Untitled

August 28, 2018 | Author: Anonymous | Category: Documents
Share Embed


Short Description

In the first detailed analysis of an extensively excavated late summer- .... You are a prince and a true friend in all w...

Description

A

FOLSOM TECHNOLOGICAL AND SOCIOECONOMIC STRATEGIES: VIEWS FROM STEWART’S CATTLE GUARD AND THE UPPER RIO GRANDE BASIN, COLORADO

by Margaret Ann Brierty Jodiy submitted to the Faculty of the College of Arts and Sciences of American University in Partial Fulfillment of the Requirements for the Degree of Doctor of Philosophy in

Chair: Dr. Richard J. Dent

c

.

D Dr. George C4rison

Dean of the College of Arts and Sciences

1999 American University Washington, D.C. 20016

' COPYRIGHT by MARGARET ANN BRIERTY JODRY 1999 ALL RIGHTS RESERVED

To Marie Wormington, Dee Ann Story, and Betty Meggers, much respected and loved friends and mentors who provided strong personal and professional role models as archaeologists, and also as women in archaeology.

When some viewed my work through the filter of Dennis Stanford, you each accepted me on my own (de)merits and provided intellectual vitality, sage advice, opportunities, good humor, and excellent conversation over many years. Muchas gracias y abrazos.

FOLSOM TECHNOLOGICAL AND SOCIOECONOMIC STRATEGIES: VIEWS FROM STEWART’S CATTLE GUARD AND THE UPPER RIO GRANDE BASIN, COLORADO BY Margaret Ann Brierty Jodry ABSTRACT The purpose of this study is to increase understanding of the Folsom Cultural Complex by linking technological organization with social dynamics, economic orientation, and environmental context. Two assumptions underlie this endeavor: primary subsistence practices strongly influence patterns of social and technological organization, and environmental circumstances hold important keys to the structure and variation of social and subsistence activities across the landscape. It is suggested that Folsom social organization and technology developed in the context of significant changes in climate and subsistence resource structure during the Younger Dryas cool episode (nearly 10,000 to 11,000 BP) when expansion of grassland ecosystems likely favored radiation in bison populations. An ecological model is proposed whereby the large size of Folsom-age bison is related to longer somatic growing seasons, favorable moisture regimes, and a postulated increase in nitrogen availability that resulted in abundant, high quality forage and a possible rise in bison fertility rates. These fundamental changes altered predator-prey relations in the aftermath of Rancholabrean extinctions and permitted economic development of bison specialization. Cooperative hunting during the late summer-early fall involved upwards of fifty animals in a single kill. It is suggested that recurrent, large-scale bison processing necessitated cooperative labor and a few specialized tools. Technological innovations included a possible woman’s fillet knife to thin-cut meat for drying and weapon tips designed to kill bison that were easily repaired and conserved raw material. High residential mobility and extensive networks of kin and other social alliances were means of offsetting risks inherent in subsistence specialization.

11

Study of sixteen raw materials at forty-five sites suggests that regular interregional communication may have characterized the period. In the first detailed analysis of an extensively excavated late summerearly fall bison kill and processing camp (Stewart’s Cattle Guard site) the archeological remains of hide working identify activities likely performed by females and weaponry repair and manufacture distinguish male tasks. Use-wear, refit, and spatial studies provide insight into the demands placed on end scrapers and projectile points. Their attrition rates may vary by season. These findings provide a useful interpretive tool for assemblages of many time periods.

