily accessible chert outcrops west of El Montañon (Perry 2003). The dispersed nature of microlith-production sites suggests opportunistic exploitation of chert and an industry that was not highly organized or specialized. Relatively high densities of island chert cores were discovered in sites on western Santa Cruz, far removed from the source area, and suggest that raw material, rather than finished microliths, was exported from eastern Santa Cruz during this interval (Arnold et al. 2001). Trapezoidal-microblade production on the northern Channel Islands waned between 800 and 650 BP, as the production of prepared triangular microblades intensified (Arnold 1987, 1990a; Arnold et al. 2001). The character and intensity of prepared triangularmicroblade production on eastern Santa Cruz differed considerably from the trapezoidal-microblade industry. Small microblade workshops (triangular with retouch) in the interior were used, but most prepared triangular-microblade production occurred at coastal villages and at a limited number of large chert quarries (Arnold 1987; Kennett 1998). Prepared triangular microblades were not produced in great numbers at ridgetop quarries along the contact zone, but it is likely that chert was mined from these locations and transported to coastal villages at China and Prisoners harbors for the production of microblades. Chert was also extracted from boulders at CA-SCrI-392, a coastal quarry located at the northern extreme of the “contact zone” on the beach at China Harbor East of El Montañon (map 15B; Arnold 1987). East of El Montañon, subsurface chert deposits were intensively mined at two known locations, CA-SCrI-610 and -611, for

map 15. Villages, chert quarries, and manufacturing stations on eastern Santa Cruz Island dating to between (A) 1,050 and 800 BP and (B) 800 and 200 BP (map produced by D. Kennett with the assistance of J. Bartruff).

L AT E H O L O C E N E

209

the purpose of manufacturing prepared triangular microliths. The most impressive of these sites is CA-SCrI-610, the largest chert quarry on the northern Channel Islands (figure 29; Kennett 1998). Extensive microlithic workshops were associated with both chert quarries, but the production of prepared microblades also occurred at coastal villages at Scorpion Anchorage and Smugglers Cove (Arnold 1990a; Kennett et al. 2000). Triangular-microblade cores are rare on western Santa Cruz (Arnold et al. 2001), and on the outer islands of Santa Rosa and San Miguel (Kennett 1998), a distributional pattern suggesting that finished microliths were exported from eastern Santa Cruz, rather than the raw material that was traded during the late Middle Period. The export of microliths indicates that these tools were distributed more efficiently during the Late Period and that groups living on eastern Santa Cruz were limiting access to this valuable source of stone (Arnold et al. 2001). More localized production of microliths, on a much larger scale, suggests a more specialized activity and is consistent with the more standardized character of microliths during this time.

Emergent Sociopolitical Complexity All human societies are complex, to a greater or lesser degree (Erlandson 2002a; McGuire 1996; Tainter 1996), but there are changes evident in the archaeological record of the northern Channel Islands that suggest increases in economic, social, and political integration during the Late Holocene. Ethnohistoric records, the geographic distribution of villages, and rank-size analysis suggest that the island Chumash were organized primarily at the village level—economically and politically (see chapter 5; Johnson 2000). Approximately 22 named communities existed on the northern Channel Islands during the historic period, and mission records indicate that they ranged in size and sociopolitical importance (see Johnson 1993). A limited number of ethnohistorical accounts suggest the periodic integration of island communities under one or two influential chiefs (Johnson 1988, 2001). On Santa Cruz Island, the village of Kaxas (CA-SCrI-240), located on the northeastern end of the island at Prisoners Harbor (see map 12), was well positioned to be an important economic center, a natural port of trade between the islands and the mainland. The village of Qshiwqshiw (CA-SRI-85, -87, -88, -187), situated at the mouth of Old Ranch Canyon

figure 29. (A) Aerial view of CA-SCrI-610, a large quarry site on eastern Santa Cruz Island where subsurface chert deposits were mined prehistorically; (B) surface of CA-SCrI-610 showing pits and the remnants of extensive prehistoric mining activity. (Photos by S. Spaulding.)

L AT E H O L O C E N E

211

on Santa Rosa, was also a natural geographic center for the outer islands (Johnson 1993). By historic contact, Kaxas was the second largest village on Santa Cruz and appears to have been an important economic and political center (Johnson 1982, 1993). At least one chief lived in this village historically, and he had two daughters: one that went on to marry a chief from another village, and a second who became a chief herself. The latter moved to Liyam (CA-SCrI-1), located on the south side of the island, and is said to have united all island communities under her authority (Arnold 2001; 290; Johnson 1982, 1993). Baptismal records suggest that Qshiwqshiw was one of the larger communities on the outer islands and had at least four chiefs in residence (Johnson 1993). It was also an important center for manufacturing plank canoes (Brown 1967, 16), watercraft that had great economic, social, and political value (Arnold 1995; see below). In island Chumash society, male chiefs were often polygynous and each of their wives would move to their native community (patrilocal residence; Johnson 1988, 2000). Owing to the preponderance of matrilocality in Chumash society (Johnson 1988, 2000), these chiefs had a great deal of influence on men from other island and mainland communities. This surely acted as an integrative force, and it is possible that Kaxas and Qshiwqshiw emerged as important political centers, perhaps each as the center of a simple chiefdom by the historic period. Several lines of evidence support the idea that the sociopolitical complexity evident at historic contact was well established on the islands by the beginning of the Late Period (~650 BP; Arnold 2001; Kennett 1998). Many of the named historic island communities were occupied at least by the beginning of this interval, suggesting continuity between the historic and Late periods. Site size hierarchies are often used by archaeologists to identify politically important villages, but the size of primary village sites on the northern Channel Islands does not vary substantially. Differential house sizes are also used in other regions to differentiate between high- and low-status households within settlements, but the houses visible at Late Period sites do not differ greatly and may not be a good predictor of wealth and status in island Chumash society (Arnold 2001, 290). More subtle differences at the site of Kaxas point to the prominence of this community during the Late Period (650–200 BP). Important households at this location appear to be differentiated by the use of rare or long-lasting materials in house construction (e.g., redwood, whalebone; Arnold

212

L AT E H O L O C E N E

2001). The artifact assemblage at the site also shows a high frequency of finished craft products from island communities and a concentration of exotic items that were manufactured on the mainland (Arnold and Graesch 2001). These data highlight the importance of this community as a center of trade during the Late Period (Arnold 2001). Direct evidence for a simple chiefdom centered on Kaxas during the Late Period is currently lacking, but if the historic record is any indication, this community was likely an important political and economic center by this time. By the Late Period (650–200 BP), there is also a substantial amount of evidence for community-based craft specialization on the northern Channel Islands, and some archaeologists have argued that such specialization required an organized labor force only afforded by political centralization and control (Arnold 2001, 288). Geographically localized production of microliths at Late Period villages and quarries on eastern Santa Cruz Island suggests that individuals living in these communities were relatively specialized craftspeople (Arnold 1987; Kennett 1998; Perry 2003), probably producing microliths on a part-time basis along with participating in other subsistence-related activities (e.g., fishing). This hypothesis is supported by the relative absence of microlith production on western Santa Cruz and the outer islands of Santa Rosa and San Miguel (Arnold 1987; Kennett 1998). Communities also specialized in the manufacture of shell beads to varying degrees. Olivella beads were manufactured in most Late Period communities, but a majority of this production occurred on western Santa Cruz and eastern Santa Rosa (Arnold 1987; Kennett 1998). Arnold (2001, 288) has convincingly argued that bead makers living on western Santa Cruz and the outer islands were dependent upon microlith manufacturers who controlled a spatially circumscribed island resource (also see Arnold 1987, 1990a). The corpus of data now available for the northern Channel Islands provides strong evidence that craft specialization was an integral component of Late Period society (Arnold 2001; Arnold and Graesch 2001; Arnold and Munns 1994; Arnold et al. 2001; Kennett 1998; Preziosi 2001). Microliths and beads were more standardized and produced in greater quantities compared to earlier times (Arnold and Graesch 2001; Kennett 1998; Preziosi 2001). It should be said that these were independent specialists that produced craft items on a part-time basis (Arnold and Munns 1994), rather than full-time (attached) specialists that are often present in more complex societies

L AT E H O L O C E N E

213

(Brumfiel and Earle 1987). Whether or not craft production was controlled and coordinated by an elite class remains an open question. The mere existence of craft specialization does not logically indicate hierarchical ranking, as tacitly assumed in the past (Clark 1995; Costin 1991; Cross 1993; Earle 1981), and such specialization can exist without centralized control (Clark 1995, 281). People may turn to craft production for maintaining exchange relations within a developing regional economy, as a diversification strategy for increasing diet breadth in the face of population-dependent reductions in higherranked prey items, or as a risk-minimization strategy for creating social connections with people living in different ecological settings —useful in the face of uncertain environmental conditions (Kennett 1998). Rather than attributing all island craft production to an elite, controlled labor force, it is more likely that island Chumash people produced craft items for a variety of economic and political reasons and that elite members of society tried to control certain aspects of production and distribution starting as early as the late Middle Period (~1,300 BP) The origin of the plank canoe has featured prominently in debates regarding the development of sociopolitical complexity in the Santa Barbara Channel region (Arnold 1995, 2001; Davenport et al. 1993; Gamble 2002). Boats of some kind (e.g., balsa rafts, dugouts) were used on the islands starting in the Terminal Pleistocene (Erlandson 2002b), but plank canoes were larger and more seaworthy, and because of this would have promoted cross-channel exchange/social contact and facilitated the hunting of large pelagic fish (e.g., swordfish) and cetaceans (e.g., dolphins). During the historic period, chiefs or other elite members of society were known to own these impressive boats and a secret men’s society (The Brotherhood of the Tomol) had developed by this time to build, maintain, and operate them (Hudson et al. 1978). Plank-canoe owners were in a good position to manipulate cross-channel exchange for their own economic and political benefit. The appearance of swordfish bones, tools used to manufacture these boats, plankcanoe parts, and canoe models suggest that the plank canoe was in full use by ~1,500 BP (Gamble 2002), and the social and political institutions centered on the use of these boats had likely solidified by the Late Period (~650 BP; Arnold 1995). When and how the sociopolitical complexity evident during the Late Period first emerged on the islands is open to debate. The evolutionary schemes put forward to explain the development of

214

L AT E H O L O C E N E

sociopolitical complexity emphasize gradual or punctuated mechanisms to varying degrees (Arnold 1987, 1991, 1992a, 2001; King 1990; Martz 1984, 1992; Raab and Larson 1997), but the most prudent models allow for both gradual and punctuated elements (Erlandson 2002a, 327). Mortuary data often provide the best evidence for the development of ranking and hierarchical social organization, because differential burial patterns often reflect ascribed- or achieved-status positions. The quantity and types of artifacts buried with infants are particularly telling because they did not have the opportunity to achieve status and wealth during their lifetimes, and, therefore, suggest some form of hereditary status. High-status burials in the Santa Barbara Channel region often contain a diverse range of funerary objects, including canoe effigies and plank-canoe parts. King (1990) used mortuary data from the northern Channel Islands to argue for the development of ranked societies and ascribed status by the end of the Early Period (~2,400 BP). The same data have been used to demonstrate that hierarchical social organization developed on the island by ~800 BP (Arnold 2001). In both instances, the raw data needed to evaluate these claims are not presented, and the resolution of this debate awaits a comprehensive analysis of mortuary remains from island contexts, coupled with a DNA study that would allow for the reconstruction of familial relationships and provide the data needed to link wealth and status with heredity. Although the precise timing for the development of ranking, or ascribed status, in Chumash society is unclear, the archaeological record does indicate significant cultural changes on the northern Channel Islands between 1,500 and 650 BP. As outlined above, these included greater settlement stability (Kennett 1998), new technologies (e.g., plank canoe, bow and arrow; Arnold 1995; Gamble 2002), economic intensification (Kennett and Kennett 2000), increased lethal violence (Lambert 1994), and control of island resources by some communities (Arnold 1987, 1997, 2001). Greater settlement stability occurred in the context of population increase and the expansion of primary villages on the islands during the Holocene. Dietary changes are also evident on the islands through time (e.g., increases in diet breadth), prior to the development of more intensive fishing after 1,300 BP. Osteological data indicate increases in sublethal violence between 2,300 and 1,300 BP, along with clear indications of subsistence stress (e.g., decreases in stature and increases in health problems). Technological developments were founded upon existing knowledge in the region: the circular

L AT E H O L O C E N E

215

fishhook evolving from compound hooks, large oceangoing boats developing from smaller watercraft (e.g., dugouts or balsa rafts), and the bow and arrow descending from the atlatl and spear (Erlandson 2002a, 325; Erlandson and Rick 2002a). Although there is a great deal of evidence for cultural continuity during the Holocene, the pace of evolutionary change increased on the islands during the Late Holocene, particularly after 1,300 BP, and the evidence for widespread change across Southern California has stimulated a spirited debate centered on why these cultural developments occurred (Arnold 1987, 2001; Kennett and Kennett 2000; Raab and Larson 1997). Central to this debate are the specific environmental and biotic changes that may have triggered sociopolitical and economic developments. In particular, the debate has focused on the relative roles of changing terrestrial and marine ecosystems on cultural developments between 800 and 650 years ago (Arnold 1997; Arnold et al. 1997; Jones and Kennett 1999; Jones et al. 1999; Lambert 1997; Lambert and Walker 1991; Raab and Larson 1997). Arnold (1992a, 1997) and others (Arnold and Tissot 1993; Arnold et al. 1997; Colten 1993, 1994, 1995; Colten and Arnold 1998) have focused on sociopolitical and economic responses to marine and terrestrial climatic stresses on the northern Channel Islands, particularly a reported interval of elevated sea-surface temperatures and low marine productivity between 800 to 650 years ago. Elevated seasurface temperatures were inferred from a long-standing marine paleoclimatic sequence for the region and faunal assemblages from Santa Cruz Island (Arnold 1992b, 2001; Arnold and Tissot 1993; Colten 1994, 1995; Pisias 1978, 1979). In contrast, Raab and Larson (1997) suggested that punctuated cultural changes during this interval were stimulated by increasing violence and competition between individuals resulting from widespread drought and associated reduced terrestrial resources (also see Jones and Kennett 1999; Jones et al. 1999; Kennett and Kennett 2000; Yatsko 2000). New paleoclimatic data for the region suggest that these economic, social, and political changes occurred during a period of sustained terrestrial drought and high marine productivity that mark the interval between 1,500 to 650 years ago (see figures 12 and 27). This was also one of the most climatically unstable intervals in the entire Holocene. Climatic instability and resource stress associated with persistent drought co-occur with an increase in lethal violence (Lambert 1994, 1997) and lend some support to the idea that competitive aggression played a role in these cultural developments. Environmental change is only one of several important variables that contributed to long- and

216

L AT E H O L O C E N E

short-term processes that contributed to the emergence of social hierarchies on the northern Channel Islands. The ability of HBE to integrate multiple contextual variables, with an emphasis on responses to social and environmental circumstances, makes it an ideal framework for exploring the causes and consequences of these changes. These variables are explored in the following chapter using a nested series of HBE models.

chapter 8

Synthesis

A series of models derived from human behavioral ecology (HBE) were presented at the outset of this book. These models established a conceptual framework for this study and provided a set of hypotheses, or predictions, regarding diet choice, intensification, etc., that are evaluated in this chapter based on the available archaeological and ethnohistorical data from the northern Channel Islands. Overall, the goodness-of-fit between the model predictions and the data are supportive of, and show substantial promise for, the application of HBE to similar problems in other maritime settings. All HBE models are contingent upon the ecological context (complex social and environmental variables) that can effectively change predicted outcomes. Therefore, fruitful avenues for continued research are discovered when data are not consistent with these models. In other words, this study is not an end in itself but points out trends in the data that appear to support model predictions, while also providing a means of defining future research questions.

Diet Breadth Several trends visible in the archaeological record of the northern Channel Islands are consistent with the predictions of the diet-breadth model. In its simplest form, this model predicts that foragers will choose the combination of foods that maximizes net intake of energy 217

218

SYNTHESIS

(energy acquired less energy expended). If forager density increases, as it appears to do on the northern Channel Islands through time, the model predicts that a disproportionate number of high-ranked prey will be killed and overall encounter rates with these animals will decrease (resource depression). This, in turn, stimulates an expansion in diet breadth to lower-ranked resources in order to compensate for overall decreases in net energy intake. Changes in shellfish assemblages on the islands through the Holocene are generally consistent with this prediction. Except for the earliest Holocene (10,000–8,000 BP; see below), most Early and Middle Holocene sites contain relatively high numbers of red (H. rufescens) and black (H. cracherodii) abalone shells. These are the largest species found in the rocky intertidal zone, and they were easy prey for early populations on the islands. Late Holocene mollusk assemblages are more diverse, and the overall abundance of abalone, particularly H. rufescens, was substantially reduced. This pattern is the clearest at sites, or site complexes, that were occupied during both the Middle and the Late Holocene (CA-SCrI-333, CA-SRI-147, CA-SMI-528/602, CASMI-261/603; Kennett 1998; Vellanoweth et al. 2000; Walker et al. 2000; Wilcoxon 1993). Environmental conditions (especially cool seasurface temperatures) likely favored red abalone in intertidal habitats during parts of the Early and Middle Holocene (Glassow 1993a; Glassow et al. 1994) and partially explain the abundance of this large species in midden deposits of this age. Reductions in both abalone species at Bay Point on San Miguel Island (CA-SMI-261/603) from the Middle to the Late Holocene, however, suggest that predation pressure also played a significant role. It appears that human-induced degradation of intertidal habitats was offset with increases in diet breadth through time. This pattern parallels increases in the dietary importance of fish through the Late Holocene (after ~3,000 BP). Starting with the earliest occupation of the northern Channel Islands at Daisy Cave (~12,500 BP), fish from nearshore kelp bed habitats were always important in the diet of these islanders (Erlandson et al. 1996a; Rick et al. 2001a). There is also some evidence from Daisy Cave for the early dominance of fish in the diet (Rick et al. 2001a; discussed below), but the general trend in the data shows increases in the quantity and types of fish targeted through the Holocene. The earliest indicators of this are the development of the single-piece shell fishhook between 2,500 and 2,100 BP (Rick et al. 2002) and increases in fish bone densities in middens after this time (Glassow

SYNTHESIS

219

1993a; Kennett 1998; Vellanoweth et al. 2000). Exponential increases in fish bone densities occur after 1,300 BP and co-occur with the development of a larger array of fishing tackle (see figure 28). The diversity of fish taxa taken during this period also expands at some locations with the development of new fishing technology (e.g., plank canoe, toggling harpoons), and fish from midwater, deep-ocean, and open-ocean habitats become more common in island faunal assemblages after 1,300 BP, at least on Santa Cruz Island ( Bernard 2001; Pletka 2001b). All of these changes represent increases in diet breadth, and, ultimately, intensified subsistence strategies and are consistent with the predictions of the diet-breadth model. In an indirect way, evidence for intensified production of nonfood trade items (e.g., beads) and increased trade through the Late Holocene are also supportive of the diet-breadth model. Edible plant foods are relatively scarce on the islands, when compared to the adjacent mainland (see chapter 3; Arnold 1987, 1991; Kennett 1998), and the ethnohistorical record suggests that islanders were heavily subsidized by mainland plant foods (e.g., acorns) obtained through trade (Arnold 2001; King 1976). Logically then, the manufacture of trade items by islanders during the Late Holocene could be interpreted as an attempt to expand their dietary breadth. Groundstone manufacture on San Miguel Island, where the availability of edible plant food was more limited than on Santa Rosa and Santa Cruz, started as early as 2,500 BP and was most intensive between 1,250 and 1,000 BP. Production of groundstone suggests that plant foods were imported to San Miguel, or that these tools were exported, possibly in exchange for plant foods (Conlee 2000). If the former is true, this represents a conscious effort by these islanders to expand their dietary choices. Increases in bead and microlith production after 1,300 BP, if exchanged for plant foods, can equally be interpreted in this way. Recovery in island contexts of plant foods native to the mainland dating to after 800 BP are supportive of this idea (Martin and Popper 2001), but more macrobotanical work is needed to confirm this result. Whether or not there was temporal overlap between humans and pygmy mammoths on the islands remains an open question, but the extinction of these large animals is worth evaluating within the framework of the diet-breadth model. These animals were the size of a small steer (Agenbroad 2000b), and they were the largest mammals on the islands during the Terminal Pleistocene. Nonhuman predators were absent, and because these mammoths were not in direct contact with

220

SYNTHESIS

human populations earlier in the Pleistocene (if they ever were) they would have been ecologically naive. If the earliest human colonists arrived on Santarosae (see chapter 3) prior to the extinction of pygmy mammoths, then they would have been high-ranked prey and targeted by human hunters. The absolute number of mammoths on these offshore islands was always small, so per capita encounter rates would have declined rapidly. The diet-breadth model predicts that as encounter rates with mammoths decreased, lower-ranked prey (e.g., pinnipeds, shellfish, and fish) would have been added to the diet. Dwarf mammoth extinction at the end of the Pleistocene (Agenbroad 2000b), and the early focus on maritime resources evident at Daisy Cave (Erlandson et al. 1996a), are both consistent with these predictions. Human-induced extirpation of the largest, ecologically naive, prey, followed by the expansion of diet breadth, is well documented in similar island settings (Anderson 1989; Kennett et al. n.d.a; Patton 1996; Steadman 1989; Steadman and Kirch 1990; Steadman and Rolett 1996; Steadman et al. 1990), but whether or not this occurred on the northern Channel Islands remains a testable hypothesis. The diet-breadth model is used primarily to predict changes in diet choice, but dietary decisions also influence where settlements are positioned and how land is used. Population-dependent decreases in diet breadth often promote residential stability, because less time is spent pursuing higher-ranked prey items and lower-ranked prey usually occur in relatively high densities near settlements. This usually corresponds with an increase in food processing and storage. Increasing diet breadth from the Middle to Late Holocene parallels reductions in settlement mobility. The more temporary logistical encampments, common during the Middle Holocene, were replaced by more permanent settlements in the Late Holocene (e.g., south side of Santa Rosa, north coast of San Miguel). Decreases in the use of interior residences signals the greater stability of coastal communities. A greater reliance on both fish and fishing tethered people to more stable settlements along the coast, because the building, maintenance, and protection of watercraft would have become more important. This was particularly true after 1,300 BP, as people started fishing more frequently in offshore habitats (Bernard 2001; Pletka 2001b). Several datasets are incongruent with the general patterns of resource depression and dietary expansion/intensification visible on the islands, and they are thus inconsistent with the predictions of the diet-breadth model as currently conceived. On the northern Channel

SYNTHESIS

221

Islands, the dietary importance of fish generally increases through the Holocene, particularly after 3,000 BP with the development of new fishing technology regionally. The initial expansion of diet breadth, followed by intensified use of certain fish species, parallels increases in human populations and the expansion of primary villages along the coasts of the larger islands, a pattern consistent with the diet-breadth model. Early evidence (10,290–8,980 BP and 9,270–8,480 BP) for the dietary importance of fish at Daisy Cave (San Miguel Island) is not consistent with these predictions (Rick et al. 2001). Edible meat estimates suggest that fish contributed 50%–65% of the meat represented by faunal remains, and overall fish-bone densities are unusually high for sites of this age. The species composition of the assemblage suggests that people fished in nearshore rocky and kelp-bed habitats. Although the species diversity was not as great as at some Late Holocene sites (see Pletka 2001b), the fish-bone densities in some of these early strata exceed many Late Holocene sites (~3,000–250 BP; Rick et al. 2001a, 609). In portions of the Early and Middle Holocene (after 8,500 BP) intertidal habitats appear to have been highly productive and contained several large species of shellfish (red abalone in particular, but also black abalone and California mussels). The low costs associated with harvesting this readily available meat source explains its prominence in the island diet through the Middle Holocene, followed by populationdependent impacts on intertidal resources and the increased reliance on fish visible in the archaeological record during the Late Holocene. Extending this logic to the Early Holocene, the data from Daisy Cave suggest that the shellfish populations in the intertidal zone within the vicinity of Bay Point were not productive enough to sustain even a small human population. The productivity of intertidal habitats is affected by human predation or unfavorable environmental conditions. The archaeological record suggests that populations on the islands were relatively small during the Early Holocene; therefore, it is most probable that environmental conditions were unfavorable at the time. Seasurface conditions between 11,000 and 9,500 BP were warm, and marine productivity was highly unstable, but generally low from 11,000 to 8,700 BP (see figure 11). Coastal habitats were also unstable through this interval because of the rising sea level. Indeed, red and black abalone are both uncommon in the earliest strata at Daisy Cave, and the shellfish assemblage includes several smaller species (e.g., small mussels, tegula, limpets). In this context, the exploitation of multiple habitats