111

ACKNOWLEDGEMENTS It is a privilege to thank friends, colleagues and relations for their contributions to field research in the upper Rio Grande Basin, to analysis of Stewart’s Cattle Guard site, and to completion of this dissertation. Just as exchanges of resources, cooperative labor, time and emotional support bound Folsom people together, I am likewise tethered, in a spirit of gratitude and reciprocity, to many people and institutions for these same reasons. The Smithsonian Institution’s intellectual and financial support enabled Dennis Stanford and myself to conduct long-term studies in Paleoindian archaeology and paleoecology in Colorado, Alaska and elsewhere. The National Museum of Natural History is a stimulating and wondrous place to conduct research and interact with the public. Among its strengths are the dedicated people who work there. Marcia Bakry comes immediately to mind because she worked long hours, nights and weekends, preparing illustrations for this dissertation. I am very grateful for her assistance and advice. Likewise, Vic Kranz photographed many lithic artifacts, and Don Hurlbert, Carl Hansen, Mike Frank and Bruno Frohlich digitized photographs and slides and provided much assistance with computer applications. Marvin Kay and students at the University of Arkansas photographed artifacts included in Marvin’s use-wear study, many of which were incorporated in Marsha’s illustrations. Marvin prepared illustrations showing photomicrographs of high magnification use-wear. His good humor and insights are appreciated. At the Smithsonian, I thank JoAllyn Archambault, Gus Van Beek, Kay Behrensmeyer, Ralph Chapman, Maggie Dittemore, Bill Fitzhugh, Mike Frank, Bruno Frohlich, Candace Greene, Deb HullWalski, Johanna and Bob Humphrey, Dave Hunt, Carole Lee Kin, Stephen Loring, Betty Meggers, Bill Merrill, Sake Homiak, Bill Melson, Bill Merrill, Dan Odess, Don Ortner, Zaboria Payne, Rick Potts, Dan Rodgers, Carolyn Rose, Dave Rosenthal, Ruth Selig, Joyce Sommers, Lynn Snyder, Dennis Stanford, Agnes Stix, Verdis Thomas, Barbara Watanabe, Kim Waters, Waldo and Mildred Wedel, and Mindy Zeder for assistance, support, and friendship. The supportive postcards are classics, Stephen. iv

American University generously supported graduate training and dissertation research with a Graduate Fellowship from the Department of Anthropology, a Deans Scholarship and a Mellon Travel Award from the College of Arts and Sciences, and a Dissertation Fellowship and Super Cohort Scholarship from the university. Thanks to Drs. Dent and McNett who served on my committee, provided advice during my semesters as a teaching assistant, and shared archeological expertise. Thanks also to Drs. Bodine, Burkhart and Leap who deepened my appreciation for sociocultural, linguistic, and engendered perspectives in archeological research. I learned a great deal serving as a teaching assistant for Drs. Leap, Bodine, Dent, and McNett’s and valued the opportunity to design and teach a section of Introduction to Archeology. Dr. Joan Gero joined the faculty recently and generously read the paleoecological portion of this dissertation despite her own heavy workload. Her careful reading and advice are appreciated. Thanks to Melanie Stright for enjoyable conversations regarding sea level change and the archeology of McFadden Beach and to Chris Jirikowic for discussions of archeological theory and interpretation and for assisting with excavations at Black Mountain and Cattle Guard. I owe a great debt of gratitude and friendship to Dr. George Frison who served on my committee and greatly benefited the final form of this dissertation. His prolific contributions to Paleoindian and High Plains archaeology are an inspiration to us all and his pioneering bison studies, in collaboration with Dr. Chuck Reher, provided a strong stimulus for this study. Financial support from the Colorado Natural History Small Grants Program and the Colorado Mountain Club Foundation for paleoecological studies at Black Mountain Lake is gratefully acknowledged. Many thanks to the Institute of Arctic and Alpine Research for assistance with the Black Mountain Lake core and with Cattle Guard sediment analysis. Mort and Joanne Turner were especially helpful and Mort’s ideas regarding nitrogen enrichment during the late glacial inspired me to look in this direction with my own research. Further appreciation is extended to Rolf Kihl, Tom Stafford, and Jocelyn Turnbull. I particularly acknowledge and thank Mel Reasoner for his fine analysis and for an enjoyable collaboration on the Black Mountain palynological record. Thanks also to those who assisted with efforts to recover lacustrine cores, particularly Phil Leggett of the Mineral County Search and Rescue Team, who provided v

winter access and Vince Spero of the Rio Grande National Forest who assisted in myriad ways large and small with Black Mountain research and who helped fund several radiocarbon dates. Many thanks to Drs, Owen Davis and David Shafer for their palynological analyses of Head and Como Lakes and for drawing my attention to the importance of Sporormiella as a proxy indicator of biomass. Steve Fryberger, Sarah Andrews, Rich Madole, Jim Benedict, Jim Jordan, Owen Mason, and Bill Melson provided additional geological expertise, field assistance and reprints for which I am very grateful. Logistics in the San Luis Valley were aided immeasurably by the good people at Great Sand Dunes National Monument, Great Sand Dunes Oasis and the Zapata Ranch. One hesitates to single out individuals for fear of forgetting someone, but I would be remiss in not acknowledging the efforts of Barbara Irwin, who has been a logistical angel of Cattle Guard research teams. Many thanks also to Bill Wellman, Fred Bunch, Bob Schultz and to Berl and Barbara Lewis, Malcolm and Rose Marie Stewart, the Vitorrio family, Jim and Joyce, Hisa Ota, Ken Klemm, Tad Carpenter, and Scott and Kevin Desplanques. The Bureau of Land Management provided interdisciplinary expertise, crew vehicles and fencing materials. John Beardsly, BLM archeologist for the San Luis Valley during most of the CG project brightened field camps with his company and extraordinary meals. Thanks to John for his support and for encouraging me to think anthropologically about things archaeological. Thanks also to Mike Casseel for sending reprints, to Ken Klemm and others whom helped build the fence around CG, which prevents modern buffalo from trampling ancient buffalo bones. This study reflects the work of many field crews in the sand dunes of the San Luis Valley, who worked diligently and well to produce an excellent body of data with which to study Folsom technological and social organization. The committed folks who returned season after season provided important continuity in the work and enabled the project to hit the ground running during short field seasons. Tim Kearns, Siste Omaha, Jerry Dawson, Chris Kugler, Jim Rancier, Paul Griegg, Stephanie Rippel, David Homer, Jerry Slack, Barbara Munford, Rick Duncan, Jim Enloe, Patty Walker Buchanan, Bob Estes, David and Barbara Breternitz, Phil LeTourneau, Beth Ann Camp, Dave Breternitz, Rusty Greaves, Sam Richings, Molly Coxson, Dave Rosenthal, Mike Frank, Dan Prikryl, Galen Burgett, Marcel Komfeld, Pete Arena, vi

Daniel Talonn, Richard Stark, Chuck Wheeler, Eldon Yellowhawk, friends from Alaska and all others are heartily thanked for assistance, camaraderie, and interesting conversations. Bob Estes is especially thanked for assistance with mapping, aerial photography and illustration. We thank Barbara Breternitz and Siste OMaha for providing such good meals and company. Thanks also to Liz Morris and Pete Eidenbach for field assistance and reprints. Finally, my grateful appreciation to members of the Denver and Pueblo chapters of the Colorado Archeological Society and to the San Luis Valley Archeological Network. Faunal analysis in recent years is a collaborative effort with Lynn Snyder. Her pragmatic views, wit, and expertise are much appreciated. When this debt comes due, may it find us pondering your assemblages in Greece. Thanks also to Janine Davis, George Frison, Larry Todd, Matt Hill, Jim Enloe, Dr. Detling, Marvin Reynolds, Mary Ann Graham and Cody. Technological study of the lithic assemblage owes much to J. B. Sollberger, Glenn Goode, Bruce Bradley, Bob Patten, Dennis Stanford, George Frison, Jack Hofman, Stan Ahler, Marvin Kay, John Tomenchuk, Mike Collins, ElTett Callahan, Chuck Reher, Dan Amick, Matt Root, Terry Del Bene, Larry Langford, Brad Vierra, and Jay Newman. I thank the participants and organizers of the Folsom workshops for helpful discussions. Special thanks to Stan Ahler and Phil Geib for allowing me to include their model of Folsom hafted design so that I could test this hypothesis with breakage data from CG. Additional thanks to Stan for insights regarding use-wear at CG, for providing rephicative end scrapers, and for your supportive friendship. My thanks also to Gene Titmus, Bruce Huckell and Bob Patten, who provided replicative bifaces for experiential use in thin-cutting meat. Special thanks to Bruce Bradley, who helped with this experiential work and provided elk meat and his kiva in which to dry the jerky. For all the times you examined CG material with me, Bruce, and for your supportive friendship, and that of Cindy, I am most beholden. To Aldan Naranjo, Bertha Grove and Everett Burch, warmest regards and appreciation are sent. I am very fortunate to have spent time with your wonderful family. Thank-you for sharing your knowledge and company with me, it helped the work very much. I look forward to my next trip to Ignacio. David Christensen also taught me many valuable things about hide working and ethnographic clothing from the vii