222

SYNTHESIS

and species (expanded diet breadth) indicated by equal reliance upon shellfish and fish is predictable based on the microeconomic theory that underpins the diet-breadth model. Also potentially at odds with the diet-breadth model is the low density of pinniped (seals, sea lions, etc.) bone in most island middens dating to the Early and Middle Holocene, suggesting that these animals played only a supplemental dietary role during this early time. Harbor seals frequent the waters around all of the islands today, and, prior to their extirpation historically, sea otters were also common in the kelp-bed habitats fringing these islands. In addition, larger pinniped species (sea lions, fur seals, elephant seals) are present in the waters surrounding the islands and are particularly common around San Miguel where large numbers aggregate in breeding colonies seasonally. Hildebrandt and Jones (1992) first noted that the larger pinniped species on the California coast also need to come ashore to breed, and usually do so in larger rookeries. One of the largest rookeries in California is located on the west end of San Miguel at Point Bennett (see chapter 3). Based on the microeconomic principles of the diet-breadth model, Hildebrandt and Jones argued that these animals would have been preferentially targeted by human hunters during the Early Holocene and, if available, would have been one of the primary meat sources for early peoples of the California coast. The argument follows that population-dependent increases in predation pressure through the Holocene would have reduced the availability of these animals and pushed breeding populations to more inaccessible islands and offshore rocks. In this context, dietary breadth would have expanded to include other resources (e.g., shellfish and fish), including smaller, more elusive marine mammals (sea otters and harbor seals) that do not aggregate in breeding colonies and can give birth at sea. Based on this model, and the trends in human population on the islands, one would expect higher densities of pinniped bone in Early/Middle Holocene sites when compared to sites dating to the Late Holocene. Thus far, the faunal records from the northern Channel Islands do not support this model. In most island villages, pinnipeds played a minor dietary role, a pattern that is not surprising given the modern-day distribution of these animals on the islands. People living on Santa Rosa and Santa Cruz would have had access to sea otters, harbor seals, and the occasional stray fur seal or sea lion. Given the high costs of capturing these smaller animals, it makes perfect economic sense for people to focus on rich shellfish beds and to

SYNTHESIS

223

ultimately turn to fishing in the Late Holocene in the face of increasing island populations and human-induced reductions in intertidal productivity. Low densities of pinniped bone in Early/Middle Holocene middens on San Miguel Island are more surprising, as is the sudden increase in pinniped bone at sites located at Point Bennett during the Late Holocene (after 1,500 BP). One possible explanation for this pattern is that marine conditions did not favor the formation of island rookeries during the Early and Middle Holocene, and it would follow that these animals were more available during the Late Holocene when conditions were more favorable. Millennial-scale fluctuations in sea-surface temperature (SST) and marine productivity occurred throughout the Holocene, but SSTs were generally warmer, and marine conditions less productive, during the Early and Middle Holocene, when compared to the Late Holocene (see figure 11). However, several cool and productive periods are evident in the Early/Middle Holocene, and it is unlikely that these environmental conditions explain the virtual absence of pinnipeds in the archaeological record. Three additional factors may account for the low frequency of pinnipeds in early sites. First, pinnipeds may have been concentrated on the western end of the San Miguel, as they are today, where the evidence for hunting at these locations would have been inundated during the final stages of the marine transgression. This could explain the absence of sites containing pinniped bone until ~7,000 BP, but not the limited amount of bone in Middle Holocene sites in the vicinity of Point Bennett (e.g., CA-SMI528). Second, human hunting during the Terminal Pleistocene and earliest Holocene reduced marine mammal populations and restricted rookeries to offshore rocks and islands (e.g., Castle Rock) where the costs of exploitation exceeded the benefits of doing so. Third, human populations were centered on the northern coast of Santa Rosa, far removed from rookeries on San Miguel, and central place foraging returns were low because of transportation costs (distance) and limitations imposed by the available maritime technology needed to transport meat back to more centrally located primary villages (see next section). This could also partially explain the sudden increase in pinniped bone in the first villages at Point Bennett that were established after 1,500 BP. Prior deposits appear to be more ephemeral and the product of logistical foraging. Future research will be required to determine the combination of factors that account for the patterns visible in the archaeological record of San Miguel.

224

SYNTHESIS

Central Place Foraging Central place foraging (CPF) theory provides another set of tools for exploring and interpreting the archaeological record on the northern Channel Islands. The same microeconomic principles underlying the diet-breadth model also form the basis of CPF, and the use of one does not negate the other. One of the benefits of the diet-breadth model is its simplicity, and it has been shown here to be a useful framework for generating hypotheses regarding prehistoric foraging behavior in this maritime context. Deviations from model expectations provide a new source of questions and hypotheses and an opportunity to evaluate the data within a different interpretive framework. One of the assumptions of the diet-breadth model is that resources are distributed evenly across space and that all foragers have the same goal: to maximize net energy intake under a given set of changing ecological circumstances. The fact that several trends in the data are consistent with these predictions suggests that, on average, people on the islands were acting rationally and with a tendency toward optimization. On the other hand, the model is not well equipped for finer-grained analyses of foraging behavior in environments where resources are not distributed evenly, such as the northern Channel Islands, and where the foraging goals of men and women, or the young and old, are potentially different. With this in mind, the primary villages that first appeared on the northern Channel Islands between 8,500 and 7,500 BP (CA-SRI-3) are conceptualized as central places and the accumulation of material at these locations represents an agglomeration of activities carried out by people with different foraging goals (men, women, elderly, children, high status, low status, etc.). Important factors for selecting primary village locations on the islands included the availability of drinking water, firewood, and well-drained flat terrain that would allow the establishment of a community. Based on these criteria, the best locations for establishing more permanent communities were on the northern and eastern sides of Santa Rosa and the western and northeastern side (Prisoners Harbor) of Santa Cruz. Perennial streams flow most vigorously in these larger drainages (drinking water) and firewood was more available from interior forests on the larger islands. Driftwood was more widely available along the coasts of all of the islands. More marginal settlement locations with respect to water availability were located at the mouths of smaller drainages around the periphery of Santa Cruz (north, south,

SYNTHESIS

225

and east shores) and Santa Rosa (south and southwestern shores), and along the north coast of San Miguel Island. Fresh drinking water was least available on Anacapa and the southern coast of San Miguel. Social factors (e.g., territoriality) limiting the location of settlements increased through the Holocene with population growth, environmental infilling, and controlled access to resource patches by certain groups (Arnold 2001; Kennett 1998). Beyond these base-level criteria, CPF theory provides a set of principles for estimating the optimal settlement locations on the islands. The model predicts that, all other variables being equal, foragers will select residential base locations that maximize the net central place foraging returns given the pursuit, handling, and transport costs of resources from different patches (Cannon 2003; see chapter 2). Given the high productivity of intertidal and kelp-bed habitats, and the depauperate nature of terrestrial resources on the islands (e.g., absence of large terrestrial mammals and limited floral communities; see chapter 3), it is predictable that the first primary villages would have been located at the mouths of the largest drainages on Santa Rosa and Santa Cruz. Reduced transport costs associated with the use of boats would have also favored coastal locations (Ames 2002). Due to increases in marine productivity from east to west along this island chain, and the higher concentrations of marine mammals on San Miguel Island (see chapter 3), primary village locations would have been favored at locations farther to the west (Santa Rosa and western Santa Cruz). The presence of a reasonably sized estuary on eastern Santa Rosa would also have favored early settlement. Population-dependent reductions in the productivity of intertidal habitats at these locations probably resulted in the expansion of diet breadth, more distant logistical forays, and more intensive subsistence activities (e.g., fishing). The archaeological data for early settlement and the expansion of primary villages through the Holocene are generally consistent with the predictions of CPF theory. The most concentrated activities during the Early Holocene occur on the northwestern shore of Santa Rosa Island, in the vicinity of Arlington Canyon (CA-SRI-3, -6, -173), and the presence of a formal cemetery at Tecolote Point (just west of Arlington Canyon) suggests that this area was a focal point of settlement by between 8,000 and 7,500 BP. With a settlement base on the northeast coast of Santa Rosa, people had access to a variety of productive intertidal and nearshore habitats locally, but they could also travel to more distant locations on San Miguel to hunt sea mammals or collect shellfish

226

SYNTHESIS

from intertidal habitats that were not suffering the same impact as those in the immediate vicinity of primary villages. Oak and Torrey Pine groves in island interiors could also be targeted from the northeast coast of Santa Rosa, along with mollusks from the estuary at the mouth of Old Ranch Canyon. The importance of the Arlington area as a primary village location persisted through the Middle Holocene as communities expanded to additional locations along the north coast of Santa Rosa and the western end of Santa Cruz. This expansion continued through the Late Holocene when the first definitive evidence for primary village settlement is visible on the south coast of Santa Rosa and the north coast of San Miguel. The initial establishment of primary villages on the north coast of Santa Rosa did not preclude the use of resources from other, more distant, communities through logistical foraging and/or sexual division of labor. This is most visible on the outer islands during the Middle Holocene with the periodic occupation of interior residences and the logistical exploitation of intertidal resources from coastal locations that were not permanently occupied (see map 9). Thin lenses of shell at multiple locations along the coast, interpreted here as logistical encampments used for collecting and processing shellfish (abalone and mussels), are by far the most common site type from the Middle Holocene on the northern Channel Islands. The most obvious of these lenses are red abalone middens (Glassow 1993b). Limited tool and faunal diversity at these sites is indicative of highly specialized subsistence activities, and seasonality data suggest periodic rather than perpetual harvesting of shellfish. The removal of heavy red abalone shells where they were harvested suggest that foragers were trying to maximize net delivery rate of this resource to primary villages located elsewhere on the islands. These data indicate that each community had a large foraging range that was relatively unrestricted by territorial behavior. The periodic establishment of small residential bases in island interiors is an interesting pattern when viewed through the lens of CPF theory. Some of these sites could easily be logistical encampments that were repeatedly reoccupied to exploit seasonally available plant foods in the interior. Others appear to be more substantial and may represent the periodic establishment of semipermanent residential bases in the interior. Faunal assemblages at these locations indicate that foods from coastal habitats (shellfish, fish, and sea mammals) were transported to these interior locations (Clifford 2001). This suggests the temporary relocation of at least a segment of the population to interior areas. The

SYNTHESIS

227

geographic distribution of these sites on Santa Rosa indicates that plant foods from native grasslands (e.g., blue dick bulbs) and coastal sage brush (e.g., sage seeds) communities were targeted, and limited seasonality data suggest that shellfish were transported to these locations during several seasons. This pattern supports the idea that plant communities were productive enough, at least periodically, to warrant the relocation of residential bases. It is possible that people harvested plant foods en masse seasonally and transported them to primary villages on the coast where they were stored for later consumption. Climatic conditions during the Early and Middle Holocene suggest more consistent rainfall on the island, and it is likely that plant communities were more productive overall when compared with the Late Holocene. These conditions may also have favored periodic relocation of settlements to island interiors. Osteological data (stable isotopic data and dental records) from island burial populations also suggest that plant foods (carbohydrates) were an important part of the Middle Holocene diet (Goldberg 1993; Walker and Erlandson 1986), and they are consistent with the evidence for interior residences dating to this interval. These data also indicate that women consumed more plant foods than men, who in turn appear to have had greater access to marine foods (fish and marine mammals). Dietary differences between the sexes may reflect a division of labor and are suggestive of varied central place foraging activities. Work parties, composed primarily of men, may have traveled to distant locations on San Miguel and southern Santa Rosa to harvest abalone en masse or hunt marine mammals that were periodically abundant in some locations. During these trips, marine foods would have been most available and may explain their dietary emphasis. Similarly, work parties composed primarily of women may have collected plant foods in island interiors and had greater access to these foods. Even if family residences relocated to the interior, men may have traveled to the coast to collect marine foods more frequently to provision their families who were collecting seasonally available plant foods for storage in bulk. Mollusk assemblages in Middle Holocene sites are highly varied, and CPF theory may help bring some of this variability into focus. Some sites are composed of large red abalone shells, others are dominated by black abalone and California mussel shells. Multicomponent sites often contain large red abalone shells in basal deposits that are absent in upper levels dating to later in the Middle Holocene (e.g., CA-SRI-147; CA-SCrI-333). Black abalone shells become less frequent in sites dating

228

SYNTHESIS

to later times and are replaced by a more diverse range of species. Some of this variability is explained by environmental changes through the Holocene or spatial variation in SSTs and/or marine productivity (e.g., cooler and more productive toward the west). Other changes were because of localized predation pressure on larger species. Based on CPF theory, predation pressure is not expected to be even across space. Human impacts on intertidal productivity would be most severe near primary villages where larger numbers of people (men, women, old, and young) could collect these easy prey. Intertidal habitats at a distance from these central places would have been less impacted. The spatial structure of mollusk assemblages through the Middle Holocene is not well established, and future work will be needed to test this hypothesis. Comparison of mollusk assemblages at CA-SRI-116 and CA-SRI109, dating to about the same period (6,500–5,000 BP), provides some support for this idea. CA-SRI-116 was a primary village site located on the northeast side of Santa Rosa (see table 12). CA-SRI-109, a red abalone site that is also located on the northeastern side of Santa Rosa, appears to have been used logistically at about the same time (see seasonality data in figure 21). The mollusk assemblage at CA-SRI-116 is composed of medium-sized California mussel shells and other smaller species. Abalone (red or black) is rare. Large abalone shells (primarily H. rufescens) are common at CA-SRI-109, where predation pressure on intertidal habitats in the vicinity was likely less than on the northeast coast of Santa Rosa near CA-SRI-116. CPF theory may also partially explain the anomalous pattern of pinniped exploitation at Point Bennett on San Miguel. The faunal assemblages at two sites in the vicinity of the modern-day pinniped breeding colony at this location (CA-SMI-528/602) show increased hunting of pinnipeds from the Middle to Late Holocene, a pattern that is inconsistent with the predictions of the diet-breadth model. The Middle Holocene strata at CA-SMI-528 are thin lenses of shell interpreted as logistical encampments for collecting and processing shellfish. The lowest stratum is a classic red abalone midden, and the stratum above it contains black abalone and other species. Pinniped bone is present in these strata and suggests that these animals were in the vicinity, probably on the most inaccessible stretches of coast and offshore rocks (see above). If the logistical encampments were left behind by people who were based out of primary villages on northeastern Santa Rosa, then travel costs to distant locations on San Miguel were high. CPF theory predicts that foragers will be more selective as the distance to a resource

SYNTHESIS

229

patch increases and pursue only the highest-ranked prey. All resources collected will be processed with attention to travel costs and toward maximizing the net delivery rate to central places. This suggests that once logistical foraging parties reached the western tip of San Miguel, the costs associated with hunting pinnipeds in the rough and dangerous waters surrounding Point Bennett, combined with the costs of transporting meat back to primary villages on Santa Rosa, exceeded the benefits of doing so, particularly when large abalone were readily available in the intertidal zone. Larger densities of pinniped bone in Late Holocene strata at the same sites, inferred to represent increased hunting, are partly explained by reduced transportation costs with the establishment of a primary village at Point Bennett by ~1,500 BP. The costs of hunting pinnipeds on offshore rocks were also reduced at this time with improved maritime (e.g., plank canoe) and hunting technology that was available by this time (e.g., bow and arrow, toggle harpoon). The Late Holocene expansion of primary villages along the coasts of Santa Rosa and Santa Cruz, and along the north coast of San Miguel, coupled with increases in territoriality, imposed new limitations on central place foraging. Foraging ranges were more restricted and the impact on intertidal habitats was more severe due to residential, rather than logistical, harvesting strategies. Within this context, the expansion of diet breadth, hunting marine mammals offshore, and a greater focus upon fishing, can be viewed as an attempt to maintain or increase central place foraging returns in the face of demographic expansion followed by population growth under more socially circumscribed conditions.

Intensification and the Ideal Free Distribution Population growth and settlement expansion can also be modeled using a concept from population ecology known as the ideal free distribution (IFD; see chapter 2). In its simplest form, the ideal free distribution provides a framework for examining settlement patterns through time on the northern Channel Islands, and when combined with central place foraging theory it can be used as an effective model for economic intensification. To reiterate the model briefly, habitats are ranked according to their suitability given a specific set of economic strategies (e.g., farming vs. foraging; shellfishing vs. fishing). The model can be applied on multiple spatial scales (e.g., mainland vs. island; coast vs.

230

SYNTHESIS

interior), but suitability in this case is defined based on the central place foraging returns offered by different drainages on the islands. The model predicts that individuals will select the most suitable habitat (ideal) and are unlimited in their ability to do so (free). If habitats have the same suitability then populations should be evenly distributed between them. Of course, interference with the effective use of habitats increases with population growth as do social restrictions in the use of some areas. Population growth in these habitats generally degrades their suitability (e.g., resource depression) as carrying capacity is approached. The model predicts that population will expand to the second-ranked habitats, when the suitability of each is equivalent. This will continue until all habitable locations are occupied in an area. The suitability of a habitat can also be increased to a certain degree with greater population densities (e.g., economies of scale associated with forest clearance for agriculture) and changes in subsistence practices or technology (e.g, improved fishing technology). As argued earlier, the top-ranked settlement locations (habitats) on the northern Channel Islands were on the coast adjacent to the mouths of the largest or best-watered drainages. These occur on northern and eastern Santa Rosa (Arlington Canyon, Cañada Verde, Lobo Canyon, Old Ranch Canyon) and western and northeastern Santa Cruz (Central Valley and Prisoners Harbor drainage). At these locations people had access to perennial water, a diversity of riparian resources (plant foods), shellfish from rocky intertidal and sandy beach habitats, and fish from nearshore kelp forests. The small estuary at the mouth of Old Ranch Canyon also provided additional brackish water species of shellfish. Logistical foraging access to various plant communities was also greater on Santa Cruz and Santa Rosa, when compared to San Miguel and Anacapa. This included oak and pine forests that provided valuable sources of food (acorns and pinenuts), fuel, and building material. Owing to increasing marine productivity from east to west, and the greater availability of marine mammals on San Miguel, settlement locations on northwestern Santa Rosa may have been preferable. Second-ranked settlement locations included smaller drainages on the northern, southern, and eastern coasts of Santa Cruz, southern and southwestern shores of Santa Rosa, and along the northern coast of San Miguel. Because of more limited water availability, and the lack of plant diversity, Anacapa and the southern coast of San Miguel are predicted to be the lowest-ranked habitats on the islands.

SYNTHESIS

231

The available settlement data from the northern Channel Islands show temporal and spatial structures that are consistent with the predictions of the ideal free distribution. Although some of the best evidence for early use of the islands comes from Daisy Cave on San Miguel (Erlandson et al. 1996a), the first clear evidence for more permanent settlement comes from CA-SRI-3, a site located on the northwestern coast of Santa Rosa, near the mouth of Arlington Canyon. The exact nature of settlement at this location awaits further research, but the presence of a residential midden and a formal cemetery (Erlandson 1994; Orr 1968) suggests that a primary village was established at this site between 8,000 and 7,500 BP, and several sites located at the mouth of Old Ranch Canyon (eastern Santa Rosa) point to early village settlement at this location (Rick et al. n.d.). The Arlington area (CA-SRI-3, -4, -5) continued to be an important locus of settlement throughout the Middle Holocene (7,500–3,000 BP), but settlements were also founded at other locations along the northern coast of Santa Rosa (CA-SRI-41, -116), and possibly at one interior location on the south side of the island (Jolla Vieja; CA-SRI-147). At least one primary village (CA-SCrI-333) was started on the western end of Santa Cruz during the Middle Holocene. During the Late Holocene (3,000–250 BP), primary village settlements expanded to the south and southwestern sides of the island. Primary villages also proliferated on Santa Cruz, with communities evenly spaced around the island periphery. Permanent settlement expanded to San Miguel, where nodes of settlement were well established at Cuyler Harbor, Otter Harbor, and Point Bennett by 1,500 BP. Archaeological evidence for primary village settlements on Anacapa is absent. This also appears to be the case on the southern coast of San Miguel. The expansion of primary villages on the islands appears to have been most rapid at ~1,500 BP, and this pattern may be more consistent with the despotic variant of the ideal free distribution (ideal despotic distribution, or IDD). One of the consequences of larger populations is competition that lowers the overall suitability of a resource patch or habitat. Conspecifics deplete shared resources, and, if foraging territories are large and ill defined, neighboring groups may compete for the same sets of resources. Territoriality, or other forms of despotic behavior by individuals or groups (see Kantner 1999a), also effects the suitability of a habitat and stimulates more rapid movement of people into adjacent, less desirable, areas. The IDD highlights differential access to resources caused by increasing territorial behavior. It predicts that

232

SYNTHESIS

populations will equilibrate with disproportionate numbers in lowerranked habitats. The relatively even distribution of primary villages around the coasts of the islands after 1,500 BP, and the rapid solidification of this pattern suggests that despotic social behavior includes the positioning of some villages on coastal promontories or seacliffs, the establishment of satellite communities, and the strategic placement of interior cemeteries. It also appears that geographically circumscribed resources started to be controlled after this time (Arnold 1987; Conlee 2000; Kennett 1998). This occurred within the context of social and environmental instabilities (Kennett and Kennett 2000; see figure 27). Lethal violence increased dramatically at this time (Lambert 1994, 1997) with the introduction of the bow and arrow, and several severe droughts impacted environmental productivity and water availability across Southern California (Jones et al. 1999). It is this context that may have stimulated the more rapid spread of primary villages on the islands at this time. This rapid spread of primary villages, starting at 1,500 BP, may also have been favored by improvements in suitability of second-ranked habitats owing to technological developments or broadscale changes in the regional economy (e.g., exchange). The development of the singlepiece shell fishhook (~2,500 BP) and the plank canoe (~1,500 BP) enhanced the fishing capabilities of islanders. With a more effective means of tapping into this seemingly endless supply of food, previously second-ranked habitats may have become more attractive locations for primary village settlement. Similarly, the plank canoe promoted crosschannel exchange and plant foods (e.g., acorns), formerly best accessible from primary villages on Santa Rosa and Santa Cruz, became more available in other locations. The best evidence for this comes from the north coast of San Miguel where the production of groundstone tools increases at ~1,500 BP and suggests the importation of plant foods from mainland sources (see previous discussion). The IDF and IDD models provide a framework that considers the dynamic character of habitat suitability along with density-dependent, and density-independent, variables influencing the location and number of primary village settlements on the northern Channel Islands. The general trends in the expansion of these villages through the Holocene are congruent with the model and suggest that some of these changes were density dependent. More rapid expansion of primary village settlements at ~1,500 BP suggests contextual changes that may be inconsistent with standard IFD predictions. This may have resulted

SYNTHESIS

233

from increases in despotic behavior or changes in the suitability of second-ranked habitats because of technological innovation or changes in the availability of resources obtained through exchange with external sources. A combination of these factors appears to have been responsible for this expansion and further work will be required to determine the varying effects of these changes throughout the island chain.

Competition and the Formation of Social Hierarchies The formation of social hierarchies in the Santa Barbara Channel region is a topic of a spirited debate (Arnold 1987, 2001; Erlandson 2002a; Erlandson and Rick 2002a; Gamble et al. 2002; Johnson 2000; King 1990), and the remarkably well-preserved archaeological record on the northern Channel Islands has featured prominently in the creation and testing of hypotheses regarding the nature of these developments (Arnold 1987, 2001). By the time of European contact, it appears that Chumash society was organized hierarchically and several simple chiefdoms existed in the region (Arnold and Green 2002). The ethnohistorical record indicates that the island Chumash people were primarily organized economically and politically at the village level (Johnson 2000), but that some communities were home to leaders (chiefs) that periodically influenced political affairs beyond their own community, perhaps into loose (and unstable) confederations of villages approximating simple chiefdoms. Two communities that stand out as important political centers were Kaxas (CA-SCrI-240) on northeastern Santa Cruz and Qshiwqshiw (CA-SRI-85, -87, -88, -187) on eastern Santa Rosa (Johnson 1993). Much of the debate has focused on when social hierarchies first developed (Arnold 1987, 2001; King 1990; Martz 1984, 1992) and the social/environmental conditions that they emerged under (Arnold 1991, 2001; Kennett and Kennett 2000; Raab and Larson 1997). Further work will be required to make such determinations, but the data currently available provide a starting point for evaluating the HBE model presented in chapter 2 for the development of social hierarchies. This model serves as a point of departure for discussing the contextual variables for these developments. In this model, the formation of social hierarchies is considered to be a density-dependent phenomenon that occurs in regions where resources are unevenly distributed. Some areas are worth defending because they contain dense resources that are predictable in time and

234

SYNTHESIS

space (economic defensibility; Boone 1992; Dyson-Hudson and Smith 1978; see chapter 2). Biotic communities in coastal settings, like the northern Channel Islands, are often concentrated spatially and highly predictable temporally. Maintaining exclusive access to locations that provide the highest rates of return (central place foraging returns in this instance) requires time and effort, these costs grow as regional populations increase and the best resource patches become a source of competitive aggression (Boone 1992, 317). Under conditions of intense competition, localized populations are expected to grow, organize, and cooperate to maintain and defend access to the best resource zones. In agricultural societies, highly localized populations may improve the suitability of these habitats, at least temporarily, because all members of the group benefit from economic improvements made by the group (e.g., forest clearance, irrigation canals; economies of scale). On the other hand, highly localized populations of nonagriculturalists (e.g., coastal foragers) degrade resource patches in a way that ultimately results in decreasing net return rates. In small populations these losses may be distributed evenly, but as the population grows, competition within the group may force some people to accept an uneven share of the burden. If unoccupied territory is available, these individuals may select to emigrate and form new communities, rather than be subjugated in their native homes. Under these conditions, social hierarchies are not expected to form. Once all habitable territories are occupied or populations become circumscribed for other reasons (e.g., greater territoriality), emigration is not an option and social hierarchies would be expected to form if population density continued to increase. Given the parameters outlined above, the IFD and IDD models provide a useful framework for predicting when social hierarchies would have been favored on the northern Channel Islands. The firsttier settlement locations were positioned on the northern and eastern coasts of Santa Rosa and the western end of Santa Cruz. Primary village locations were established in these areas during the Early and Middle Holocene, but a variety of second-tier settlement locations that were used logistically during this time were not permanently occupied until after 3,000 BP (e.g., south coasts of Santa Rosa and the north coast of San Miguel). Under these demographic conditions social hierarchies would not have been favored. The expansion of primary villages to most second-tier settlement locations during the Late Holocene would have been more favorable for their development, particularly after 1,300 BP when villages were evenly distributed around the coast

SYNTHESIS

235

of Santa Rosa and Santa Cruz and along the north coast of San Miguel. Aggressive competition for resources and increased territoriality through the Late Holocene are also predicted by the model, as are all-around increases in population for maintaining and defending the best settlement locations. Population-resource imbalances caused by increasing populations or environmental degradation also would have stimulated competition within each community and created a situation where individuals would have differential access to key resources. Because of the lack of unoccupied territory on the islands after 1,300 BP, people were more likely to accept lower-quality lifestyles, compared to others in the community, rather than colonizing the most marginal parts of the islands. Once all viable settlement locations were occupied, it would take continued population growth over the course of several generations for social hierarchies to develop. In other words, social hierarchies would not be expected to emerge immediately under these conditions. Continued research will be required to test this model, but the available data are consistent with its predictions. It is still unclear how social hierarchies functioned within island Chumash society, but most archaeologists would agree that they were present in some island communities by at least 650 BP (Arnold 2001). Many of the named historic communities were in place by this time and the communities that had chiefs in residence in the contact period contain materials suggesting that they were important political and economic centers going back to at least this time (e.g., Kaxas; Arnold 2001). Burials containing status markers are also present on the islands (Arnold 2001), but more detailed study of mortuary remains will be required before a definitive statement regarding differential burial practices can be made. Although there is evidence for cultural continuity through the Holocene, several datasets indicate significant cultural changes between 1,300 and 650 BP that likely played a role in the development of social hierarchies. Settlement data and the distribution of interior cemeteries suggest that the territories surrounding primary villages solidified and became more rigid after 1,300 BP. Increases in lethal violence suggest that the social landscape was unstable and that people were competing aggressively with one another for resources. There is also evidence for economic intensification (e.g., fishing), specialized production of nonfood trade items, and greater intervillage exchange. The potential for violence would have increased the costs of colonizing already marginal, unoccupied habitats on the islands.

236

SYNTHESIS

Population increase and demographic expansion were clearly at the core of the cultural changes evident on the islands between 1,500 and 650 BP, but new paleoclimatic data for the region suggest that this interval was highly unstable climatically and that the already dry conditions prevailing through this period were interrupted by megadroughts that impacted much of Southern California (Stine 1994). Dry conditions further circumscribed populations, and environmental instability stimulated conflict and competition for access to perennial water sources. Violence was exacerbated by the introduction of the bow and arrow between 1,500 and 1,300 BP. Larger and more stable settlements emerged on the islands during this period and suggest that populations increased and that aggregations of people were favored within this context. These conditions also promoted economic intensification, craft specialization, and exchange as alternatives to aggressive competition (Kennett and Kennett 2000). Cooperative strategies, such as trade, associated with a new political system became dominant after 650 BP as violent interaction decreased regionally. It is within this context that social hierarchies were favored and certain individuals were able to control aspects of the political and economic system for their own benefit. The social ranking and hierarchical political structure likely solidified in island Chumash society under these conditions, and two simple chiefdoms may have existed on the northern Channel Islands by the historic period.

Human Behavioral Ecology and Maritime Societies The HBE models employed in this study provide a well-defined set of hypotheses that have been evaluated using the ethnohistorical and archaeological data currently available from the northern Channel Islands. This analysis supports the conclusion that the social and political complexity evident at historic contact was ultimately a product of individual behavioral responses, both competitive and cooperative, to demographic expansion, human-induced impacts to marine habitats, and periods of rapid paleoclimatic change. These trends in demography, dietary expansion, economic intensification, and increasing sociopolitical complexity evident on these islands from the Terminal Pleistocene to Late Holocene are not unique. The expansion in dietary breadth from the Pleistocene to the Holocene, commonly known as the “broad-spectrum” revolution (Flannery 1969), parallels population increase and precedes the shift from foraging to farming in several

SYNTHESIS

237

important centers of plant and animal domestication (Piperno and Pearsall 1998; Russell 1988). Mixed foraging and food producing economies often emerged in these regions and persisted for thousands of years before more intensive forms of agriculture were adopted (lowlevel food production; Smith 2001). Demographic expansion and the emergence of more intensive forms of food production provided a context for the formation of social hierarchies and the ultimate emergence of state-level societies. HBE provides a framework for analyzing dietary expansion, plant and animal domestication, mixed foraging/farming strategies, and the ultimate development of agriculture. For instance, Piperno and Pearsall (1998) have argued that the trends in resource diversification evident at the close of the Pleistocene in several regions are consistent with the diet-breadth model, the expansion in diet breadth occurring in response to decreases in the availability of highranked prey. It is also useful for explaining why farming or mixed foraging/farming strategies persisted in some areas where social and environmental circumstances favored stability, rather than change. There is clear evidence for human use of coastal and aquatic resources extending back to 150,000 years, and the exploitation of aquatic resources is tied closely to the expansion of anatomically modern humans out of Africa (Erlandson 2001). The use of coastal habitats increased in many areas after 10,000 BP (Kennett et al. n.d.b; Pálsson 1988; Yesner 1980), and maritime resources were added to an expanding set of resources targeted by foragers living close to the coast. Similar to the development of food production in other regions, economic intensification is also visible in many coastal areas (Moseley 1975; Yesner 1980). In several parts of the world (e.g., Mesoamerica, Andes, Near East, Thailand, Japan), economic intensification entailed the adoption of domesticated plants that resulted in mixed maritime foraging and food-producing strategies ultimately leading to more intensive forms of food production (e.g., Kennett et al. n.d.b). In other settings (e.g., California coast, Pacific Northwest coast), this involved the development of more sophisticated fishing technologies and intensified exploitation of near- and offshore fisheries. The positive correlation between coastal resource use and a variety of measures for social complexity (Pálsson 1988) suggests that the diversification and economic intensification in coastal settings represents one pathway toward the emergence of social hierarchies. HBE holds substantial promise for the study of maritime societies. The diet-breadth model has been used effectively to analyze resource

238

SYNTHESIS

intensification in a variety of maritime settings (Broughton 1999; Butler 2001; Jones and Hildebrandt 1995; Porcasi et al. 2000). These models highlight three variables that are crucial for the study of cultural change in maritime settings: human demography, environmental degradation, and changes in diet breadth. Even small groups of maritime foragers change the resource structure of coastal habitats and population increase, perhaps stimulated by competition for the most ideal habitats, degrade coastal biomes in predictable ways, with the highest-ranked prey (based on net foraging returns) declining more rapidly than lowerranked resources. Declining encounter rates with higher-ranked prey stimulate changes in diet breadth (diversification and intensification). The archaeological record from the northern Channel Islands is consistent with heavy exploitation and depletion of high-ranking resources and the consequent expansion of diet breadth. These patterns are also evident in New Zealand (Anderson 1983, 1989; Nagaoka 2002a, 2002b) and other islands in Oceania (Butler 2000; Kennett et al. n.d.a), northern California (Broughton 1999), and the Pacific Northwest coast (Butler 2000). Resource intensification models provide a useful framework for generating hypotheses that can be tested in other maritime settings. Deviations from model expectations provide a new source of questions and hypotheses (see Hildebrandt and McGuire 2002). Owing to the patchy and discontinuous nature of resources in coastal habitats (Yesner 1980), the transportation benefits of boat travel, and the tendency for maritime foragers to collect resources logistically, leaving large accumulations of debris at residential locations, central place foraging theory is particularly promising for archaeological research in coastal settings. In this study, CPF theory was useful for developing alternative hypotheses when the archaeological record was inconsistent with the predictions of the diet-breadth model (e.g., pinniped exploitation at Point Bennett). These hypotheses require further testing, but some aspects of the record conform well to the predictions of CPF theory. When combined with the ideal free distribution or the ideal despotic distribution, CPF theory also provides a model for the expansion of primary villages on the islands and was used here to predict when social hierarchies would have been favored on the northern Channel Islands. These models provide a useful set of tools for analyzing maritime subsistence/settlement strategies and the development of social hierarchies in other of coastal settings.

References

Adam, D. P., R. Byrne, and E. Luther. 1981. A Pleistocene and Holocene pollen record from Laguna de las Trancas, northern coastal Santa Cruz County, California. Madrono 28: 255–77. Adams, R. M. 1981. The Heartland of Cities. Chicago: University of Chicago Press. Agenbroad, L. D. 1998. New pygmy mammoth (Mammuthus exilis) localities and radiocarbon dates from San Miguel, Santa Rosa, and Santa Cruz Islands, California. In Contributions to the Geology of the Northern Channel Islands, Southern California, ed. P. W. Weigand, 169–76. Bakersfield: Pacific Section, American Association of Petroleum Geologists. 2000a. California’s Channel Islands: A one-way trip in the tunnel of doom. In The Fifth California Islands Symposium, ed. D. R. Brown, K. C. Mitchell, and H. W. Chaney, 518–24. U.S. Department of the Interior Minerals Management Service, Pacific OCS Region. 2000b. New localities, chronology, and comparisons for the pygmy mammoth (Mammuthus exilis): 1994–1998. In The Fifth California Islands Symposium, ed. D. R. Brown, K. C. Mitchell, and H. W. Chaney, 518–24. U.S. Department of the Interior Minerals Management Service, Pacific OCS Region. Alvard, M. S., and L. Kuznar. 2001. Deferred harvests: The transition from hunting to animal husbandry. American Anthropologist 103(2): 295–311. Ambrose, R. F., J. M. Engle, J. A. Coyer, and B. V. Nelson. 1993. Changes in urchin and kelp densities at Anacapa Island. In Third California Islands Symposium: Recent Advances in Research on the 239

240

REFERENCES

California Islands, ed. F. G. Hochberg, 199–209. Santa Barbara, CA: Santa Barbara Museum of Natural History. Ames, K. M. 2002. Going by boat: The forager-collector continuum at sea. In Beyond Foraging and Collecting: Evolutionary Change in Hunter-Gatherer Settlement Systems, ed. B. Fitzhugh and J. Habu, 19–52. New York: Kluwer/ Plenum. Anderson, A. J. 1983. Faunal depletion and subsistence change in the early prehistory of southern New Zealand. Archaeology in Oceania 18: 1–10. 1989. Prodigious Birds: Moas and Moa Hunting in Prehistoric New Zealand. Cambridge: Cambridge University Press. Anderson, D. G., and J. C. Gillam. 2000. Paleoindian colonization of the Americas: Implications from an examination of physiography, demography, and artifact distribution. American Antiquity 65: 43–66. Antevs, E. 1948. Climatic changes and pre-white man in the Great Basin, with emphasis on glacial times and post-glacial times. University of Utah Bulletin 38: 168–91. 1952. Climatic history and the antiquity of man in California. University of California Archaeological Survey Reports 16: 23–31. 1955. Geologic-climatic dating in the West. American Antiquity 20: 317–35. Arnold, J. E. 1985. The Santa Barbara Channel Islands bladelet industry. Lithic Technology 14(2): 71–80. 1987. Craft Specialization in the Prehistoric Channel Islands, California. Berkeley: University of California Press. 1990a. Lithic resource control and economic change in the Santa Barbara Channel region. Journal of California and Great Basin Anthropology 12(2): 158–72. 1990b. An archaeological perspective on the historic settlement pattern on Santa Cruz Island. Journal of California and Great Basin Anthropology 12(1): 112–27. 1991. Transformation of a regional economy: Sociopolitical evolution and the production of valuables in Southern California. Antiquity 65: 953–62. 1992a. Complex hunter-gatherer-fishers of prehistoric California: Chiefs, specialists, and maritime adaptations of the Channel Islands. American Antiquity 57: 60–84. 1992b. Cultural disruption and the political economy in Channel Islands prehistory. In Essays on the Prehistory of Maritime California, ed. T. Jones,

REFERENCES

241

129–44. Center for Archaeological Research at Davis, 10. Davis: University of California Press. 1993. Labor and the rise of complex hunter-gatherers. Journal of Anthropological Archaeology 12: 75–119. 1995. Transportation innovation and social complexity among maritime hunter-gatherer societies. American Anthropologist 97(4): 733–47. 1996. The archaeology of complex hunter-gatherers. Journal of Archaeological Method and Theory 3(2): 77–126. 1997. Bigger boats, crowded creekbanks: Environmental stresses in perspective. American Antiquity 62(2): 337–39. ed. 2001. The Origins of a Pacific Coast Chiefdom: The Chumash of the Channel Islands. Salt Lake City: University of Utah Press. Arnold, J. E., R. H. Colten, and S. Pletka. 1997. Contexts of cultural change in insular California. American Antiquity 62(2): 300–18. Arnold, J. E., and A. P. Graesch. 2001. The evolution of specialized shellworking among the island Chumash. In The Origins of a Pacific Coast Chiefdom: The Chumash of the Channel Islands, ed. J. E. Arnold, 71–112. Salt Lake City: University of Utah Press. Arnold, J. E., and T. M. Green. 2002. Mortuary ambiguity: The Ventureno Chumash case. American Antiquity 67(4): 760–71. Arnold, J. E., and A. Munns. 1994. Independent or attached specialization: The organization of shell bead production in California. Journal of Field Archaeology 21: 473–89. Arnold, J. E., A. M. Preziosi, and P. Shattuck. 2001. Flaked stone craft production and exchange in island Chumash territory. In The Origins of a Pacific Coast Chiefdom: The Chumash of the Channel Islands, ed. J. E. Arnold, 113–32. Salt Lake City: University of Utah Press. Arnold, J. E., and B. N. Tissot. 1993. Measurement of significant marine paleotemperature variation using black abalone shells from prehistoric middens. Quaternary Research 39: 390–94. Aswani, S. 1998. Patterns of marine harvest effort in southwestern New Georgia, Solomon Islands: Resource management or optimal foraging? Ocean and Coastal Management 40: 207–35. 1999. Common property models of sea tenure: A case study from the Roviana and Vonavona Lagoons, New Georgia, Solomon Islands. Human Ecology 27(3): 417–53.

242

REFERENCES

Aswani, S., and M. W. Graves. 1998. The Tongan maritime expansion: A case in the evolutionary ecology of social complexity. Asian Perspectives 37(2): 135–64. Atwater, T. M. 1998. Plate tectonic history of southern California with emphasis on the western transverse ranges and northern Channel Islands. In Contributions to the Geology of the Northern Channel Islands, Southern California, ed. P. W. Weigand, 1–8. Bakersfield, CA: Pacific Section, American Association of Petroleum Geologists. Avila, F. A., and D. W. Weaver. 1969. Mid-Tertiary stratigraphy, Santa Rosa Island. In Geology of the Northern Channel Islands: AAPG and SEPM Pacific Section, ed. D. W. Weaver, D. P. Doerner, and B. Nolf, 48–67. Los Angeles, CA: Pacific Section, AAPG. Axelrod, D. I. 1967. Geologic history of the Californian insular flora. In Proceedings of the Symposium on the Biology of the California Islands, ed. R. N. Philbrick, 267–315. Santa Barbara, CA: Santa Barbara Museum of Natural History. 1981. Holocene climatic changes in relation to vegetation disjunction and speciation. American Naturalist 117: 847–70. Axelrod, R., and W. D. Hamilton. 1981. The evolution of cooperation. Science 211: 1390–96. Bailey, G. N., and A. S. Craighead. 2003. Late Pleistocene and Holocene coastal paleoeconomies: A reconsideration of the molluscan evidence from northern Spain. Geoarchaeology: An International Journal 18(2): 175–204. Bakun, A. 1990. Global climate change and intensification of coastal ocean upwelling. Science 247: 198–201. Bamforth, D. B. 2002. Evidence and metaphor in evolutionary archaeology. American Antiquity 67: 435–52. Bamforth, D. B., and A. C. Spaulding. 1982. Human behavior, explanation, archaeology, history and science. Journal of Anthropological Archaeology 1: 179–95. Barlow, K. R. 1997. Foragers who farm: Economics of maize agriculture and the ecology of the Fremont. Ph.D. dissertation, University of Utah, Salt Lake City. 2002. Predicting maize agriculture among the Fremont: An economic comparison of farming and foraging in the American Southwest. American Antiquity 67(1): 65–88.

REFERENCES

243

Barlow, K. R., P. R. Henrikson, and D. Metcalfe. 1993. Estimation load size in the Great Basin: Data from conical burden baskets. Utah Archaeology 27–37. Barlow, K. R., and D. Metcalfe. 1996. Plant utility indices: Two Great Basin examples. Journal of Archaeological Science 23(3): 351–71. Bartholomew, G. A. 1967. Seal and sea lion populations of the California Islands. In Proceedings of the Symposium on the Biology of the California Channel Islands, ed. R. N. Philbrick, 229–44. Santa Barbara, CA: Santa Barbara Botanic Garden. Barton, K. R., and G. A. Clark, eds. 1997. Rediscovering Darwin: Evolutionary Theory in Archaeological Explanation. Washington, D.C.: American Anthropological Association. Basgall, M. E. 1987. Resource intensification among hunter-gatherers: Acorn economies in prehistoric California. Research in Economic Anthropology 9: 21–52. Beaton, J. M. 1991. Extensification and intensification in central California prehistory. Antiquity 65(249): 946–52. Beckerman, S. 1983. Carpe diem: An optimal foraging approach to Barí fishing and hunting. In Adaptive Responses of Native Amazonians, ed. R. B. Hames and W. T. Vickers, 269–99. New York: Academic Press. 1991. Barí spear fishing: Advantages to group formation? Human Ecology 19(4): 529–54. Begossi, A. 1992. The use of optimal foraging theory in the understanding of fishing strategies: A case from Sepetiba Bay (Rio de Janeiro State, Brazil). Human Ecology 20(4): 463–75. Behl, R. J., and J. P. Kennett. 1996. Brief interstadial events in the Santa Barbara Basin, NE Pacific, during the past 60 kyr. Nature 379: 243–46. Bellwood, P. 2001. Early agriculturalist population diasporas? Farming, languages, and genes, Annual Review of Anthropology 30: 181–207. Bemis, B. E., H. J. Spero, J. Bijima, and D. W. Lea. 1998. Reevaluation of the oxygen isotopic composition of planktonic foraminifera: Experimental results and revised paleotemperature equations. Paleoceanography 13(2): 150–60.

244

REFERENCES

Bennyhoff, J. A., and R. E. Hughes. 1987. Shell Bead and Ornament Exchange Networks Between California and the Western Great Basin. New York: The American Museum of Natural History. Benson, A. S. 1982. The Noontide Sun: The Fieldnotes and Unpublished Manuscripts of the Rev. Stephen Bowers. Master’s thesis, Department of Anthropology, California State University, Northridge. Berger, R. 1980. Early Man on Santa Rosa Island. In The California Islands: Proceedings of a Multidisciplinary Symposium, ed. D. M. Power, 73–78. Santa Barbara, CA: Santa Barbara Museum of Natural History. 1982. The woolly mammoth site, Santa Rosa Island, California. In Peopling of the New World, ed. J. E. Ericson, R. E. Taylor, and R. Berger. Ballena Press Anthropological Papers 23: 163–70. Berger, R., and P. Orr. 1966. Early Man on Santa Rosa Island, California, II. Proceedings of the National Academy of Sciences 56: 1678–82. Berger, R., and R. Protsch. 1989. UCLA Radiocarbon dates XI. Radiocarbon 31: 55–67. Bernal, P. A., and J. A. McGowan. 1981. Advection and upwelling in the California Current. In Coastal Upwelling, ed. F. A. Richards, 381–99. Washington, D.C.: American Geophysical Union. Bernard, J. L. 2001. The origins of open-ocean and large species fishing in the Chumash region of Southern California. Master’s thesis, Department of Anthropology, University of California, Los Angeles. Berry, B. J. 1961. City size distributions and economic development. Economic Development and Culture Change 9: 573–88. Bettinger, R. L. 1991. Hunter-Gatherers: Archaeological and Evolutionary Theory. New York: Plenum Press. Bettinger, R. L., R. Malhi, and H. McCarthy. 1997. Central place models of acorn and mussel processing. Journal of Archaeological Science 24: 887–99. Bettinger, R. L., and P. J. Richerson. 1996. The state of evolutionary archaeology: Evolutionary correctness, or the search for common ground. In Darwinian Archaeologies, ed. H. D. G. Maschner, 221–32. New York: Plenum Press.

REFERENCES

245

Bickel, P. 1978. Changing sea levels along the California coast: Anthropological implications. Journal of California Anthropology 5: 6–20. Binford, L. R. 1968. Post-Pleistocene adaptations. In New Perspectives in Archaeology, ed. S. R. Binford and L. R. Binford, 313–41. Chicago: Aldine. 1980. Willow smoke and dog’s tails: Hunter-gatherer settlement systems and archaeological site formation. American Antiquity 45: 4–20. 1983. In Pursuit of the Past. London: Thames and Hudson. 1989. “New Archaeology,” then and now. In Archaeological Thought in America, ed. C. C. Lamberg-Karlovsky, 50–62. New York: Cambridge University Press. 2001. Constructing Frames of Reference: An Analytical Method for Archaeological Theory Building Using Hunter-Gatherer and Environmental Data Sets. Berkeley: University of California Press. Biondi, F., D. R. Cayan, and W. H. Berger. 1997. Dendroclimatoloy of Torrey Pine (Pinus torreyana Parry ex Carr.). The American Midland Naturalist 138(2): 237–51. Bird, D. W. 1997. Behavioral ecology and the archaeological consequences of central place foraging among the Meriam. In Rediscovering Darwin: Evolutionary Theory in Archaeological Explanation, ed. C. M. Barton and G. A. Clark, 291–306. Washington, D.C.: American Anthropological Association. Bird, D. W., and R. L. Bliege Bird. 1997. Contemporary shellfish gathering strategies among the Meriam of the Torres Strait Islands, Australia: Testing predictions of a central place foraging model. Journal of Archaeological Science 24: 39–63. 2000. The ethnoarchaeology of juvenile foragers: Shellfishing strategies among Meriam children. Journal of Anthropological Archaeology 19: 461–76. Birdsell, J. B. 1953. Some environmental and cultural factors influencing the structuring of Australian aboriginal populations. American Naturalist 87: 171–207. Blackburn, T. C., ed. 1975. December’s Child: A Book of Chumash Oral Narratives, collected by J. P. Harrington. Berkeley: University of California Press. Blanton, R. E. 1976. The role of symbiosis in adaptation and sociocultural change in the Valley of Mexico. In The Valley of Mexico: Studies in Prehispanic Ecology and Sociology, ed. E. R. Wolf, 181–201. Albuquerque: University of New Mexico Press.

246

REFERENCES

Bliege Bird, R., and D. W. Bird. 1997. Delayed reciprocity and tolerated theft: The behavioral ecology of food sharing strategies. Current Anthropology 38(1): 49–78. 2001. The hunting handicap: Costly signaling in human foraging strategies. Behavioral Ecology Sociobiology 50: 9–19. Bliege Bird, R., E. A. Smith, and D. W. Bird. 2001. The hunting handicap: Costly signaling in male foraging strategies. Behavioral Ecology and Sociobiology 50: 9–19. Blitz, J. H. 1988. Adoption of the bow in prehistoric North America. North American Archaeologist 9(2): 123–45. Boles, J. R. 1997. Geologic introduction and lithologic units. In Santa Cruz Island: Geology Field Trip Guide, ed. J. R. Boles and W. Landry, 3–18. San Diego, CA: San Diego Association of Geologists. Boone, J. L. 1992. Competition, conflict, and the development of social hierarchies. In Evolutionary Ecology and Human Behavior, ed. E. A. Smith and B. Winteralder, 301–38. New York: Aldine de Gruyter. Boone, J. L., and E. A. Smith. 1998. Is it evolution yet? A critique of evolutionary archaeology. Current Anthropology 39: 141–73. Boserup, E. 1965. The Conditions of Agricultural Growth: The Economics of Agrarian Change Under Population Pressure. Chicago, IL: Aldine. Bowser, B. 1993. Dead fish tales: Analysis of fish remains from two Middle Period sites on San Miguel Island. In Archaeology on the Channel Islands of California: Studies in Subsistence, Economics, and Social Organization, ed. M. A. Glassow, 95–135. Salinas, CA: Coyote Press. Bremner, C. 1932. Geology of Santa Cruz Island, Santa Barbara County, California. Santa Barbara Museum of Natural History Occasional Papers 1: 1–33. Bright, R. C. 1966. Pollen and seed stratigraphy of Swan Lake, southeastern Idaho: Its relation to regional vegetational history and to Lake Bonneville history. Tebiwa 9: 1–47. Brink, K. H., D. Halpern, A. Huyer, and R. L. Smith. 1983. The physical environment of the Peruvian upwelling system. Progress in Oceanography 12: 285–305.

REFERENCES

247

Broughton, J. M. 1994a. Late Holocene resource intensification in the Sacramento Valley, California: The vertebrate evidence. Journal of Archaeological Science 21: 501–14. 1994b. Declines in mammalian foraging efficiency during the Late Holocene, San Francisco Bay, California. Journal of Anthropological Archaeology 13: 371–401. 1997. Widening diet breadth, declining foraging efficiency, and prehistoric harvest pressure: Icthyofaunal evidence from the Emeryville shellmound. Antiquity 71: 845–62. 1999. Resource depression and intensification during the Late Holocene, San Francisco Bay. Berkeley: University of California Press. Broughton, J. M., and J. F. O’Connell. 1999. On evolutionary ecology, selectionist archaeology, and behavioral archaeology. American Antiquity 64(1): 153–65. Brown, A. K. 1967. The Aborginal Population of the Santa Barbara Channel. Berkeley: University of California Archaeological Survey Reports. Brown, J. L. 1969. The buffer effect and productivity in tit populations. American Naturalist 103: 347–54. Browne, D. R. 1994. Understanding the oceanic circulation in and around the Santa Barbara Channel. In The Fourth California Islands Symposium: Update on the Status of Resources, ed. W. L. Halvorson and G. J. Maender, 27–34. Santa Barbara, CA: Santa Barbara Museum of Natural History. Brumfiel, E. M. 1992. Distinguished lecture in archaeology: Breaking and entering the ecosystem-gender, class and faction steal the show. American Anthropologist 94: 551–66. 1994. Factional competition and political development in the New World: An introduction. In Factional Competition and Political Development in the New World, ed. E. M. Brumfiel and J. W. Fox, 3–14. Cambridge: Cambridge University Press. Brumfiel, E. M., and T. K. Earle. 1987. Specialization, exchange, and complex societies: An introduction. In Specialization, Exchange, and Complex Societies, ed. E. M. Brumfiel and T. K. Earle, 1–9. Cambridge: Cambridge University Press. Buikstra, J. E. 1981. Mortuary practices, paleodemography, and paleopathology: A case study. In The Archaeology of Death, ed. R. I. Chapman, I. Kinnes, and K. Randsborg, 123–32. Cambridge: Cambridge University Press.

248

REFERENCES

Burley, D. V., and W. R. Dickinson. 2001. Origin and significance of a founding settlement in Polynesia. Proceedings of the National Academy of Sciences of the United States of America 98(20): 11829–31. Butler, V. L. 2000. Resource Depression on the Northwest Coast of North America. Antiquity 74: 649–61. 2001. Changing fish use on Mangaia, southern Cook Islands: Resource depression and the prey choice model. Osteoarchaeology 11(1–2): 88–100. Butterworth, J. B., J. A. Jones, and S. Jones. 1993. Soil forming factors, morphology and classification—Santa Cruz Island, California. In Third California Islands Symposium: Recent Advances in Research on the California Islands, ed. F. G. Hochberg, 39–56. Santa Barbara, CA: Santa Barbara Museum of Natural History. Byrne, R. 1979. Commentary on “Archaeology and California’s Climate.” Journal of California and Great Basin Anthropology 1(1): 196–98. Byrne, R., C. Busby, and R. F. Heizer. 1979. The Altithermal revisited: Pollen evidence from the Leonard Rockshelter. Journal of California and Great Basin Anthropology 1(2): 280–94. Cannariato, K. G., J. P. Kennett, and R. J. Behl. 1999. Biotic response to late Quaternary rapid climate switches in Santa Barbara Basin: Ecological and evolutionary implications. Geology 27(1): 63–66. Cannon, M. 2000. Large mammal relative abundance in Pithouse and Pueblo Period archaeofaunas from south western New Mexico: Resource depression in the Mimbres-Mogollon? Journal of Anthropological Archaeology 19: 317–47. 2003. A model of central place forager prey choice and an application to faunal remains from the Mimbres Valley, New Mexico. Journal of Anthropological Archaeology 22: 1–25. Carneiro, R. L. 1970. A Theory of the Origin of the State. Science 169(3947): 733–38. 1978. Political expansion as an expression of the principle of competitive exclusion. In Origins of the State: The Anthropology of Political Evolution, ed. R. Cohen and E. R. Service, 205–23. Philadelphia, PA: Institute for the Study of Human Issues. 1988. The Circumscription Theory, Challenge and Response. American Behavioral Scientist 31(4): 497–511. Cashdan, E. 1983. Territoriality among human foragers: Ecological models and an application to four Bushman groups. Current Anthropology 24(1): 47–66.

REFERENCES

249

ed. 1990. Risk and Uncertainty in Tribal and Peasant Economies. San Francisco, CA: Westview Press. 1992. Spatial organization and habitat use. In Evolutionary Ecology and Human Behavior, ed. E. A. Smith and B. Winterhalder, 237–68. New York: Aldine de Gruyter. Chagnon, N. A. 1997. Yanomamö. 5th ed. Fort Worth, TX: Harcourt Brace College. Chapman, R., I. Kinnes, and K. Randsborg, eds. 1981. The Archaeology of Death. Cambridge: Cambridge University Press. Charnov, E. L. 1976. Optimal foraging, the marginal value theorem. Theoretical Population Biology 9: 367–90. Chess, K. A., K. McEachern, and D. H. Wilken. 2000. Island jepsonia (Jepsonia malvifolia) demography on Santa Rosa and Santa Cruz Islands, California. In The Fifth California Islands Symposium, ed. D. R. Brown, K. C. Mitchell, and H. W. Chaney, 155–57. U.S. Department of the Interior Minerals Management Service, Pacific OCS Region. Chinn, B. D., and P. W. Weigand. 1994. Petrology and geochemistry of volcanic clasts from the Miocene Becher’s Bay Formation, Santa Rosa Island, California. In The Fourth California Islands Symposium: Update on the Status of Resources, ed. W. L. Halvorson and G. J. Maender, 256–65. Santa Barbara, CA: Santa Barbara Museum of Natural History. Christy, J. 2004. Bead Manufacturing on the Northern Channel Islands: An Evolutionary Perspective. Master’s thesis, Department of Anthropology, California State University, Long Beach. Claassen, C. 1986. Shellfishing seasons in the prehistoric southeastern United States. American Antiquity 51(1): 21–37. Clark, J. E. 1995. Craft specialization as an archaeological category. Research in Economic Anthropology 16: 267–94. Clark, P. U., J. X. Mitrovica, G. A. Milne, and M. E. Tamisiea. 2002. Sea-level fingerprinting as a direct test for the source of global meltwater pulse IA. Science 295: 2438–41. Clifford, R. A. 2001. Middle Holocene Hilltop and Ridgeline Settlement on the Northern Channel Islands of California: A Study of Evolutionary Stability. Master’s thesis, Department of Anthropology, California State University, Long Beach.

250

REFERENCES

Cobb, M., and L. A. K. Mertes. 2000. Mapping gradations among vegetation communities on Santa Cruz Island with field and remote sensing data. In The Fifth California Islands Symposium, ed. D. R. Brown, K. C. Mitchell, and H. W. Chaney, 143–54. U.S. Department of the Interior Minerals Management Service, Pacific OCS Region. Cohen, M. N., and G. J. Armelagos. 1984. Paleopathology at the Origins of Agriculture. New York: Academic Press. Cole, K. L., and G. W. Lui. 1994. Holocene paleoecology of an estuary on Santa Rosa Island, California. Quaternary Research 41: 326–35. Cole, K. L., and E. R. Wahl. 1996. Mid- to Late Holocene stability in both natural populations of Torrey Pine. Abstracts of the 14th Biennial Meeting of the American Quaternary Association, Northern Arizona University. 2000. A Late Holocene paleoecological record from Torrey Pines State Reserve, California. Quaternary Research 53: 341–51. Collins, P. W. 1991a. Interaction between island foxes (Urocyon littoralis) and Indians on islands off the coast of southern California: I. Morphologic and archaeological evidence of human assisted dispersal. Journal of Ethnobiology 11(1): 51–81. 1991b. Interaction between island foxes (Urocyon littoralis) and Native Americans on islands off the coast of southern California: II. Ethnographic, archaeological, and historical evidence. Journal of Ethnobiology 11(2): 205–29. Colten, R. H. 1993. Prehistoric Subsistence, Specialization, and Economy in a Southern California Chiefdom. Ph.D. dissertation, Archaeology Program, University of California, Los Angeles. 1994. Prehistoric animal exploitation, environmental change, and emergent complexity on Santa Cruz Island, California. In The Fourth California Islands Symposium: Update on the Status of Resources, ed. W. L. Halvorson and G. J. Maender, 201–14. Santa Barbara, CA: Santa Barbara Museum of Natural History. 1995. Faunal exploitation during the Middle to Late Period Transition on Santa Cruz Island, California. Journal of California and Great Basin Anthropology 17(1): 93–120. 2001. Ecological and Economic Analysis of Faunal Remains from Santa Cruz Island. In The Origins of a Pacific Coast Chiefdom: The Chumash of

REFERENCES

251

the Channel Islands, ed. J. E. Arnold, 199–220. Salt Lake City: University of Utah Press. Colten, R. H., and J. E. Arnold. 1998. Prehistoric marine mammal hunting on California’s Northern Channel Islands. American Antiquity 63(4): 679–701. 2000. Native uses of marine mammals on Santa Cruz and San Miguel Islands. In The Fifth California Islands Symposium, ed. D. R. Brown, K. C. Mitchell, and H. W. Chaney, 623–27. U.S. Department of the Interior Minerals Management Service, Pacific OCS Region. Colvin, W. I., III, and S. R. Gliessman. 2000. Fennel (Foeniculum vulgare) management and native species enhancement on Santa Cruz Island, California. In The Fifth California Islands Symposium, ed. D. R. Brown, K. C. Mitchell, and H. W. Chaney, 184–89. U.S. Department of the Interior Minerals Management Service, Pacific OCS Region. Conlee, C. A. 2000. Intensified middle period ground stone production on San Miguel Island. Journal of California and Great Basin Anthropology 22(2): 374–91. Connolly, T., J. M. Erlandson, and S. E. Norris. 1995. Early Holocene basketry from Daisy Cave, San Miguel Island, California. American Antiquity 60: 309–18. Cook, S. F. 1964. The aboriginal population of upper California. Actas y Memorias, XXXV International Congress of Americanists, Mexico 1962. Vol. 3: 397–403. 1976. The Population of California Indians, 1769–1970. Berkeley: University of California Press. Cooke, I. D. 1940. Indian trails. Santa Barbara Museum of Natural History Talk 15: 5–7. Corry, P. M., and A. K. McEachern. 2000. Native shrub recovery in nonnative annual grasslands, California Channel Islands. In The Fifth California Islands Symposium, ed. D. R. Brown, K. C. Mitchell, and H. W. Chaney, 182–83. U.S. Department of the Interior Minerals Management Service, Pacific OCS Region. Costin, C. L. 1991. Craft specialization: Issues in defining, documenting, and explaining the organization of production. In Archaeological Method and Theory, Vol. 3, ed. M. B. Schiffer, 1–56. Tucson: University of Arizona Press. Cross, J. R. 1993. Craft specialization in nonstratified societies. Research in Economic Anthropology 14: 61–84.

252

REFERENCES

Dall, W. H. 1874. The Lord of the Isles. Overland Monthly 12: 522–26. Davenport, D., J. Timbrook, and J. R. Johnson. 1993. Chumash and the swordfish. Antiquity 67(255): 257–72. Davis, O. K. 1992. Rapid climatic change in coastal Southern California inferred from pollen analysis of San Joaquin Marsh. Quaternary Research 37: 89–100. Dayton, P. K., M. J. Tegner, P. E. Parnell, and P. B. Edwards. 1992. Temporal and spatial patterns of distribution and recovery in a kelp forest community. Ecological Monographs 62: 421–45. de Boer, W. F. 2000. Between the Tides: The Impact of Human Exploitation on an Intertidal. Veenendal, The Netherlands: Universal Press. de Boer, W. F., and F. A. Longamane. 1996. The exploitation of intertidal food resources in Inhaca Bay, Mozambique, by shorebirds and humans. Biological Conservation 78: 295–303. Deith, M. R. 1983. Seasonality of shell collecting determined by oxygen isotope analysis of marine shells from Asturian sites in Cantabria. In Animals and Archaeology: 2. Shell Middens, Fishes, and Birds, ed. C. Grigson and J. CluttonBrock, 67–76. Oxford: BAR International Series 183. 1985. Seasonality from shells: An evaluation of two techniques for seasonal dating of marine molluscs. In Palaeobiological Investigations Research Design, Methods and Data Analysis, ed. N. R. J. Fieller, D. D. Gilbertson, and N. G. A. Ralph, 119–30. Oxford: BAR International Series 266. 1988. Shell seasonality: An appraisal of the oxygen isotope technique, recent developments. In Environmental Analysis of Old and New World Archaeology, ed. R. E. Webb, 37–49. Oxford: BAR International Series 416. Delaney-Rivera, C. 2001. Groundstone tools as indicators of changing subsistence and exchange patterns in the coastal Chumash Region. In The Origins of a Pacific Coast Chiefdom: The Chumash of the Channel Islands, ed. J. E. Arnold, 165–82. Salt Lake City: University of Utah Press. DeLong, R. L., and S. R. Melin. 2000. Thirty years of pinniped research at San Miguel Island. In The Fifth California Islands Symposium, ed. D. R. Brown, K. C. Mitchell, and H. W. Chaney, 401–6. U.S. Department of the Interior Minerals Management Service, Pacific OCS Region. DeNiro, M. J. 1985. Postmortem preservation and alteration of in vivo bone collagen isotope ratios in relation to paleodietary reconstruction. Nature 317: 806–9.

REFERENCES

253

Dibblee, T. W., Jr., and H. E. Ehrenspeck. 1998. General geology of Santa Rosa Island, California. In Contributions to the Geology of the Northern Channel Islands, Southern California, ed. P. W. Weigand, 49–76. Bakersfield, CA: Pacific Section, American Association of Petroleum Geologists. 2000. Geology of Santa Rosa Island, California. In The Fifth California Islands Symposium, ed. D. R. Brown, K. C. Mitchell, and H. W. Chaney, 115–26. U.S. Department of the Interior Minerals Management Service, Pacific OCS Region. Dillehay, T. D. 1997. Monte Verde, A Late Pleistocene Settlement in Chile, Volume 2, The Archaeological Context and Interpretation. Washington, D.C.: Smithsonian Institution Press. 2000. The Settlement of the Americas: A New Prehistory. New York: Basic Books. Divale, W. T. 1974. Migration, external warfare, and matrilocal residence. Behavior Science Research 9: 75–133. 1984. Matrilocal Residence in Pre-Literate Society. Ann Arbor: University Microfilms International Research Press. Dixon, E. J. 2001. Human colonization of the Americas: Timing, Technology, and Process. Quaternary Science Reviews 20: 277–99. Dorman, C. E., and D. P. Palmer. 1981. Southern California summer coastal upwelling. In Coastal upwelling, ed. F. A. Richards, 44–56, Washington, D.C.: American Geophysical Union. Dunnell, R. C. 1971. Systematics in Prehistory. New York: Free Press. 1980. Evolutionary theory and archaeology. In Advances in Archaeological Method and Theory 3, ed. M. B. Schiffer, 35–99. New York: Academic Press. 1989. Aspects of the application of evolutionary theory in archaeology. In Archaeological Thought in America, ed. C. C. Lamberg-Karlovsky, 35–49. Cambridge: Cambridge University Press. Dwyer, P. 1985. A hunt in New Guinea: Some difficulties for optimal foraging theory, Man 20: 243–53. Dyson-Hudson, R., and E. A. Smith. 1978. Human territoriality: An ecological reassessment. American Anthropologist 80: 21–41.

254

REFERENCES

Earle, T. 1981. CA Comment on “Evolution of Specialized Pottery Production: A Trial Model” by Prudence M. Rice. Current Anthropology 22: 230–31. Eastwood, A. 1924. Journal of the Vancouver expedition. California Historical Society Quarterly 2: 264–340. Eerkens, J. W. 1999. Common pool resources, buffer zones, and jointly owned territories: Hunter-gatherer land and resource tenure in Fort Irwin, southeastern California. Human Ecology 27(2): 297–318. Ehorn, B., and D. Cornell. 1979a. California marine waters, areas of biological significance. Reconnaissance survey report, Anacapa Island, Ventura County, California: Resources Control Board. 1979b. Environmental assessment. Channel Islands National Monument, California: National Park Service. 1980. Environmental assessment. Channel Islands National Monument, California: National Park Service. Ember, M., and C. R. Ember. 1971. The conditions favoring matrilocal vs. patrilocal residence. American Anthropologist 73: 571–94. Engle, J. M. 1993. Distribution patterns of rocky subtidal fishes around the California Islands. In Third California Islands Symposium: Recent Advances in Research on the California Islands, ed. F. G. Hochberg, 475–84. Santa Barbara, CA: Santa Barbara Museum of Natural History. 1994. Perspectives on the structure and dynamics of nearshore marine assemblages of the California Channel Islands. In The Fourth California Islands Symposium: Update on the Status of Resources, ed. W. L. Halvorson and G. J. Maender, 13–26. Santa Barbara, CA: Santa Barbara Musuem of Natural History. Erlandson, J. M. 1984. A case study in faunalturbation: Delineating the effects of the burrowing pocket gopher on the distribution of archaeological materials. American Antiquity 49: 785–90. 1988a. Role of shellfish in prehistoric economies: A protein perspective. American Antiquity 53(1): 102–9. 1988b. Cultural evolution and paleogeography on the Santa Barbara coast: A 9,600 year 14C record from southern California. Radiocarbon 30: 25–39. 1991a. Early maritime adaptations on the Northern Channel Islands. In Hunter-Gatherers of Early Holocene Coastal California, ed. J. M. Erlandson

REFERENCES

255

and R. H. Colten, 101–11. Perspectives in California Archaeology 1. Los Angeles: Institute of Archaeology, University of California. 1991b. Shellfish and seeds as optimal resources: Early Holocene subsistence on the Santa Barbara coast. In Hunter-Gatherers of Early Holocene Coastal California, ed. J. M. Erlandson and R. H. Colten, 89–100. Perspectives in California Archaeology 1: Los Angeles: Institute of Archaeology, University of California. 1991c. Antiquity of CA-SMI-1: A multi-component site on San Miguel Island. Journal of California and Great Basin Anthropology 13(2): 273–79. 1993a. California’s coastal prehistory: A circum-Pacific perspective. Proceedings of the Society for California Archaeology 6: 23–36. 1993b. Evidence for a terminal Pleistocene human occupation of Daisy Cave, San Miguel Island, California. Current Research in the Pleistocene 10: 17–21. 1993c. Summary and conclusions. In Archaeological Investigations at CA-SBA-1731: A Transitional Middle-to-Late Period Site on the Santa Barbara Channel, ed. J. L. Gerber, 187–96. Report prepared for Exxon Company, Goleta, CA. Santa Barbara, CA: Dames and Moore. 1994. Early Hunter-Gatherers of the California Coast. New York: Plenum Press. 1997a. The Middle Holocene along the California coast. In Archaeology of the California Coast during the Middle Holocene, ed. J. M. Erlandson and M. A. Glassow, 1–10. Perspectives in California Archaeology 4. Los Angeles: Institute of Archaeology, University of California. 1997b. The Middle Holocene on the western Santa Barbara coast. In Archaeology of the California Coast during the Middle Holocene, ed. J. M. Erlandson and M. A. Glassow, 91–109. Perspectives in California Archaeology 4. Los Angeles: Institute of Archaeology, University of California. 2001. The archaeology of aquatic adaptations: Paradigms for a new millennium. Journal of Archaeological Research 9: 287–350. 2002a. Cultural Change, Continuity, and variability along the Late Holocene California coast. In Catalysts to Complexity: Late Holocene Societies of the California Coast, ed. J. M. Erlandson and T. L. Jones, 320–29. Perspectives in California Archaeology 6. Los Angeles: Cotsen Institute of Archaeology, University of California. 2002b. Anatomically modern humans, maritime voyaging, and the Pleistocene colonization of the Americas. Memoirs of the California Academy of Sciences, No. 27 (The First Americans), California Academy of Sciences. Erlandson, J. M., and K. Bartoy. 1995. Cabrillo, the Chumash, and Old World Diseases. Journal of California and Great Basin Anthropology 17: 1–15. 1996. Protohistoric California: Paradise or pandemic? Proceedings of the Society for California Archaeology 9: 304–9.

256

REFERENCES

Erlandson, J. M., and R. H. Colten. 1991. An archaeological context for early Holocene studies on the California coast. In Hunter-Gatherers of Early Holocene Coastal California, ed. J. M. Erlandson and R. H. Colten, 1–10. Perspectives in California Archaeology 1. Los Angeles: Institute of Archaeology, Los Angeles: University of California. Erlandson, J. M., T. Cooley, and R. Carrico. 1987. A fluted point fragment from the Southern California Coast: Chronology and context at CA-SBA-1951. Journal of California and Great Basin Anthropology 9: 120–28. Erlandson, J. M., and M. A. Glassow, eds. 1997. Archaeology of the California Coast during the Middle Holocene. Perspectives in California Archaeology 4. Los Angeles: Institute of Archaeology, University of California. Erlandson, J. M., and T. L. Jones, eds. 2002. Catalysts to Complexity: Late Holocene Societies of the California Coast. Perspectives in California Archaeology 6. Los Angeles: Cotsen Institute of Archaeology, University of California. Erlandson, J. M., D. J. Kennett, R. J. Behl, and I. Hough. 1997. The Cico chert source on San Miguel Island, California. Journal of California and Great Basin Anthropology 19: 124–30. Erlandson, J. M., D. J. Kennett, B. L. Ingram, D. Guthrie, D. Morris, and G. J. West. 1996a. A radiocarbon chronology for the archaeology and paleontology of Daisy Cave, San Miguel Island, California. Radiocarbon 38(2): 355–73. Erlandson, J. M., and T. C. Rick. 2002a. Late Holocene cultural developments along the Santa Barbara coast. In Catalysts to Complexity: Late Holocene Societies of the California Coast, ed. J. M. Erlandson and T. L. Jones, 166–82. Perspectives in California Archaeology 6. Los Angeles: Cotsen Institute of Archaeology, University of California. 2002b. A 9,700-year-old shell midden on San Miguel Island, California. Antiquity 76: 315–16. Erlandson, J. M., T. C. Rick, D. J. Kennett, and P. L. Walker. 2001. Dates, demography, and disease: Cultural contacts and possible evidence for Old World epidemics among the Island Chumash. Pacific Coast Archaeological Society Quarterly 37(3): 1–16. Erlandson, J. M., T. C. Rick, and R. L. Vellanoweth. 2004. Human impacts on ancient environments: A case study from California’s Northern Channel Islands, In Voyages of Discovery: The Archaeology of Islands, ed. S. Fitzpatrick. Westport, CT: Praeger/Greenwood Press.

REFERENCES

257

Erlandson, J. M., T. C. Rick, R. L. Vellanoweth, and D. J. Kennett. 1999. Marine subsistence patterns at CA-SRI-6: A 9,300-year-old midden from Santa Rosa Island, California. Journal of Field Archaeology 26(3): 255–65. Erlandson, J. M., M. A. Tveskov, and R. S. Byram. 1998. The development of maritime adaptations on the southern Northwest Coast of North America. Arctic Anthropology 35: 6–22. Erlandson, J. M., M. A. Tveskov, D. J. Kennett, and B. L. Ingram. 1996b. Further evidence for a Terminal Pleistocene occupation of Daisy Cave, San Miguel Island, California. Current Research in the Pleistocene 13: 161–68. Fagan, B. 1993. Cruising Guide to Southern California’s Offshore Islands. Santa Barbara, CA: Caractacus Corporation. Fages, P. 1937. A historical, political, and natural description of California (November 20, 1775), translated by H. E. Priestly. Berkeley: University of California Press. 1965. The Chumash Indians of Santa Barbara. In The California Indians: A Source Book, ed. R. F. Heizer and M. A. Whipple, 255–61. Berkeley: University of California Press. Fairbanks, R. G. 1989. A 17,000 year glacio-eustatic sea level record: Influence of glacial melting rates on the Younger Dryas event and deep ocean circulation. Nature 342: 637–41. Falconer, S. E., and S. H. Savage. 1995. Heartlands and hinterlands: Alternative trajectories of early urbanization in Mesopotamia and the southern Levant. American Antiquity 60(1): 37–58. Fausett, L. L. 1993. Activity and movement of the island fox, Urocyon littoralis, on Santa Cruz Island, California. In Third California Islands Symposium: Recent Advances in Research on the California Islands, ed. F. G. Hochberg, 391–404. Santa Barbara, CA: Santa Barbara Museum of Natural History. Fiedel, S. J. 1999. Older than we thought: Implications of corrected dates for Paleoindians. American Antiquity 64: 95–115. 2000. The peopling of the New World: Present evidence, new theories, and future directions. Journal of Archaeological Research 8: 39–103. Fitzgerald, R. T., and T. L. Jones. 2003. On the weight of the evidence from Cross Creek. American Antiquity 68(2): 396–99.

258

REFERENCES

Flannery, K. V. 1969. Origins and ecological effects of early domestication in Iran and the Near East. In The Domestication and Exploitation of Plants and Animals, ed. P. J. Ucko and G. W. Dimbleby, 73–100. London: Gerald Duckworth. Gamble, L. H. 1991. Organization of Activities at the Historic Settlement of Helo’: A Chumash Political, Economic, and Religious Center. Ph.D. dissertation, Department of Anthropology, University of California, Santa Barbara. 2002. Archaeological evidence for the origin of the plank canoe in North America. American Antiquity 67: 301–15. Gamble, L. H., and G. S. Russell. 2002. A view from the mainland: Late Holocene cultural developments among the Ventureño Chumash and the Tongva, In Catalysts to Complexity: Late Holocene Societies of the California Coast, ed. J. M. Erlandson and T. L. Jones, 101–26. Perspectives in California Archaeology 6. Los Angeles: Cotsen Institute of Archaeology, University of California. Gamble, L. H., P. L. Walker, and G. S. Russell. 2001. An integrative approach to mortuary analysis: Social and symbolic dimensions of Chumash burial practices. American Antiquity 66(2): 185–212. 2002. Further considerations on the emergence of Chumash chiefdoms. American Antiquity 67(4): 772–77. Geiger, M., and C. W. Meighan. 1976. As the Padres saw them: California Indian life and customs as reported by the Franciscan missionaries, 1813–1815. Santa Barbara Mission Archive Library. Gibbons, A. 2002. Humans’ head start: New views of brain evolution. Science 296: 835–37. Glassow, M. A. 1977. An Archaeological Overview of the Northern Channel Islands, California, including Santa Barbara Island. Salinas, CA: Coyote Press. 1980. Recent developments in the archaeology of the Channel Islands. In The California Islands: Proceedings of a Multidisciplinary Symposium, ed. D. M. Power, 79–99. Santa Barbara, CA: Santa Barbara Museum of Natural History. 1992. The relative dietary importance of marine foods through time in western Santa Barbara County. In Essays on the Prehistory of Maritime California, ed. Terry L. Jones, 115–28. Center for Archaeological Research at Davis. Publication no. 10. Davis: University of California Press. 1993a. Changes in subsistence on marine resources through 7,000 years of prehistory on Santa Cruz Island. In Archaeology on the Northern

REFERENCES

259

Channel Islands of California, ed. M. A. Glassow, 75–94. Salinas, CA: Coyote Press. 1993b. The occurrence of red abalone shells in Northern Channel Island archaeological middens: Implications for climatic reconstruction. In Third California Islands Symposium: Recent Advances in Research on the California Islands, ed. F. G. Hochberg, 567–78. Santa Barbara Museum of Natural History. 1996. Purisimeno Chumash Prehistory: Maritime Adaptations Along the Southern California Coast. Fort Worth, TX: Hartcourt Brace College. 1997. Middle Holocene cultural development in the central Santa Barbara Channel region. In Archaeology of the California Coast during the Middle Holocene, ed. J. M. Erlandson and M. A. Glassow, 73–90. Perspectives in California Archaeology 4. Los Angeles: Institute of Archaeology, University of California. 2000. Prehistoric chronology and environmental change at the Punta Arena site, Santa Cruz Island, California. In The Fifth California Islands Symposium, ed. D. R. Brown, K. C. Mitchell, and H. W. Chaney, 555–62. U.S. Department of the Interior Minerals Management Service, Pacific OCS Region. Glassow, M. A., D. J. Kennett, J. P. Kennett, and L. R. Wilcoxon. 1994. Confirmation of Middle Holocene ocean cooling inferred from stable isotopic analysis of prehistoric shells from Santa Cruz Island, California. In The Fourth California Islands Symposium: Update on the Status of Resources, ed. W. L. Halvorson and G. J. Maender, 223–32. Santa Barbara: Santa Barbara Museum of Natural History. Glassow, M. A., and L. R. Wilcoxon. 1988. Coastal adaptations near Point Conception, California, with particular regard to shellfish exploitation. American Antiquity 53: 36–51. Glassow, M. A., L. R. Wilcoxon, and J. Erlandson. 1988. Cultural and environmental change during the early period of Santa Barbara Channel prehistory. In The Archaeology of Prehistoric Coastlines, ed. G. Bailey and J. Parkington, 64–77. Cambridge: Cambridge University Press. Goldberg, C. F. 1993. The Application of Stable Carbon and Nitrogen Isotope Analysis to Human Dietary Reconstruction in Prehistoric Southern California. Ph.D. dissertation, University of California, Los Angeles. Gordon, J. H., and P. W. Weigand. 1997. Willows plutonic complex: A slice of immature arc rocks marooned on Santa Cruz Island, California. In Santa Cruz Island: Geology Field Trip Guide, ed. J. R. Boles and W. Landry, 59–71. San Diego, CA: San Diego Association of Geologists.

260

REFERENCES

Gray, R. S., and W. H. Harz. 1998. Pygmy mammoths from San Miguel Island, California. In Contributions to the Geology of the Northern Channel Islands, Southern California, ed. P. W. Weigand, 177–85. Bakersfield, CA: Pacific Section, American Association of Petroleum Geologists. Grayson, D. 1991. Alpine faunas from the White Mountains, California: Adaptive change in the Great Basin? Journal of Archaeological Science 18: 483–506. Greenwood, R. S. 1978. Archaeological Survey and Investigation, Channel Islands National Monument, California. Vols. 1 and 2. Prepared for the National Park Service, Contract Report. Gremillion, K. J. 1996. Diffusion and Adoption of Crops in Evolutionary Perspective. Journal of Anthropological Archaeology 15: 183–204. n.d. Central place foraging and food production on the Cumberland Plateau, Eastern Kentucky. In Behavioral Ecology and the Origins of Agriculture, ed. D. J. Kennett and B. Winterhalder. Berkeley, CA: University of California Press. Forthcoming. Haggett, P. 1971. Locational Analysis in Human Geography. New York: St. Martin’s. Hames, R. B., and W. T. Vickers. 1982. Optimal diet-breadth theory as a model to explain variability in Amazonian hunting. American Ethnologist 9: 358–78. Hansen, H. P. 1947. Postglacial forest succession, climate and chronology in the Pacific Northwest. Transactions of the American Philosophical Society 37: 1–130. Hawkes, K., and R. Bliege Bird. 2002. Showing off, handicap signaling, and the evolution of men’s work. Evolutionary Anthropology 11: 58–67. Hawkes, K., K. Hill, and J. F. O’Connell. 1982. Why hunter-gather: Optimal foraging and the Ache of eastern Paraguay. American Ethnologist 9: 378–97. Hawkes, K., J. O’Connell, and N. Blurton Jones. 1991. Hunting income patterns among the Hadza: Big game, common goods, foraging goals and the evolution of the human diet. In Foragers’ Strategies and Natural Diet of Monkeys, Apes, and Humans, ed. A. Whiten and E. Widdowson, 243–51, Proceedings of the Royal Society of London 334. Oxford: Clarendon Press.

REFERENCES

261

Heizer, R.F. 1955. California Indian linguistic records: The Mission Indian vocabularies of H. W. Henshaw. University of California Anthropological Records 15(2): 85–202. Heizer, R. F., and A. B. Elsasser, eds. 1956. Archaeological Investigations on Santa Rosa Island in 1901, by Philip Mills Jones. University of California Anthropological Records 17(2). Heizer, R. F., and M. A. Whipple, eds. 1971. The California Indians: A Source Book. 2nd ed. Berkeley: University of California Press. Hendy I. L., and J. P. Kennett. 1999. Latest Quaternary North Pacific surface-water responses imply atmosphere-driven climate instability. Geology 27(4): 291–94. 2000. Dansgaard-Oeschger cycles and the California Current system: Planktonic foraminiferal response to rapid climate change in Santa Barbara Basin, Ocean Drilling Program Hole 893A. Paleoceanography 15(1): 30–42. Heusser, C. J., L. E. Heusser, and D. M. Peteet. 1985. Late Quaternary climatic change on the American North Pacific Coast. Nature 315: 485–87. Heusser, L. E. 1978. Pollen in the Santa Barbara Basin, California: A 12,000 year record. Geological Society of America Bulletin 89: 673–78. Heusser, L. E., and F. Sirocko. 1997. Millennial pulsing of environmental change in southern California from the past 24 k.y.: A record of Indo-Pacific ENSO events? Geology 25(3): 243–46. Hickey, B. M. 1992. Circulation over the Santa Monica–San Pedro basin and shelf. Progressive Oceanography 30: 37–115. Hildebrandt, W. R., and T. L. Jones. 1992. Evolution of marine mammal hunting: A view from the California and Oregon coasts. Journal of Anthropological Archaeology 11: 360–401. Hildebrandt, W. R., and K. R. McGuire. 2002. Ascendance of hunting during the California middle Archaic: An evolutionary perspective. American Antiquity 67(2): 231–56. Hill, K., H. Kaplan, K. Hawkes, M. Hurtado. 1987. Foraging decisions among Ache hunter-gatherers: New data and implications for optimal foraging models. Ethology and Sociobiology 8: 1–36.

262

REFERENCES

Hockey, P. A. R., A. L. Bosman, and W. R. Siegfried. 1988. Patterns and correlates of shellfish exploitation by coastal people in Transkei: An enigma of protein production. Journal of Applied Ecology 25: 353–63. Hoover, R. L. 1971. Some Aspects of Santa Barbara Channel Prehistory. Ph.D. dissertation, Department of Anthropology, University of California, Berkeley. Huckins, J. 1995. Weather Data Summary. San Miguel, San Miguel, and Santa Cruz Islands: 1–23. Report on file at Channel Islands National Park, Ventura, California. Hudson, T., and T. C. Blackburn. 1982. The Material Culture of the Chumash Interaction Sphere. Vol. 1: Food Procurement and Transportation. Ballena Press Anthropological Papers, No. 25. Ballena Press and Santa Barbara Museum of Natural History Cooperative Publication, Los Altos and Santa Barbara, CA. Hudson, T., J. Timbrook, and M. Rempe, eds. 1978. Tomol: Chumash Watercraft as Described in the Ethnographic Notes of John P. Harrington. Anthropological Papers, No. 9. Socorro, NM: Ballena Press. Huyer, A. 1983. Coastal upwelling in the California Current system. Progress in Oceanography 12: 259–84. Ingram, B. L., and J. P. Kennett. 1995. Radiocarbon chronology and planktonic-benthic foraminiferal 14C age differences in Santa Barbara Basin sediments, Hole 893A. In Proceedings of the Ocean Drilling Program, Scientific Results, Vol. 146, ed. J. P. Kennett et al., 19–27. Ocean Drilling Program, College Station, TX. Ingram, B. L., and J. R. Southon. 1996. Reservoir ages in eastern Pacific coastal and estuarine waters. Radiocarbon 38(3): 573–82. Inman, D. 1983. Application of coastal dynamics to the reconstruction of paleo-coastlines in the vicinity of La Jolla, California. In Quaternary Coastlines and Marine Archaeology, ed. P. Masters and N. C. Flemming, 1–49. New York: Academic Press. Janetski, J. C. 1997. Fremont hunting and resource intensification in the eastern Great Basin. Journal of Anthropological Science 24: 1075–88. Jennings, C. W. 1977. Geologic Map of California. Division of Mines and Geology, Sacramento.

REFERENCES

263

Jochim, M. A. 1976. Hunter-Gatherer Subsistence and Settlement: A Predictive Model. New York: Academic Press. 1981. Strategies for Survival: Cultural Behavior in an Ecological Context. New York: Academic Press. 1988. Optimal foraging and the division of labor. American Anthropologist 90(1): 130–36. Johnson, A. W., and T. Earle. 1987. The Evolution of Human Societies from Foraging Group to Agrarian State. Stanford, CA: Stanford University Press. Johnson, D. L. 1972. Landscape Evolution on San Miguel Island, California. Ph.D. dissertation, University of Kansas. 1977. The Late Quaternary climate of coastal California: Evidence for an Ice Age refugium. Quaternary Research 8: 154–79. Johnson, G. A. 1977. Aspects of regional analysis in archaeology. Annual Review of Anthropology 6: 476–508. Johnson, J. R. 1982. An Ethnohistoric Study of the Island Chumash. Master’s thesis, Department of Anthropology, University of California, Santa Barbara. 1988. Chumash Social Organization: An Ethnohistoric Perspective. Ph.D. dissertation, Department of Anthropology, University of California, Santa Barbara. 1993. Cruzeño Chumash social geography. In Archaeology on the Northern Channel Islands of California, ed. M. A. Glassow, 19–46. Archives of California Prehistory 34. Salinas, CA: Coyote Press. 1999. Lineal descendants from the Northern Channel Islands. In Cultural Affiliation and Lineal Descent of Chumash Peoples in the Channel Islands and Santa Monica Mountains, ed. S. McLendon and J. R. Johnson, 185–261. Report prepared for the National Park Service, Washington, D.C. 2000. Social responses to climate change among the Chumash Indians of South-Central California. In The Way the Wind Blows: Climate, History, and Human Action, ed. R. J. McIntosh, J. A. Tainter, and S. K. McIntosh, 301–27. New York: Columbia University Press. 2001. Ethnohistoric reflections of Cruzeno Chumash society. In The Origins of a Pacific Coast Chiefdom: The Chumash of the Channel Islands, ed. J. E. Arnold, 53–70. Salt Lake City: University of Utah Press. Johnson, J. R., and S. McLendon. 2000. The social history of native islanders following missionization. In The Fifth California Islands Symposium, ed. D. R. Brown, K. C. Mitchell,

264

REFERENCES

and H. W. Chaney, 646–53. U.S. Department of the Interior Minerals Management Service, Pacific OCS Region. Johnson, J. R., T. W. Stafford, Jr., H. O. Ajie, and D. P. Morris. 2000. Arlington Springs revisited. In The Fifth California Islands Symposium, ed. D. R. Brown, K. C. Mitchell, and H. W. Chaney, 541–45. U.S. Department of the Interior Minerals Management Service, Pacific OCS Region. Johnson, J. R., and P. L. Walker. 1999. Effects of contact on the Chumash Indians. In Disease and Demography in the Americas, ed. J. Verano and D. Ubelaker, 127–39, Washington, D.C.: Smithsonian Institution Press. Jones, J. A., and D. Grice. 1993. A computer-generated soil map of Santa Cruz Island, California. In Third California Islands Symposium: Recent Advances in Research on the California Islands, ed. F. G. Hochberg, 45–56. Santa Barbara, CA: Santa Barbara Museum of Natural History. Jones, J. A., S. A. Junak, and R. J. Paul. 1993. Progress in mapping vegetation on Santa Cruz Island and a preliminary analysis of relationships with environmental factors. In Third California Islands Symposium: Recent Advances in Research on the California Islands, ed. F. G. Hochberg, 97–104. Santa Barbara, CA: Santa Barbara Museum of Natural History. Jones, K. T., and D. B. Madsen. 1989. Calculating the cost of resource transportation: A Great Basin example. Current Anthropology 30: 529–34. Jones, T. L., G. M. Brown, J. McVicar, M. Raab, D. J. Kennett, G. Spaulding, and A. York. 1999. Environmental imperatives reconsidered: Demographic crises in western North America during the medieval climatic anomaly. Current Anthropology 40(2): 137–70. Jones, T. L., R. T. Fitzgerald, D. J. Kennett, C. Misicek, J. Fagan, J. Sharp, and J. E. Erlandson. 2002. Stratigraphic evidence for a separate Late Pleistocene migration of coastal foragers into the New World. American Antiquity 67(2): 213–30. Jones, T. L., and W. R. Hildebrandt. 1995. Reasserting a prehistoric tragedy of the commons: Reply to Lyman. Journal of Anthropological Archaeology 14: 78–98. Jones, T. L., and D. J. Kennett. 1999. Late Holocene sea temperatures along the central California Coast. Quaternary Research 51: 74–82.

REFERENCES

265

Jones, T. L., and J. R. Richman. 1995. On mussels: Mytilus californianus as a prehistoric resource. North American Archaeologist 16(1) 33–58. Junak S., T. Ayers, R. Scott, D. Wilken, and D. Young. 1995. A Flora of Santa Cruz Island. Santa Barbara, CA: Santa Barbara Botanic Garden. Junger, S. T., and D. L. Johnson. 1980. Was there a Quaternary land bridge to the Northern Channel Islands? In The California Islands: Proceedings of a Multidisciplinary Symposium, ed. D. M. Power, 33–40. Santa Barbara, CA: Santa Barbara Natural History Museum. Kantner, J. 1996. Political competition among the Chaco Anasazi of the American Southwest. Journal of Anthropological Archaeology 15: 41–105. 1999a. Survival cannibalism or sociopolitical intimidation? Explaining perimortem mutilation in the American Southwest. Human Nature: An Interdisciplinary Biosocial Perspective 10(1): 1–50. 1999b. Anasazi mutilation and cannibalism in the American Southwest. In The Anthropology of Cannibalism, ed. L. R. Goldman, 75–104, Westport, CT: Bergin and Garvey. 2002. Complexity. In Darwin and Archaeology: A Handbook of Key Concepts, ed. J. P. Hart and J. E. Terrell, 89–106. Westport, CT: Bergin and Garvey. Kaplan, H., and K. Hill. 1992. The evolutionary ecology of food acquisition. In Evolutionary Ecology and Human Behavior, ed. E. A. Smith and B. Winterhalder, 167–202. New York: Aldine de Gruyter. Keefer, D. K., M. E. Moseley, and S. D. deFrance. 1998. Early Maritime Economy and el Nino Events at Quebrada Tacahauy, Peru. Science 281(5384): 1833–35. 2003. A 30,000-year record of floods and debris flows in the Ilo region of southern Peru and its relation to el Niño events and great earthquakes. Palaeogeography, Palaeoclimatology, Paleoecology 194: 41–77. Keegan, W. F., and J. M. Diamond. 1987. Colonization of islands by humans: A biogeographical perspective. Advances in Archaeological Method and Theory 10: 49–91. Keene, A. S. 1981. Optimal foraging in a non-marginal environment: A model of prehistoric subsistence strategies in Michigan. In Hunter-Gatherer Foraging Strategies, ed. E. A. Smith and B. Winterhalder, 171–93. Chicago: University of Chicago Press.

266

REFERENCES

Kelly, R. L. 1983. Hunter-gatherer mobility strategies. Journal of Anthropological Research 39: 277–306. 1995. The Foraging Spectrum: Diversity in Hunter-Gatherer Lifeways. Washington, D.C.: Smithsonian Institution Press. 1996. Foraging and fishing. In Prehistoric Hunter-Gatherer Fishing Strategies, ed. M. G. Plew, 208–14. Boise, ID: Boise State University. Kennett, D. J. 1998. Behavioral Ecology and the Evolution of Hunter-Gatherer Societies on the Northern Channel Islands, California. Ph.D. dissertation, Department of Anthropology, University of California, Santa Barbara. Kennett, D. J., and R. Clifford. 2004. Flexible strategies for resource defense on the Northern Channel Islands of California: An actor-based model. In Voyages of Discovery: The Archaeology of Islands, ed. S. Fitzpatrick. Westport, CT: Praeger/Greenwood Press. Kennett, D. J., and C. A. Conlee. 2002. Emergence of Late Holocene sociopolitical complexity on Santa Rosa and San Miguel Islands. In Catalist to Complexity: The Late Holocene Archaeology of the California Coast, ed. J. M. Erlandson and T. L. Jones, Perspectives in California and Archaeology 6, Los Angeles: University of California. Kennett, D. J., B. L. Ingram, J. Erlandson, and P. Walker. 1997. Evidence for temporal fluctuations of marine radiocarbon reservoir ages in the Santa Barbara Channel Region, California. Journal of Archaeological Science 24: 1051–59. Kennett, D. J., J. R. Johnson, T. C. Rick, D. P. Morris, and J. Christy. 2000. Historic Chumash settlement on eastern Santa Cruz Island, Southern California. Journal of California and Great Basin Anthropology 22(2): 212–22. Kennett, D. J., and J. P. Kennett. 2000. Competitive and cooperative responses to climatic instability in Southern California. American Antiquity 65(2): 379–95. Kennett, D. J., and B. Voorhies. 1996. Oxygen isotopic analysis of archaeological shells to detect seasonal use of wetlands on the southern Pacific coast of Mexico. Journal of Archaeological Science 23: 689–704. Kennett, D. J., A. Anderson, and B. Winterhalder. n.d.a. The ideal free distribution, food production, and the colonization of Oceania. In Behavioral Ecology and the Origins of Agriculture, ed. D. J. Kennett and B. Winterhalder. Berkeley, CA: University of California Press. Forthcoming.

REFERENCES

267

Kennett, D. J., B. Voorhies, and D. Martorana. n.d.b. An evolutionary model for the origins of agriculture on the Pacific coast of southern Mexico. In Behavioral Ecology and the Origins of Agriculture, ed. D. J. Kennett and B. Winterhalder. Berkeley, CA: University of California Press. Forthcoming. Kennett, D. J., and B. Winterhalder, ed. n.d. Behavioral Ecology and the Origins of Agriculture, Berkeley, CA: University of California Press. Forthcoming. Kennett, J. P., and B. L. Ingram. 1995a. Paleoclimatic evolution of Santa Barbara basin during the last 20 k.y.: Marine evidence from hole 893A. In Proceedings of the Ocean Drilling Program, Scientific Results, Vol. 146, ed. J. P. Kennett, 309–25. Ocean Drilling Program, College Station, TX. 1995b. A 20,000 year record of ocean circulation and climate change from the Santa Barbara basin. Nature 377: 510–14. Killingley, J. S. 1981. Seasonality of mollusk collecting determined from O-18 profiles of midden shells. American Antiquity 46: 152–58. King, C. D. 1976. Chumash inter-village economic exchange. In Native Californians: A Theoretical Perspective, ed. L. J. Bean and T. C. Blackburn, 289–318. Ramona, CA: Ballena Press. 1982. The evolution of Chumash society. Ph.D. dissertation, Department of Anthropology, University of California, Davis. 1990. Evolution of Chumash Society: A Comparative Study of Artifacts Used for Social System Maintenance in the Santa Barbara Channel Region before A.D. 1804. New York: Garland. King, L. 1969. The Medea Creek cemetery (LAN-243): An investigation of social organization from mortuary practices. UCLA Archaeological Survey Annual Report 11: 27–68. 1982. Medea Creek Cemetery: Late Inland Chumash Patterns of Social Organization, Exchange and Warfare. Ph.D. dissertation, Department of Anthropology, University of California, Los Angeles. Kirch, P. V. 2000. On the Road of the Winds: An Archaeological History of the Pacific Islands before European Contact. Berkeley: University of California Press. Klinger, R. C., and I. Messer. 2000. Composition and structure of a grassland community following prescribed burns on Santa Cruz Island. In The Fifth California Islands Symposium, ed. D. R. Brown, K. C. Mitchell, and H. W. Chaney, 190–200.

268

REFERENCES

U.S. Department of the Interior Minerals Management Service, Pacific OCS Region. Kowalewski, S. 1982. The evolution of primate regional systems. Comparative Urban Research 9(1): 60–68. Kroeber, A. L. 1925. Handbook of the Indians of California. Washington, D.C.: Bureau of American Ethnology Bulletin (No. 78). La Marche, V. C. 1973. Holocene climatic variations inferred from treeline fluctuations in the White Mountains, California. Quaternary Research 3: 632–60. 1974. Paleoclimatic inferences from long tree-ring records. Science 183: 1043–48. Lambert, P. M. 1994. War and Peace on the Western Front: A Study of Violent Conflict and Its Correlates in Prehistoric Hunter-Gatherer Societies of Coastal Southern California. Ph.D. dissertation, Department of Anthropology, University of California, Santa Barbara. 1997. Patterns of violence in prehistoric hunter-gatherer societies of coastal Southern California. In Troubled Times, ed. D. L. Martek and D. W. Frayer, 77–109. Amsterdam: Gordon and Breach. Lambert, P. M., and P. L. Walker. 1991. Physical anthropological evidence for the evolution of social complexity in coastal Southern California. Antiquity 65: 963–73. Landberg, L. C. W. 1965. The Chumash Indians of Southern California. Southwest Museum Papers 19. Los Angeles: Highland Park. 1975. Fishing effort in the aboriginal fisheries of the Santa Barbara region, California: An ethnohistorical appraisal. In Maritime Adaptations of the Pacific, ed. R. W. Casteel and G. I. Quimby, 145–70. Paris: Mouton. Larson, D., J. R. Johnson, and J. C. Michaelson. 1994. Missionizing among the coastal Chumash of central California: A study of risk minimization strategies. American Anthropologist 96: 263–99. Larson, D., and J. C. Michaelson. 1989. Climatic Variability: A Compounding Factor Causing Culture Change among Prehistoric Coastal Populations of Central California. Manuscript on file, Department of Anthropology, California State University, Long Beach. Le Boeuf, B. J., and M. L. Bonnel. 1980. Pinnipeds of the California islands: Abundance and distribution. In The California Islands: Proceedings of a Multidisciplinary Symposium, ed.

REFERENCES

269

D. M. Power, 475–96. Santa Barbara, CA: Santa Barbara Museum of Natural History. Lee, R. B. 1968. What hunters do for a living, or, how to make out on scarce resources. In Man the Hunter, ed. R. Lee and I. DeVore, 30–48. 1972. Population growth and the beginning of sedentary life among the !Kung bushman. In Population Growth, ed. B. Spooner, 329–42. Cambridge, MA: MIT Press. Lieth, H., and R. H. Whittaker. 1975. Primary Productivity of the Biosphere. New York: Springer. Lindström, S. 1990. Submerged tree stumps as indicators of mid-Holocene aridity in the Lake Tahoe Basin. Journal of California and Great Basin Anthropology 12(2): 146–57. 1996. Great Basin fisherfolk: Optimal diet-breadth modeling the Truckee River aboriginal subsistence fishery. In Prehistoric Hunter-Gatherer Fishing Strategies, ed. M. G. Plew, 114–79. Boise, ID: Boise State University. Littler, M. M. 1980. Overview of the rocky intertidal systems of Southern California. In The California Islands: Proceedings of a Multidisciplinary Symposium, ed. D. M. Power, 265–306. Santa Barbara, CA: Santa Barbara Museum of Natural History. Lock, G. R., and T. M. Harris. 1996. Danebury revisited: An English Iron Age hillfort in a digital landscape. In Anthropology, Space, and Geographic Information Systems, ed. M. Aldenderfer and H. D. G. Maschner, 214–40. Oxford: Oxford University Press. Love, M. 1996. Probably More Than You Want to Know about the Fishes of the Pacific Coast. Santa Barbara, CA: Really Big Press. Lyman, R. L. 1995. On the evolution of marine mammal hunting on the west coast of North America. Journal of Anthropological Archaeology 14: 45–77. Lyman, R. L., and M. J. O’Brien. 1998. The goals of evolutionary archaeology: History and explanation. Current Anthropology 39(5): 615–52. MacArthur, R. H. 1960. On the relation between reproductive value and optimal predation. Proceedings of the National Academy of Sciences 46: 143–45. MacArthur, R. H., and E. R. Pianka. 1966. On optimal use of a patchy environment. American Naturalist 100(916): 603–9.

270

REFERENCES

Madsen, D. B., and D. Rhode. 1990. Early Holocene pinyon (Pinus monophylla) in the northeastern Great Basin. Quaternary Research 33: 94–101. Madsen, D. B., and D. N. Schmitt. 1998. Mass collecting and the diet-breadth model: A Great Basin example. Journal of Anthropological Archaeology 25: 445–55. Mandryk, C. A. S., H. Josenshans, D. W. Fedje, and R. W. Mathewes. 2001. Late Quaternary paleoenvironments of northwestern North America: Implications for inland versus coastal migration routes. Quaternary Science Reviews 20: 301–14. Martin, S. L., and V. S. Popper. 2001. Paleoethnobotanical investigations of archaeological sites on Santa Cruz Island. In The Origins of a Pacific Coast Chiefdom: The Chumash of the Channel Islands, ed. J. E. Arnold, 245–59. Salt Lake City: University of Utah Press. Martz, P. C. 1984. Social Dimensions of Chumash Mortuary Populations in the Santa Monica Mountains Region. Ph.D. dissertation, Department of Anthropology, University of California, Riverside. 1992. Status distinctions reflected in Chumash mortuary populations in the Santa Monica Mountains region. In Essays on the Prehistory of Maritime California, ed. T. L. Jones, 145–56. Center for Archaeological Research at Davis Publication, No. 10. Davis: University of California Press. Maschner, H. D. G., ed. 1996. Darwinian Archaeologies. New York: Plenum Press. Maschner, H. D. G., and J. Q. Patton. 1996. Kin selection and the origins of hereditary social inequality: A case study from the northern Northwest coast. In Darwinian Archaeologies, ed. H. D. G. Maschner, 89–108. New York: Plenum Press. Masters, P. M., and D. R. Gallegos. 1997. Environmental change and coastal adaptations in San Diego County during the Middle Holocene. In Archaeology of the California Coast during the Middle Holocene, ed. J. M. Erlandson and M. A. Glassow, 11–22. Perspectives in California Archaeology 4. Los Angeles: Institute of Archaeology, University of California. McGuire, R. H. 1996. Why complexity is too simple. In Debating Complexity: Proceedings of the 26th Annual Chacmool Conference, ed. D. A. Meyer, P. C. Dawson, and D. T. Hanna, 23–29. Alberta, Canada: Archaeological Association of the University of Calgary. Meehan, B. 1977. Man does not live by calories alone: The role of shellfish in a coastal cuisine. In Sunda and Sahul: Prehistoric Studies in Southeast Asia, Melanesia,

REFERENCES

271

and Australia, ed. J. Allen, J. Golson, and R. Jones, 493–531. New York: Academic Press. 1982. Shell Bed to Shell Midden. Canberra, New South Wales: Australian Institute of Aboriginal Studies. Mehringer, P. J., Jr. 1985. Late Quaternary pollen records from the interior Pacific Northwest and northern Great Basin of the United States. In Pollen Records of Late-Quaternary North America Sediments, ed. V. M. Bryant, Jr., and R. G. Holloway, 167–89. Dallas, TX: American Association of Stratigraphic Palynologists. Meire, P., and E. Kuyken. 1984. Relations between the distribution of waders and the intertidal benthic fauna of the Oosterschelde, Netherlands. In Coastal Waders and Wildfowl in Winter, ed. P. R. Evans, J. D. Goss-Custard, and W. G. Hale, 57–68. Cambridge: Cambridge University Press. Melin, S. R., and R. L. DeLong. 2000. At-sea distribution and diving behavior of California sea lion females from San Miguel Island, California. In The Fifth California Islands Symposium, ed. D. R. Brown, K. C. Mitchell, and H. W. Chaney, 407–12. U.S. Department of the Interior Minerals Management Service, Pacific OCS Region. Meltzer, D. J., D. K. Grayson, G. Ardila, A. W. Barker, D. F. Dincauze, C. V. Haynes, F. Mena, L. Nunez, and D. J. Stanford. 1997. On the Pleistocene antiquity of Monte Verde, Southern Chile. American Antiquity 62: 659–63. Metcalfe, D., and K. R. Barlow. 1992. A model for exploring the optimal trade-off between field processing and transport. American Anthropologist 94: 340–56. Minnich, R. A. 1980. Vegetation of Santa Cruz and Santa Catalina Islands. In The California Islands: Proceedings of a Multidisciplinary Symposium, ed. D. M. Power, 123–37, Santa Barbara, CA: Santa Barbara Museum of Natural History. Mithen, S. 1989. Evolutionary theory and post-processual archaeology. Antiquity 63: 483–94. Molesworth, J. L., and J. R. Sloan. 1998. Foraminifera of the Eocene Canada Formation, Santa Cruz Island, Southern California. In Contributions to the Geology of the Northern Channel Islands, Southern California, ed. P. Weigand, 133–60. Bakersfield, CA: Pacific Section, American Association of Petroleum Geologists. Mooers, C. N., and A. R. Robinson. 1984. Turbulent jets and eddies in the California Current and inferred crossshore transports. Science 223: 51–53.

272

REFERENCES

Moratto, M. J. 1984. California Archaeology. New York: Academic Press. Morgan, A., L. S. Cummings, and J. L. Rudolph. 1991. Paleoenvironmental Change. In Resource Use and Settlement in the Santa Ynez River Valley, ed. C. F. Woodman, J. L. Rudolph, and T. P. Rudolph, 65–101. Santa Barbara, CA: Science Applications International Corporation. Morris, D., and J. M. Erlandson. 1993. 9,500-year-old human burial from CA-SRI-116, Santa Rosa Island. Journal of California and Great Basin Anthropology 15(1): 129–34. Moseley, M. E. 1975. The Maritime Foundations of Andean Civilization. Menlo Park, CA: Cummings. Moser, M. E. 1988. Limits to the numbers of grey plovers Pluvialis squatarola wintering on British estuaries: An analysis of long-term trends. Journal of Applied Ecology 25: 475–86. Moss, M. L. 1993. Shellfish, gender, and status on the Northwest Coast: Reconciling archeological, ethnographic, and ethnohistorical records of the Tlingit. American Anthropologist 95(3): 631–52. Munns, A. M., and J. E. Arnold. 2002. Late Holocene Santa Cruz Island: Patterns of continuity and change. In Catalysts to Complexity: Late Holocene Societies of the California Coast, ed. J. M. Erlandson and T. L. Jones, 127–46. Perspectives in California Archaeology 6. Los Angeles: Cotsen Institute of Archaeology, University of California. Murdock, G. P. 1967. Ethnographic Atlas. Pittsburgh, PA: University of Pittsburgh Press. Murray, S. N., M. M. Littler, and I. A. Abbott. 1980. Biogeography of the California marine algae with emphasis on the Southern California Islands. In Proceedings of the Symposium on the Biology of the California Islands, ed. R. N. Philbrick, 325–39. Santa Barbara, CA: Santa Barbara Botanic Garden. Nagaoka, L. 2002a. The effects of resource depression on foraging efficiency, diet breadth, and patch use in southern New Zealand. Journal of Anthropological Archaeology 21: 419–42. 2002b. Explaining subsistence change in southern New Zealand using foraging theory models. World Archaeology 34: 84–102.

REFERENCES

273

Neff, H. 2000. On evolutionary ecology and evolutionary archaeology: Some common ground? Current Anthropology 41(3): 427–29. Neiman, F. 1997. Conspicuous consumption as wasteful advertising: A Darwinian perspective on spatial patterns in Classic Maya terminal monument dates. In Rediscovering Darwin: Evolutionary Theory in Archaeological Explanation, ed. C. M. Barton and G. A. Clark, 267–90, Washington, D.C.: American Anthropological Association. Netting, R. McC. 1977. Cultural Ecology. Menlo Park, CA: Cummings. Nettle, D. 1997. On the status of methodological individualism. Current Anthropology 38(2): 283–85. Neushul, M., W. D. Clarke, and D. W. Brown. 1967. Subtidal plant and animal communities of the Southern California Islands. In Proceedings of the Symposium on the Biology of the California Islands, ed. R. N. Philbrick, 37–55. Santa Barbara, CA: Santa Barbara Botanic Garden. O’Brian, J. J. 1983. The physical environment in coastal upwelling regions. Progress in Oceanography 12: 221–57. O’Brien, M. J., and T. D. Holland. 1992. The role of adaptation in archaeological explanation. American Antiquity 57: 36–59. 1995. The nature and premise of a selection-based archaeology. In Evolutionary Archaeology: Methodological Issues, ed. P. A. Teltser, 175–200. Tucson: University of Arizona Press. O’Brien, M. J., and R. L. Lyman. 2002. Evolutionary archeology: Current status and future prospects. Evolutionary Anthropology 11(1): 26–36. O’Connell, J. F., and K. Hawkes. 1981. Alyawara plant use and optimal foraging theory. In Hunter-Gatherer Foraging Strategies: Ethnographic and Archaeological Analyses, ed. B. Winterhalder and E. A. Smith, 99–125. Chicago: University of Chicago Press. Odell, D. K. 1971. Census of pinnipeds breeding on the California Channel Islands. Journal of Mammals 52: 187–90. Odum, E. P. 1971. Fundamentals of Ecology. Philadelphia: Saunders.

274

REFERENCES

Olson, E. A., and W. S. Broecker. 1961. Lamont natural radiocarbon measurements VII. Radiocarbon 3: 141–75. Olson, R. L. 1928. Manuscript on file at the Santa Barbara Museum of Natural History. Ono, K. A., D. J. Boness, O. T. Oftedal, and S. J. Iverson. 1993. The effects of El Niño on mother-pup biology in the California Sea Lion. In Third California Islands Symposium: Recent Advances in Research on the California Islands, ed. F. G. Hochberg, 495–500. Santa Barbara: Santa Barbara Museum of Natural History. Orians, G. H., and N. E. Pearson. 1979. On the theory of central place foraging. In Analysis of Ecological Systems, ed. D. J. Horn, B. R. Stairs, and R. D. Mitchell, 155–77. Columbus: Ohio State University Press. Orr, P. C. 1951. Ancient population centers of Santa Rosa Island. American Antiquity 16: 221–26. 1956a. Radiocarbon, mammoths, and man on Santa Rosa Island. Geological Society of America Bulletin 67: 17–77. 1956b. Dwarf mammoth and man on Santa Rosa Island. University of Utah Anthropological Papers 26: 75–81. 1960. Radiocarbon dates from Santa Rosa Island, II. Santa Barbara Museum of Natural History, Department of Anthropology Bulletin 3: 1–10. 1962. The Arlington Spring site, Santa Rosa Island, California. American Antiquity 27(3): 417–19. 1967. Geochronology of Santa Rosa Island, California. In Proceedings of the Symposium on the Biology of the California Islands, ed. R. N. Philbrick, 317–25. Santa Barbara, CA: Santa Barbara Botanic Garden. 1968. Prehistory of Santa Rosa Island. Santa Barbara: Santa Barbara Museum of Natural History. Orr, P. C., and R. Berger. 1966. The fire areas on Santa Rosa Island, California. Proceedings of the National Academy of Sciences 56(5): 1409–16. Osborn, A. 1977. Strandloopers, mermaids, and other fairy tales: Ecological determinants of marine resource utilization—The Peruvian case. In For Theory Building in Archaeology, ed. L. R. Binford, 157–205. New York: Academic Press. Ostoja, S. M., and R. C. Klinger. 2000. The relationship of bishop pine cone morphology to serotiny on Santa Cruz Island, California. In The Fifth California Islands Symposium, ed.

REFERENCES

275

D. R. Brown, K. C. Mitchell, and H. W. Chaney, 167–71. U.S. Department of the Interior Minerals Management Service, Pacific OCS Region. Otterbein, K. F. 1968. Internal war: A cross-cultural study. American Anthropologist 70: 277–89. Pak, D., J. P. Kennett, and M. Kashgarian. 1997. Planktonic foraminiferal proxies of hydrographic conditions in the Santa Barbara Basin. EOS, Transactions, American Geophysical Union 78: 359. Palmer, D. F. 1997. Channel Islands National Park. In Geology of National Parks, ed. A. G. Harris, E. Tuttle, and S. D. Tuttle, 652–65. Dubuque, Iowa: Kendall/Hunt. Pálsson, G. 1988. Hunters and gatherers of the sea. In Hunters and Gatherers 1: History, Evolution and Social Change, ed. T. Ingold, D. Riches, and J. Woodburn, 189–204. New York: Berg. Patton, M. 1996. Islands in Time: Island Sociogeography and Mediterranean Prehistory. New York: Routledge. Pearlman, S. M. 1980. An optimum diet model, coastal variability and hunter-gatherer behavior. In Advances in Archaeological Method and Theory 3, ed. M. B. Schiffer, 257–310. New York: Academic Press. Pederson, P. O. 1938. Investigations into dental conditions of about 3,000 ancient and modern Greenlanders. The Dental Record 58: 191–98. Perry, J. E. 2003. Changes in Prehistoric Land and Resource Use among Complex Hunter-Gatherer-Fishers on Eastern Santa Cruz Island, California. Ph.D. dissertation, Department of Anthropology, University of California, Santa Barbara. Peterson, R. R. 1994. Archaeological settlement dynamics on the south side of Santa Cruz Island. In The Fourth California Islands Symposium: Update on the Status of Resources, ed. W. L. Halvorson and G. J. Maender, 215–22. Santa Barbara, CA: Santa Barbara Museum of Natural History. 2000. Settlement and subsistence at the middlelate transition on the south shore of Santa Cruz Island. In The Fifth California Islands Symposium, ed. D. R. Brown, K. C. Mitchell, and H. W. Chaney, 615–22. U.S. Department of the Interior Minerals Management Service, Pacific OCS Region.

276

REFERENCES

Philbrick, R. 1980. Distribution and evolution of endemic plants of the California Islands. In The California Islands: Proceedings of a Multidisciplinary Symposium, ed. D. M. Power, 173–88. Santa Barbara, CA: Santa Barbara Museum of Natural History. Philbrick, R., and J. Haller. 1977. The Southern Channel Islands. In Terrestrial Vegetation of California, ed. M. Barbour and J. Major, 893–906. New York: John Wiley and Sons. Pinter, N., C. C. Sorlien, and A. T. Scott. 1998. Late Quaternary folding and faulting of Santa Cruz Island, California. In Contributions to the Geology of the Northern Channel Islands, Southern California, ed. P. W. Weigand, 111–22. Bakersfield, CA: Pacific Section, American Association of Petroleum Geologists. Piperno, D. R. n.d. The origins of plant cultivation and domestication in the lowland Neotropics: A Behavioral Ecological perspective. In Behavioral Ecology and the Origins of Agriculture, ed. D. J. Kennett and B. Winterhalder. Berkeley, CA: University of California Press. Forthcoming. Piperno, D. R., and D. M. Pearsall. 1998. The Origins of Agriculture in the Lowland Neotropics, San Diego, CA: Academic Press. Pisias, N. 1978. Paleoceanography of the Santa Barbara basin during the last 8,000 years. Quaternary Research 10: 366–84. 1979. Model for paleoceanographic reconstructions of the California Current during the last 8,000 years. Quaternary Research 11: 373–86. Pletka, S. M. 2001a. Bifaces and the institutionalization of exchange relationships in the Chumash sphere. In The Origins of a Pacific Coast Chiefdom: The Chumash of the Channel Islands, ed. J. E. Arnold, 133–49. Salt Lake City: University of Utah Press. 2001b. The economics of island Chumash fishing practices. In The Origins of a Pacific Coast Chiefdom: The Chumash of the Channel Islands, ed. J. E. Arnold, 221–44. Salt Lake City: University of Utah Press. Porcasi, J. F., T. L. Jones, and L. M. Raab. 1999. Trans-Holocene marine mammal exploitation on San Clemente Island, California: A tragedy of the commons revisited. Journal of Anthropological Archaeology 19: 200–220. Porcasi, P., J. F. Porcasi, and C. O’Neill. 1999. Early Holocene coastlines of the California Bight: The Channel Islands as first visited by humans. Pacific Coast Archaeological Society Quarterly 35(2,3): 1–24.

REFERENCES

277

Power, D., ed. 1979. Natural Resources Study of the Channel Islands National Monument, California. Report to the National Park Service, Denver Service Center. Santa Barbara Museum of Natural History, Santa Barbara, California (contract no. cx-2000-8-0040). Preston, W. 1996. Serpent in Eden: Dispersal of foreign diseases into pre-mission California. Journal of California and Great Basin Anthropology 18: 2–37. 2002. Portents of plague from California’s Protohistoric. Ethnohistory 49: 69–121. Preziosi, A. M. 2001. Standardization and specialization: The island Chumash microdrill industry. In The Origins of a Pacific Coast Chiefdom: The Chumash of the Channel Islands, ed. J. E. Arnold, 151–63. Salt Lake City: University of Utah Press. Price, T. D., and J. A. Brown. 1985. Aspects of hunter-gatherer complexity. In Prehistoric Hunter-Gatherers: The Emergence of Cultural Complexity, ed. T. D. Price and J. A. Brown, New York: Academic Press. Raab, L. M. 1992. An optimal foraging analysis of prehistoric shellfish collecting on San Clemente Island, California. Journal of Ethnobiology 12(1): 63–80. 1994. The Dead at Calleguas Creek: A Study of Punctuated Cultural Evolution During the Middle–Late Period Transition in Southern California. Report for the Environmental Division, Naval Air Weapons Station, Point Mugu, California. Raab, L. M., and K. Bradford. 1997. Making nature answer to interpretivism: Response to J. E. Arnold, R. H. Colten, and S. Pletka. American Antiquity 62(2): 340–41. Raab, L. M., K. Bradford, J. F. Porcasi, and W. J. Howard. 1995. Return to Little Harbor, Santa Catalina Island, California: A critique of the marine paleotemperature model. American Antiquity 60(2): 287–308. Raab, L. M., and D. O. Larson. 1997. Medieval climatic anomaly and punctuated cultural evolution in coastal Southern California. American Antiquity 62(2): 319–36. Raab, L. M., and A. Yatsko. 1992. Ancient maritime adaptations of the California Bight: A perspective from San Clemente Island. In Essays on the Prehistory of Maritime California, ed. T. L. Jones, 173–94. Davis: Center for Archaeological Research, University of California.

278

REFERENCES

Ramage, C. S. 1986. El Niño. American Scientist 254(6): 77–83. Renfrew, C. 1973. Monuments, mobilization, and social organization in Neolithic Wessex. In The Explanation of Culture Change: Models in Prehistory, ed. C. Renfrew, 539–58. London: Duckworth. 1976. Megaliths, territories, and populations. Dissertationes Archaeologicae Gadenses 16: 198–220. Rick, T. C. 2001. AMS radiocarbon dating of a shell fishhook from Santa Rosa Island, California. Radiocarbon 43(1): 83–86. Rick, T. C., J. M. Erlandson, and R. L. Vellanoweth. 2001a. Paleocoastal marine fishing on the Pacific coast of the Americas: Perspectives from Daisy Cave, California. American Antiquity 66(4): 595–613. Rick, T. C., C. E. Skinner, J. M. Erlandson, and R. L. Vellanoweth. 2001b. Obsidian source characterization and human exchange systems on California’s Channel Islands. Pacific Coast Archaeological Society Quarterly 37(3): 1–18. Rick, T. C., R. L. Vellanoweth, J. M. Erlandson, and D. J. Kennett. 2002. On the antiquity of the single-piece shell fishhook: AMS radiocarbon evidence from the Southern California coast. Journal of Archaeological Science 29: 933–42. Rick, T. C., D. J. Kennett, and J. M. Erlandson. n.d. Archaeology and Paleoecology of the Abalone Rocks Estuary, Santa Rosa Island, California. In Proceedings of the Sixth California Islands Symposium. Santa Barbara, CA: Santa Barbara Museum of Natural History. Forthcoming. Rindos, D. 1984. The Origins of Agriculture: An Evolutionary Perspective. New York: Academic Press. Rogers, D. B. 1929. Prehistoric Man of the Santa Barbara Coast. Santa Barbara, CA: Santa Barbara Museum of Natural History. Roosevelt, A. C., M. Lima Da. Costa, C. Lopes Machado, M. Michado, M. Michab, N. Mercier, H. Vallada, J. Feathers, W. Barnet, M. Imazio da Silveira, A. Henderson, J. Sliva, B. Chernoff, D. S. Reese, J. A. Holman, N. Toth, and K. Schick. 1996. Paleoindian cave dwellers in the Amazon: The peopling of the Americas. Science 272: 373–84.

REFERENCES

279

Roth, V. L. 1993. Dwarfism and variability in the Santa Rosa Island mammoth (Mammuthus exilis): An interspecific comparison of limb-bone sizes and shapes in elephants. In Third California Islands Symposium: Recent Advances in Research on the California Islands, ed. F. G. Hochberg, 433–42. Santa Barbara, CA: Santa Barbara Museum of Natural History. Rozaire, C. E. 1978. Archaeological Investigations on San Miguel Island, California. Manuscript, Los Angeles, CA: Los Angeles County Museum of Natural History. 1983. Mortar and pestle manufacturing on San Miguel Island. Masterkey 57(4): 131–43. 1993. The bladelet industry on Anacapa and San Miguel Islands. In Archaeology on the Northern Channel Islands of California: Studies of Subsistence, Economics, and Social Organization, ed. M. A. Glassow, 63–73. Salinas, CA: Coyote Press. Russell, K. W. 1988. After Eden: The Behavioral Ecology of Early Food Production in the Near East and North Africa. Oxford: British Archaeological Reports, International Series 391. Rypins, S., S. L. Reneau, R. Byrne, and D. R. Montgomery. 1989. Palynologic and geomorphic evidence for environmental change during the Pleistocene–Holocene transition at Point Reyes Peninsula, Central Coastal California. Quaternary Research 32: 72–87. Salls, R. A. 1988. Prehistoric Fisheries of the California Bight. Ph.D. dissertation, Department of Anthropology, University of California, Los Angeles. Sandweiss, D. H. 2003. Terminal Pleistocene through mid-Holocene archaeological sites as paleoclimatic archives for the Peruvian coast. Paleogeography, Paleoclimatology, Paleoecology 194(1–3): 23–40. Sandweiss, D. H., K. A. Maasch, R. L. Burger, J. B. Richardson, H. B. Rollins, A. Clement. 2001. Variations in Holocene El Nino frequencies: Climate records and cultural consequences in ancient Peru. Geology 29(7): 603–6. Sandweiss, D. H., H. McInnis, R. L. Burger, A. Cano, B. Ojeda, R. Paredes, M. del Carmen Sandweiss, M. D. Glascock. 1998. Quebrada Jaquay: Early South American maritime adaptations. Science 281: 1830–32. Sandweiss, D. H., J. Richardson III, E. Reitz, J. Hsu, and R. Feldman. 1989. Early maritime adaptations in the Andes: Preliminary studies at the Ring Site. In Ecology, Settlement, and History in the Osmore Drainage,

REFERENCES

280

Peru, ed. D. Rice, C. Stanish, and P. Scarr, 35–84. British Archaeological Reports International Series No. 545, Oxford. Sandweiss, D. H., J. B. Richardson, E. J. Reitz, H. B. Rollins, and K. A. Maasch. 1996. Geoarchaeological evidence from Peru for a 5,000 years B.P. onset of El Nino. Science 273(5281): 1531–33. Schiffer, M. B. 1996. Some relationships between behavioral and evolutionary archaeologies. American Antiquity 61(4): 643–62. Schoenher, A. A., C. R. Feldmeth, and M. J. Emerson. 1999. Natural History of the Islands of California. Berkeley: University of California Press. Schumacher, P. 1875. Ancient graves and shell heaps of California. Smithsonian Institution Annual Report 1874: 353–55. Seapy, R. R., and M. M. Littler. 1980. Biogeography of the rocky intertidal macroinvertebrates of the Southern California islands. In The California Islands: Proceedings of a Multidisciplinary Symposium, ed. D. M. Power, 307–23. Santa Barbara, CA: Santa Barbara Museum of Natural History. Shapiro, R. S. 1998. Paleogene–Early Neogene macrofossils of southwestern Santa Cruz Island. In Contributions to the Geology of the Northern Channel Islands, Southern California, ed. Peter Weigand, 123–32. Bakersfield, CA: Pacific Section, American Association of Petroleum Geologists. Sharp, J. 2000. Shellfish analysis from a Santa Cruz Island red abalone midden: Reevaluating the marine cooling hypothesis. In The Fifth California Islands Symposium, ed. D. R. Brown, K. C. Mitchell, and H. W. Chaney, 563–72. U.S. Department of the Interior Minerals Management Service, Pacific OCS Region. Shennan, S. 1989. Cultural transmission and cultural change. In What’s New? A Closer Look at the Process of Innovation, ed. S. E. van der Leeuw and R. Torrence, 330–46. London: Unwin Hyman. 2002. Genes, Memes and Human History: Darwinian Archaeology and Cultural Transmission. London: Thames and Hudson. Sidwell, V. D. 1981. Chemical and Nutritional Composition of Finfishes, Whales, Crustaceans, Mollusks, and Their Products. NOAA Technical Memorandum NMFS FSEC-11. Washington, D.C.: U.S. Department of Commerce.

REFERENCES

281

Simms, S. R. 1987. Behavioral Ecology and Hunter-Gatherer Foraging: An Example from the Great Basin. British Archaeological Reports, International Series 381. Oxford: B.A.R. Simpson, L. B., ed. 1939. California in 1792: The Expedition of Longinos Martinez. Translated by L. B. Simpson, San Marino, CA: Huntington Library Publications. Sjögren, K.-G. 1996. Kinship, labor, and land in Neolithic southwest Sweden: Social aspects of megalithic graves. Journal of Anthropological Archaeology 5(3): 229–65. Smith, B. D. 2001. Low-level food production. Journal of Archaeological Research 9(1): 1–43. Smith, C. 1952. A Flora of Santa Barbara. Santa Barbara, CA: Santa Barbara Botanic Garden. Smith, C. A. 1976. Regional economic systems linking geographical models and socioeconomic problems. In Regional Economic Systems, Vol. 1, ed. C. Smith, 3–68. New York: Academic Press. Smith, E. A. 1991. Inujjuamiut Foraging Strategies: Evolutionary Ecology of an Arctic Hunting Economy. New York: Aldine de Gruyter. 2000. Three styles in the evolutionary study of human behavior. In Human Behavior and Adaptation: An Anthropological Perspective, ed. L. Cronk, W. Irons, and N. Chagnon, 27–46. New York: Aldine de Gruyter. Smith, E. A., and R. Bliege Bird. 2000. Turtle hunting at Tombstone Opening: Public generosity as costly signaling. Evolution of Human Behavior 21: 245–61. Smith, E. A., R. Bliege Bird, and D. W. Bird. 2003. The benefits of costly signaling: Meriam turtle hunters and spearfishers. Behavioral Ecology 14: 116–26. Smith, E. A., and B. Winterhalder. 1981. New perspectives on hunter-gatherer socioecology. In Hunter-Gatherer Foraging Strategies: Ethnographic and Archaeological Analyses, ed. B. Winterhalder and E. A. Smith, 1–12. Chicago, IL.: University of Chicago Press. 1992. Natural selection and decision-making: Some fundamental principles. In Evolutionary Ecology and Human Behavior, ed. E. A. Smith and B. Winterhalder, 25–60. New York: Aldine de Gruyter. Smith, E. A., and M. Wishnie. 2000. Conservation and subsistence in small-scale societies. Annual Review of Anthropology 29: 493–524.

282

REFERENCES

Snethkamp, P., and D. Guthrie. 1988. The Early Holocene Occupations of Daisy Cave, San Miguel Island, California. Paper presented, Society for California Archaeology Annual Meeting. Sosis, R. 2000. Costly signaling and torch fishing on Ifaluk Atoll. Evolution and Human Behavior 21: 223–44. 2001. Sharing consumption and patch choice on Ifaluk Atoll: Evaluating an explanatory hypothesis. Human Nature 22(3): 221–45. 2002. Patch choice decisions among Ifaluk fishers. American Anthropologist 104(2): 583–98. Sorlien, C. C., and N. Pinter. 1997. Faulting and folding of Santa Cruz Island, California. In Santa Cruz Island: Geology Field Trip Guide, ed. J. R. Boles and W. Landry, 72–90. San Diego, CA: San Diego Association of Geologists. Speth, J., and K. Spielman. 1983. Energy source, protein metabolism, and hunter-gatherer subsistence strategies. Journal of Anthropological Archaeology 2: 1–31. Spriggs, M. 1997. The Island Melanesians. Oxford: Blackwell. Steadman, D. W. 1989. Extinction of birds in Eastern Polynesia: A review of the record, and comparisons with other Pacific Island groups. Journal of Archaeological Science 16: 177–205. Steadman, D. W., and P. V. Kirch. 1990. Prehistoric extinction of birds on Mangaia, Cook Islands, Polynesia. Proceedings of the National Academy of Sciences 87: 9605–9. Steadman, D. W., D. S. Pahlavan, and P. V. Kirch. 1990. Extinction, biogeography, and human exploitation of birds on Tikopia and Anuta, Polynesian Outliers in the Solomon Islands. B.P. Bishop Museum Occasional Papers 30: 118–53. Steadman, D. W., and B. Rolett. 1996. Chronostratigraphic analysis of landbird extinction on Tahuata, Marquesas Islands. Journal of Archaeological Science 23(1): 81–94. Stephens, D. W., and J. R. Krebs. 1986. Foraging Theory. Princeton, N J: Princeton University of Press. Steward, J. 1955. Theory of Culture Change. Urbana: University of Illinois Press.

REFERENCES

283

Stewart, B. S., P. K. Yochem, R. L. DeLong, and G. A. Antonelis. 1993. Trends in abundance and status of pinnipeds on the Southern California Channel Islands. In Third California Islands Symposium: Recent Advances in Research on the California Islands, ed. F. G. Hochberg, 501–18. Santa Barbara, CA: Santa Barbara Museum of Natural History. Stine, S. 1994. Extreme and persistent drought in California and Patagonia during medieval time. Nature 369: 546–49. Stuiver, M., and T. F. Braziunas. 1993. Modeling atmospheric 14C influences and 14C ranges of marine samples to 10,000 BC. Radiocarbon 35: 137–89. Stuiver, M., and P. J. Reimer. 1993. Extended 14C database and revised Calib 3.0 program. Radiocarbon 35(1): 215–30.

14

C age calibration

Sutherland, W. J. 1996. From Individual Behavior to Population Ecology. Oxford: Oxford University Press. Tainter, J. A. 1996. Valuing complexity. In Debating Complexity: Proceedings of the 26th Annual Chacmool Conference, ed. D. A. Meyer, P. C. Dawson, and D. T. Hanna, 10–15. Alberta, Canada: Archaeological Association of the University of Calgary. Tegner, M. J., and P. K. Dayton. 1987. El Niño effects on Southern California kelp forest communities. Advances in Ecological Research 17: 243–79. 1991. Sea Urchins, El Niños and the long-term stability of Southern California kelp forest communities. Marine Ecology Progress Series 77: 49–63. Teltser, P. A. 1995a. The methodological challenge of evolutionary theory in archaeology. In Evolutionary Archaeology: Methodological Issues, ed. P. A. Teltser, 1–11. Tucson: University of Arizona Press. 1995b. Culture history, evolutionary theory, and frequency seriation. In Evolutionary Archaeology: Methodological Issues, ed. P. A. Teltser, 51–68. Tucson: University of Arizona Press. Thaler, P. F. 1998. Mammoth remains on the Channel Islands of California: A summary of geologic, biogeographic, and evolutionary implications. Contributions to the Geology of the Northern Channel Islands, Southern California, ed. P. W. Weigand. 161–85. Bakersfield, CA: Pacific Section, American Association of Petroleum Geologists.

284

REFERENCES

Thomas, F. R. 2001. Remodeling marine tenure on the atolls: A case study from western Kiribati, Micronesia. Human Ecology 29(4): 399–423. Thompson, R. S. 1992. Late Quaternary environments in Ruby Valley, Nevada. Quaternary Research 37: 1–15. Thompson, R. S., C. Whitlock, P. J. Bartlein, S. P. Harrison, and W. G. Spaulding. 1993. Climatic changes in the western United States since 18,000 years B.P. In Global Climates since the Last Glacial Maximum, ed. H. E. Wright, Jr. et al., 468–513. Minneapolis: University of Minnesota Press. Timbrook, J. 1980. A wooden artifact from Santa Cruz Island. Journal of California and Great Basin Anthropology 2: 272–79. 1982a. Use of wild cherry pits as food by the California Indians. Journal of Ethnobiology 2(2): 162–76. 1982b. Vegetation of California and Great Basin anthropology. Journal of California and Great Basin Anthropology 4(2): 163–86. 1986. Chia and the Chumash: A reconsideration of sage seeds in Southern California. Journal of California and Great Basin Anthropology 8(1): 50–64. 1987. Virtuous herbs: Plants in Chumash medicine. Journal of Ethnobiology 7(2): 171–80. 1990. Ethnobotany of Chumash Indians, California, based on collections by John P. Harrington. Economic Botany 44(2): 236–53. 1993. Island Chumash ethnobotany. In Archaeology on the Northern Channel Islands of California: Studies of Subsistence, Economics and Social Organization, ed. M. A. Glassow, 47–62. Salinas, CA: Coyote Press. Timbrook, J., J. R. Johnson, and D. D. Earle. 1982. Vegetation burning by the Chumash. Journal of California and Great Basin Anthropology 4(2): 163–86. Trigger, B. G. 1989. A History of Archaeological Thought. Cambridge: Cambridge University Press. Trivers, R. L. 1971. The evolution of reciprocal altruism. Quarterly Review of Biology 46: 35–57. True, D. L. 1990. Site location and water supply: A perspective from northern San Diego County. Journal of New World Archaeology 7(4): 37–60.

REFERENCES

285

Tucker, B. T. 2001. The Behavioral Ecology and Economics of Variation, Risk and Diversification among Mikea Forager-Farmers of Madagascar. Ph.D. dissertation, University North Carolina at Chapel Hill. Turner, C. G. 2003. Three ounces of sea shells and one fish bone do not a coastal migration make. American Antiquity 68(2): 391–95. Ugan, A., and J. Bright. 2001. Measuring foraging efficiency with archaeological faunas: The relationship between relative abundance indices and foraging returns. Journal of Archaeological Science 28: 1309–21. Van Geen, A., S. N. Luoma, C. C. Fuller, R. Anima, H. E. Clifton, S. Trumbore. 1992. Evidence from Cd/Ca ratios in foraminifera for greater upwelling off California 4,000 years ago. Nature 358: 54–56. Vellanoweth, R. L. 1998. Earliest island fox remains on the southern Channel Islands: Evidence from San Nicolas Island, California. Journal of California Great Basin Anthropology 20(1): 100–8. Vellanoweth, R. L., T. C. Rick, and J. M. Erlandson. 2000. Middle and Late Holocene maritime adaptations on northeastern San Miguel Island, California. In The Fifth California Islands Symposium, ed. D. R. Brown, K. C. Mitchell, and H. W. Chaney, 607–614. U.S. Department of the Interior Minerals Management Service, Pacific OCS Region. Voorhies, B., D. J. Kennett, J. G. Jones, and T. A. Wake. 2002. A Middle Archaic archaeological site on the West Coast of Mexico. Latin American Antiquity 13(2): 179–200. Wagner, H. R. 1929. Spanish voyages to the Northwest Coast of North America in the sixteenth century. California Historical Society Special Publications 4. San Francisco. Wake, T. A. 2001. Bone tool technology on Santa Cruz Island and implications for exchange. In The Origins of a Pacific Coast Chiefdom: The Chumash of the Channel Islands, ed. J. E. Arnold, 183–98. Salt Lake City: University of Utah Press. Walker, P. L. 1978. A quantitative analysis of dental attrition rates in the Santa Barbara Channel area. American Journal of Physical Anthropology 48: 101–6. 1986. Porotic hyperostosis in a marine-dependent California Indian population. American Journal of Physical Anthropology 69: 345–54.

286

REFERENCES

1989. Cranial injuries as evidence of violence in prehistoric Southern California. American Journal of Physical Anthropology 80: 313–23. Walker, P. L., and M. J. DeNiro. 1986. Stable nitrogen and carbon isotope ratios in bone collagen as indices of prehistoric dietary dependence on marine and terrestrial resources in Southern California. American Journal of Physical Anthropology 71: 51–61. Walker, P. L., and J. Erlandson. 1986. Dental evidence for prehistoric dietary change on the Northern Channel Islands. American Antiquity 51(2): 375–83. Walker, P. L., and S. Hollimon. 1989. Changes in osteoarthritis associated with the development of maritime economy among Southern California Indians. International Journal of Anthropology 4(1): 171–83. Walker, P. L. and T. Hudson 1993 Chumash Healing: Changing Health and Medical Practices in an American Indian Society. Banning, CA: Malki Museum Press. Walker, P. L., D. J. Kennett, T. Jones, and R. Delong. 2000. Archaeological investigations of the Point Bennett pinniped rookery on San Miguel Island. In The Fifth California Islands Symposium, ed. D. R. Brown, K. C. Mitchell, and H. W. Chaney, 628–32. U.S. Department of the Interior Minerals Management Service, Pacific OCS Region. Walker, P. L., and P. E. Snethkamp. 1984. Archaeological investigations on San Miguel Island—1982, prehistoric adaptations to the marine environment, Vol. 1, manuscript on file, Central Coast Information Center, University of California, Santa Barbara. Wallace, W. J. 1955. A suggested chronology for Southern California coastal archaeology. Southwestern Journal of Anthropology 11(3): 214–30. Walter, H. S., and L. A. Taha. 2000. Regeneration of bishop pine (Pinus muricata) in the absence and presence of fire: A case study from Santa Cruz Island, California. In The Fifth California Islands Symposium, ed. D. R. Brown, K. C. Mitchell, and H. W. Chaney, 172–81. U.S. Department of the Interior Minerals Management Service, Pacific OCS Region. Watt, B. K., and A. L. Merrill. 1975. Composition of Foods, Agriculture Handbook 8, Washington, D.C.: United States Department of Agriculture.

REFERENCES

287

Weaver, D. W. 1969. General geology of the region. In Geology of the Northern Channel Islands: AAPG and SEPM Pacific Section, ed. D. W. Weaver, D. P. Doerner, and B. Nolf, 9–10. Los Angeles, CA: Pacific Section, AAPG. Weaver, D. W., and D. P. Doerner. 1969. Lower Tertiary stratigraphy, San Miguel and Santa Rosa islands. In Geology of the Northern Channel Islands: AAPG and SEPM Pacific Section, ed. D. W. Weaver, D. P. Doerner, and B. Nolf, 30–47. Los Angeles, CA: Pacific Section, AAPG. Weaver, D. W., and G. L. Meyer. 1969. Stratigraphy of northeastern Santa Cruz Island. In Geology of the Northern Channel Islands: AAPG and SEPM Pacific Section, ed. D. W. Weaver, D. P. Doerner, and B. Nolf, 95–104. Los Angeles, CA: Pacific Section, AAPG. Weigand, P. W. 1993. Geochemistry and origin of middle Miocene volcanic rocks from Santa Cruz and Anacapa islands, Southern California borderland. In Third California Islands Symposium: Recent Advances in Research on the California Islands, ed. F. G. Hochberg, 21–37. Santa Barbara, CA: Santa Barbara Museum of Natural History. 1997. Volcanic rocks. In Santa Cruz Island: Geology Field Trip Guide, ed. J. R. Boles and W. Landry, 44–49. San Diego, CA: San Diego Association of Geologists. ed. 1998. Contributions to the Geology of the Northern Channel Islands, Southern California. Bakersfield, CA: Pacific Section, American Association of Petroleum Geologists. Weigand, P. W., and K. L. Savage. 2000. Summary of the miocene igneous rocks of the Channel Islands, southern California. In The Fifth California Islands Symposium, edited by D. R. Brown, K. C. Mitchell, and H. W. Chaney, 106–114. U.S. Department of the Interior Minerals Management Service, Pacific OCS Region. Wenner, A. M., and D. L. Johnson. 1980. Land vertebrates on the California Channel Islands: Sweepstakes or bridges? In California Islands: Proceedings of a Multidisciplinary Symposium, ed. D. M. Power, 497–530. Santa Barbara, CA: Santa Barbara Museum of Natural History. West, G. J., and J. M. Erlandson. 1994. A Late Pleistocene pollen record from San Miguel Island, California: Preliminary results. American Quaternary Association Program and Abstracts. 13th biennial meeting, Minneapolis, MN: 156.

288

REFERENCES

White, T. D. 1991. Human Osteology. New York: Academic Press. Whitlock, C. 1992. Vegetational and climatic history of the Pacific Northwest during the last 20,000 years: Implications for understanding present day biodiversity. The Northwest Environmental Journal 8: 5–28. Whitlock, C., and P. J. Bartlein. 1997. Vegetation and climate change in northwest America during the past 125 kyr. Letters to Nature 388: 57–61. Wickham, J. B. 1975. Observations of the California countercurrent. Journal of Marine Research 33(3): 325–40. Wilcoxon, L. R. 1993. Subsistence and site structure: An approach for deriving cultural information from coastal shell middens. In Archaeology on the Northern Channel Islands of California, ed. M. A. Glassow, 137–50. Salinas, CA: Coyote Press. Winterhalder, B., and C. Goland. 1993. On population, foraging efficiency, and plant domestication. Current Anthropology 34(5): 710–15. 1997. An evolutionary ecology perspective on diet choice, risk, and plant domestication. In People, Plants, and Landscapes: Studies in Paleoethnobotany, ed. K. J. Gremillion, 123–60. Tuscaloosa: University of Alabama Press. Winterhalder, B., and F. Lu. 1997. A forager-resource population ecology model and implications for indigenous conservation. Conservation Biology 11(6): 1354–62. Winterhalder, B., and E. A. Smith, eds. 1981. Hunter-Gatherer Foraging Strategies: Ethnographic and Archaeological Analyses. Chicago: University of Chicago Press. 1992. Evolutionary ecology and the social sciences. In Evolutionary Ecology and Human Behavior, ed. E. A. Smith and B. Winterhalder, 3–24. New York: Aldine de Gruyter. 2000. Analyzing adaptive strategies: Human behavioral ecology at twenty-five. Evolutionary Anthropology 9(2): 51–72. Winterhalder, B., F. Lu, and B. Tucker. 1999. Risk-sensitive adaptive tactics: Models and evidence from subsistence studies in biology and anthropology. Journal of Archaeological Research 7(4): 301–48. Winterhalder, B., and D. J. Kennett. n.d. Human behavioral ecology and the origins of agriculture: Volume introduction. In Behavioral Ecology and the Origins of Agriculture, ed.

REFERENCES

289

D. J. Kennett and B. Winterhalder. Berkeley, CA: University of California Press. Forthcoming. Winterhalder, B. P. 1977. Foraging strategies and adaptations of the Boreal Forest Cree: An evaluation of theory and models from evolutionary ecology. Ph.D. dissertation, Cornell University. 1981a. Optimal foraging strategies and hunter-gatherer research in anthropology: Theory and models. In Hunter-Gatherer Foraging Strategies: Ethnographic and Archaeological Analyses, ed. B. Winterhalder and E. A. Smith, 13–35. Chicago: University of Chicago Press. 1981b. Foraging strategies in the boreal forest: An analysis of Cree hunting and gathering. In Hunter-Gatherer Foraging Strategies: Ethnographic and Archaeological Analyses, ed. B. Winterhalder and E. A. Smith, 66–98. Chicago: University of Chicago Press. 1986. Diet choice, risk, and food sharing in a stochastic environment. Journal of Anthropological Archaeology 5: 369–92. 1994. Concepts in historical ecology: The view from evolutionary theory. In Historical Ecology: Cultural Knowledge and Changing Landscapes, ed. C. L. Crumley, 17–42. Santa Fe, NM: School of American Research Press. 1997. Gifts given, gifts taken: The behavioral ecology of nonmarket, intragroup exchange. Journal of Archaeological Research 5(2): 121–68. Wobst, H. M. 1978. The archaeo-ethnology of hunter-gatherers or the tyranny of the ethnographic record in archaeology. American Archaeology 43(2): 303–9. Woolley, J. J. 1998. Aspects of the Quaternary geology of Santa Rosa Island, California. In Contributions to the Geology of the Northern Channel Islands, Southern California, ed. P. W. Weigand, 103–10. Bakersfield, CA: Pacific Section, American Association of Petroleum Geologists. Yatsko, A. 2000. Late Holocene paleoclimatic stress and prehistoric human occupation on San Clemente Island. Ph.D. dissertation, Department of Anthropology, University of California, Los Angeles. Yellen, J. E. 1998. Barbed bone points: Tradition and continuity in Saharan and subSaharan Africa. African Archaeological Review 15: 173–98. Yellen, J. E., A. S. Brooks, E. Cornelissen, M. J. Mehlman, and K. Stewart. 1995. A Middle Stone Age worked bone industry from Katanda, Upper Semliki Valley, Zaire. Science 268: 553–56.

290

REFERENCES

Yesner, D. R. 1980. Maritime hunter-gatherers: Ecology and prehistory. Current Anthropology 21: 727–50. 1987. Life in the “Garden of Eden”: Constraints of marine diets for human societies. In Food and Evolution, ed. M. Harris and E. G. Ross, 285–310. Philadelphia, PA: Temple University Press. 2001. Human dispersal into interior Alaska: Antecedent conditions, mode of colonization, and adaptations. Quaternary Science Reviews 20: 315–27. York, A. L. 1996. An Archaeological Survey of Jolla Vieja Canyon, Santa Rosa Island, California. Report, Ventura, CA: Channel Islands National Park. Zeanah, D. W. 2000. Transport costs, central-place foraging, and hunter-gatherer alpine land-use strategies. In Intermountain Archaeology: Selected Papers of the Rock Mountain Anthropological Conference, ed. D. B. Madsen and M. D. Metcalfe, 1–14. Salt Lake City: University of Utah Press. Zipf, G. K. 1965. Human Behavior and the Principle of Least Effort: An Introduction to Human Ecology. New York: Hafner.

Index

Abalone black, 58 red, 58 species significance, 188 Aggression, cultural development and, 215 Aggressive competition resources, 235 territoriality, 236 Altithermal conditions, in California environments, 70 Animal resources, postencounter return rates and, 19 Aquatic resources, 237 Arlington Canyon, 123 villages, 139 Arnold, Jeanne, 97 Artifacts Late Holocene, 158–163, 165–168, 175 Santa Rosa Island, 142 settlement permanence, 139 Beads glass, historic period and, 102 manufacturing, 202, 203 production increase, 206 Olivella shell, 202, 212 bead production, 199 distribution, 203 intensity, 203 radiocarbon dates, 204–205 Bee Rock, Late Holocene, 102

Behavioral ecologists 1970’s foraging models, 11 fish exploitation, 24 Behavioral variability, HBE, 3 Bone densities, exponential increases in fish, 219 Bowers, Steven, 99 Cabrillo, Juan Rodríguez, 4 California. See also Southern California coast archaeological record, 37–38 sea level, 67 environments, altithermal conditions and, 70 mussels, 58, 144, 179, 221 sea lions, 196 White Mountains, 71 California Current, marine productivity and, 56 Cañada Verde, 100 Carbon isotopic analysis, 148–149 Carbonized plant material, 118 Cemeteries, placement of, 231 Central place foraging (CPF) theory, 30, 224, 225 and foraging, 31–32 predictions, 228 Chalcedonic chert (Cico), 200 Channel Islands. See also Northern Channel Islands

291

292

INDEX

Channel Islands (continued) adaptive viability, 5 book synthesis subjects, 5–6 chronological framework, 9 evolutionary process, 5 Spanish explorers, 4 Channel Islands National Park, 5 Chert bifaces, Late Holocene, 176 Chert quarries, 209 Cheumen, 98 Chiefdoms, in Chumash society, 109 Chumash communities convex rank-size distribution, 109f distribution of, 74 division of labor, 77 history, 111 market economy, 86 primate rank-size distribution, 109f rank-size distribution, 110 seasonal settlement shifts, 76 dialects, 185 hierarchy, 77 island society, 152 occupation, history of, 91, 97 population cultural geography, 180 levels, 73–75 Old World disease, 157 society, 187 evolutionary history, 4 exchange, 78 fishing, 192 marine traditions, 152 political leadership, 211 social complexity, 92 specialization, 78 Cico chert, 200 Climatic change, and cultural complexity, 6–9 Coast California archaeological record, 37–38 sea level, 67 population ethnohistoric summary, 80 Southern California major current systems, 55 political complexity, 6 Coastal habitats, 221 maritime foragers, 15 resource distribution, 29

Coastal settings diet-breadth model, 18–19 population density and group sizes, 36–37 social hierarchies competition and formation, 36–39 Coastal settlement Late Holocene, 169 Middle Holocene, 170 Community-based craft specialization, 212 Communities, formation of HBE, 37–38 Competition, and population growth, 234 Competitive aggression, and economically defendable resources, 39 Convex distribution, 108 CPF. See Central place foraging (CPF) theory Cultural context, 72–90 Cultural ecological principles vs. HBE, 12 Daisy Cave, faunal assemblage, 122 Darwinian approach, HBE, 12 Demographic expansion, and population growth, 229 Deposits ephemeral, 129 Pleistocene shell, 119 residential, 137 Despotic behavior, 185 Dialects, Chumash, 185 Dichclostemma, 54 Diet breadth model, 217–219 application, 16 choice, 220 coastal settings, 18–19 prey choice predictor, 20, 27 Dietary differences, 192 Middle Holocene diet, 227 Dietary reconstructions, 122 Diversification, rank size analysis, 108 Early Holocene, 82, 118 burial sites, 127 climatic conditions, 227 cultural developments, 154 habitats, 221 King’s Chronology, 84 Lobo canyon, 137 record, 122 settlement, 128

INDEX

shellfish, 124 site positioning, 126 Ecological context, HBE, 217 Ecological interaction, HBE, humans and environment, 12 Economic Intensification, 187–188 El Niño Southern Oscillation (ENSO), 60 Elehuascui, 103 Emergent Sociopolitical Complexity, 209 Endogamous marriages, 103 Environment, California, altithermal conditions, 70 Environmental context, 41–71 Environmental debate economic developments, 7 sociopolitical developments, 7 Ethnographic Atlas, 37 Ethnohistoric data, and Chumash society, 73 European disease, and depopulation, 93 Faunal assemblages Daisy Cave, 122 fishing, 124 Late Holocene, 129 marine mammals, 29 Fish bone densities, exponential increases in, 219 exploitation, behavioral ecologists and, 24 faunal data, 190–191 modern distribution, 192 Fishing Chumash society, 192 strategies, 23 resource return rates, 25 techniques, 192 variety of, 24 Food return rates, 18 Foragers. See also Maritime foragers Inujjuamiut, 27 marine resources, 20 predictions in CPF theory, 228 strategies, 33 Geographic analysis, 104–105 Geographic Information System, 6 Geographic position in village history, 105

293

Glass beads, during historic period occupation, 102 Glassow, Michael, on Santa Cruz Island, 6 Globigerina bulloides, 64, 67 Harbor seals, 59 breeding, 57 HBE. See Human behavioral ecology (HBE) He’lewashkuy (Elehuascui), 103 Hichimin (Cheumen), 98 High-ranked prey species, population dependence on, 28 Historic communities intervisibility, 107 political integration, 107 social integration, 108 Historic island communities, 91 villages, 93 Historic village problems in identifying, 96 rank-size analysis, 109 Holocene climatic conditions, 69 cultural evolution, 8 marine mammals, 28 pollen, lake level, 70 population growth, 225 stages, climatic conditions of, 185 House depressions, Late Holocene, 172 Human behavioral ecology (HBE) adaptive design, 3 archaeological Darwinian approach, 12 behavioral variability, 3 community formation prediction, 37–38 ecological context, 217 ecological interaction, humans and environment, 12 ecology, 236–237 integration, 216 maritime settings, 217 framework, 1–2 maritime societies, 10–40, 236–238 models, ethnographers and, 13 neo-Darwinian principles, 3 optimal foraging theory, 14 resource patchiness, 29 vs. cultural ecological principles, 12

294

INDEX

Human induced degradation intertidal habitats, 218 Hunting historic period, 196 mammoth, 120 marine mammal, 197 ethnography, 27 importance, 26 Hydrology, of northern Channel Islands, 48 Ideal despotic distribution (IDD) models, 231, 232, 234 and IFD, 35 Ideal free distribution (IFD) model, 229, 232, 234 and intensification, 32–36 and resource patches, 34, 36 predictions, 34–35 Intensification free distribution, 229–231 Interior cemeteries, placement of, 231 Intertidal habitats, 221 human induced degradation, 218 Inujjuamiut foragers, 27 Island Chumash, 1–9 society, 152 study area, 4–6 Island populations, 162 social organization, 127 Jolla Vieja canyon, Santa Rosa Island, 146 Jones, Philip Mills, 99 Kaxas, 211 King’s Chronology, of Early Holocene, 84 King, Chester, 173 Lambert, Patricia, 183 Landberg’s model, from Point Conception region, 76 Late Holocene, 82, 85, 89, 220 archaeological sites, 171 artifact assemblage, 212 associated artifacts, 158–163, 165–168 associated domestic features, 158–163 Bee Rock, 102 chert bifaces, 176 coastal settlement, 169 community differences, 211 craft items, 212 diagnostic artifacts, 175

faunal assemblages, 129 house depressions, 172 population increase, 37–38, 158 radiocarbon dates, 164 San Miguel Island, 172 settlement disruptions, 174 terrestrial conditions, 188 village distribution, 156 expansion, 229 history, 99 Late Holocene Record, 154–216 Lethal violence, 183 increases, 185 Liyam, 211 Logistical encampments, 129 Mainland coast population, ethnohistoric summary of, 80 Mammoth absolute number, 220 extinction, 120 hunting, 120 Marine climate history, 64 Marine foods, northern Channel Islands, 151 Marine mammals availability, 26 bone, absence of, 198 exploitation, 28 faunal assemblages, 29 hunting, 197 ethnography, 27 importance, 26 return rates, 28 secondary products, 26 Marine productivity, 184, 189 California Current, 56 Marine resources, 201 modern foragers, 20 return rates, 20–29 shellfish, 21 spatial distribution, 54–58 variety, 75 Marine traditions, in Chumash society, 152 Maritime foragers, 193 archaeological data, 30 central place foraging, 29–32 coastal habitats, 15 ethnographic data, 30 strategies, 13–16, 193 wide diet breadth, 31–32

INDEX

Maritime settings diet choice, 16–20 HBE, 217 Maritime societies, 236 HBE, 10–40, 236–238 Marriage records, and village population, 106 Marriages, endogamous, 103 Matrilocality, 187 Medea Creek, Santa Monica Mountains, 87 Metamorphic stratigraphic sequences, 45 Microlith production, 206 Middle Holocene, 61, 82 artifact types, 128 bird bones, 147, 148 burial patterns, 173 climatic conditions, 227 coastal settlement, 170 diet, 148 dietary differences, 227 faunal data, 149 interior settlement, 126 Santa Cruz Island, 169 settlement, 144, 152 village locations, 175 village sites, 170 Miocene sediments, 45 Mollusk assemblage, 218 comparison, 228 Monterey chert, eastern Santa Cruz Island, 46 Monterey formation, volcanics in vicinity of, 46 Morris, Don, 100 Mortuary data, 214 Museum of Natural History, 100 Mussels, 58, 144, 146, 179, 221 Nahuani, 101, 102 Nanawani, 97 Natural selection theory, HBE, 11 Nawani (Nahuani), 101, 102 Needle drilled shell beads historical indicators, 92 Neo-Darwinian principles, HBE, 3 Neogloboquadrina pachyderma, 67 Niaqla (Niacla), 100 Nilal’uy (Nilalui), 102 Nimkilkil (Nimquelquel), 101 Nitrogen isotopic analysis, 148 Niwoyomi (Niuoiomi), 104 Northern Channel Islands, 112, 208, 217

295

adjacent mainland, map of, 2, 95 bathymetry, 62 breeding populations, 197 Chumash population, 4 coastal communities, 173 colonization history, 1 environmental differences, 41 evolutionary history, 1 general physiography, 42 geographic history, 111 geographic position, 106 geology, 42–45 hydrology, 48 kelp forests, 57 physiographic differences, 41 plant communities, 50 Pleistocene occupation, 118 resource availability, 59 topographic map, 43 Olivella biplicata, 98, 100, 127, 175, 202 shell beads, 202, 212 distribution, 203 production, 199 radiocarbon dates, 204–205 Olson, R.L., 96 Optimal foraging theory, HBE, 14 Osteological data burial populations, 148, 227 sublethal violence, 185 Otter Creek, Niwoyomi (Niuoiomi), 104 Oxygen isotopes, 175 Paleo-Indian foraging strategies, 121 population, 121 Paleoclimatic, 174 marine productivity, 215 records, 65 Patrilocality, 187 Pico, Juan Estevan, 104 Pinus muricata, 54 Pinus torreyana, 54 Plank canoe, 193 Plant foods ethnohistorical accounts, 200 consumption by women, 227 foraging range, 169 scarcity, 218 trade, 200–201 Pleistocene eolianite sandstone, 45

296

INDEX

Pleistocene (continued) human colonists, 219 human population, 220 shell deposits, 119 Pollen zones, union project, 68 Population density and group sizes in coastal settings, 36–37 vs. suitability, 35 Population-dependent degradation, 38 Population growth coastal villages, 155 cultural developments, 10 demographic expansion, 155–156, 229 social restrictions, 230 Primary villages distribution, 231 expansion, 230 Primate distribution, 108 Prime movers, 10 Primo convex distribution, 108 Punta Arena. 138 Pygmy mammoths, 49, 219 high-ranked prey, 220 Qshiwqshiw (Siucsiu), 211 Old Ranch Canyon, 98–99 Quercus agrifolia, 54 Radiocarbon dating, 143 AMS, 195 chronological information, 158 northern Channel Islands, 155 frequency, 157 Olivella shell beads, 204–205 Rainfall fluctuation, 60 Rank and size analysis, 107, 108–111 Red abalone, 58 species significance, 188 Resources dietary utility, 17 distribution coastal habitats, 29 foraging behavior, 17 intensification, 33 spatial component, 33–34 patches IFD, 38 procurement energetics, 32 ranking, 17

Rincon formation, 45 Rocky near shore environments, 122 Rogers, David Banks, 10 San Miguel Island, 112 breeding colonies, 222 Daisy Cave, 114 interior settlements, 147–148 intertidal habitats, 145 Late Holocene, 172 plant food, 219 primary productivity, 57 quaternary terraces, 45 Tuqan (Toan), 104 villages, 181 Santa Barbara coastal population, 76 ethnohistoric records, 72–90 Santa Barbara Channel, 4, 50, 68 chronology, 8 Chumash people, 75 ethnohistoric record, 72 populations, 89 Portolá expedition, 73 prehistoric record, 80 prehistory, 158 region, 10, 82, 125, 187 animal nutritional information, 22 burial data, 87 chronological schemes, 80, 83 Chumash society, 50, 154 climate, 46–48 cultural changes, 89 ethnohistoric data, 75 plant nutritional information, 22 warfare, 180 Santa Cruz Island, 6 Abalone Point, 101 aerial view, 208, 210 chert, 206 groundstone, 201 Kaxas, 209, 211, 212 L’akayamu, 101 Punta Arena, 123 Scorpion Anchorage, 97 sea otters, 222 topographic relief, 53 Santa Miguel Island, 146 groundstone, 218 Santa Monica Mountains Chumash society, 88 Medea Creek, 87

INDEX

Santa Rosa Island, 62, 97–98, 112 grasslands, 148 Highlander phase, 140 interior locations, 183 intervillage relationships, 103 Jolla Vieja canyon, 146 logistical foraging, 226 Niaqla, 106 Nimkilkil, 106 northwest coast, 118 Old Ranch Canyon, 99 pygmy mammoths, 119 Qshiwqshiw, 105 territory boundaries, 183 villages, 178, 181 expansion, 158 water reliability, 49 Satellite communities, 232 Scorpion drainage, 143 Sea level California coast, 67 changes global, 63 local, 63 Early Holocene, 117 Late Pleistocene, 117 stabilization, 32–33 Sea lions, California, 196 Sea mammals absence, 121 large concentrations, 59 Sea surface temperature (SST), 64, 184, 189 kelp growth, 60 millennial-scale fluctuations, 223 monthly average, 56 seasonal fluctuations, 58 Seasonal shellfish harvesting profiles, 151 Sespe formations, 45 Settlement disruptions, Late Holocene, 174 Settlement packing, 36–37 Shellfish assemblages, 218 beds, 21 carbohydrates and calories, 21 dietary importance, 147 faunal data, 190–191 harvesting, 151, 179 logistical encampments, 226 return rates, 23 Silmihi (Silimi), 99–100

297

Siucsiu, 211 Old Ranch Canyon, 98–99 Skull Gulch, 100, 101 Skunk Point, 103 Social and environmental instabilities, 232 Social hierarchies competition and formation, 233–236 coastal settings, 36–39 Social instability, 8 Sociopolitical complexity, development of, 86 Sociopolitical evolution, climatic influences and, 7 Sociopolitical organization, 77–78 Southern California coastal ranges, 71 ENSO, 60 major coastal current systems, 55 marine snails, 19 SST. See Sea surface temperature (SST) Stanton Ranch precipitation, 47 temperature, 47 Stylistic changes, in shell fishhooks, 195 Sublethal violence, 183 Swaxil, 96 Synthesis, 217–238 Terminal pleistocene, 80, 112–113 Terrestrial conditions, Late Holocene, 188 Terrestrial microenvironments northern Channel Islands, 49 Terrestrial resources postencounter return rates, 19 spatial distribution, 49 Territorial defense, forms of, 38 Territoriality, 230 Ethnohistoric evidence, 79 Tertiary sedimentary sequences, 45 Toan, San Miguel Island, 104 Trade, 198 Trade items, distribution of, 200 Transportation, energetics of, 32 Trapezoidal microblade production, 207 Tricdacna gigas, 22 Tuqan (Toan), 99 San Miguel Island, 104 Unification consumption, 107 rank size analysis, 108

298

INDEX

Vaqueros formations, 45 Vertical stratification, productivity and, 66 Viewshed analysis, 105 Village Arlington Canyon, 139 history baptismal records, 105 geographic position, 105 intervisibility, 106 marriage records, 106 Late Holocene distribution, 156 expansion, 229 history, 99 locations

Middle Holocene, 175 population, 96 growth, 155 northern Channel Islands, 105 rank-size analysis, 109 San Miguel Island, 181 Santa Cruz Island, 158 Santa Rosa Island, 158, 178, 181 Violence, lethal, 183 increases, 185 Warfare, ethnohistoric evidence of, 79 Water temperature, 184 Weathering, chemical, 119 Whale bones, 196 White Mountains, California, 71