Plains. I will not forget standing in the tipi and realizing for the first time that such a dwelling enables one to live within the heart of a buffalo herd. Thanks also to Jim Riggs, Matt Richards and Larry Belitz for answering my queries regarding tanning. Special appreciation and abrazos to my friend Jan Motriuk for never doubting I would finish this dissertation, for providing support throughout, and for never losing sight of the humor in any situation. Your feast in DC is remembered fondly by many. Thanks to Ken Brown, whose dissertation inspires me and whose keen feedback is always thought provoking. I also thank Dan Odess who arrived with Dennis on Friday nights to remind me that "enough is enough, time to drink a beer". For all the times you provided a sounding board as I thought out loud (and challenged you to get a word in edgewise), muchisimas gracias. It helped the work, as did your editing and critique. I thank members of my immediate and extended families that grace my life and provide moral support. My mother, Mary Brierty, exemplifies every day how to do more than is expected, how to do it cheerfully and generously, and how to nurture others. My father, Ed Brierty, taught me to strive for scholarship and integrity, to cultivate wit, and to embrace our Irish heritage. Thank you both for so many things throughout my life. Many thanks also to Patty and Kevin for support and interest in this research. To Donna and Brandy I can only say that I feel as if I hit the jackpot when you two entered my life. Thankyou for help and adventures in the field, and for providing encouragement and feedback during the writing of this dissertation -- and otherwise. My life is greatly enriched by our bonds of family, friendship and good humor. To Cody, a special hug and repayment of hiking IOUs. Last and most important -- I extend my deepest regards and gratitude to Dennis Stanford, who shared all aspects of this journey. A spectacular display of Northern Lights is needed to accurately convey my appreciation for the spirit with which you supported this endeavor. My debt is great and only we can take full measure of it. You are a prince and a true friend in all weather.

vi"

CONTENTS

ABSTRACT

. ii

ACKNOWLEGMENTS..............................................................................................................................iv LISTOF TABLES ......................................................................................................................................xiv LIST OF ILLUSTRATIONS ....................................................................................................................xvii Chapter 1. THE RESEARCH PROBLEM ............................................................................................... Introduction....................................................................................................................

1

Approachesto the Study ................................................................................................. 5 Organization of the Dissertation .....................................................................................

8

2. UPPER RIO GRANDE BASIN PALEOENVIRONMENT .................................................. 14 Introduction...................................................................................................................

14

The San Luis Valley and Upper Rio Grande Basin ....................................................... 15 Physiography...................................................................................................

15

Hydrologyand Geology .................................................................................. 18 ModernClimate ............................................................................................... 21 Younger Dryas Climatic Oscillation ............................................................................. 23 Palynological Studies in the Upper Rio Grande Basin .................................................. 24 Introduction.....................................................................................................

24

BlackMountain Lake ...................................................................................... 24 HeadLake ........................................................................................................ 33 Sporormiella and Large Animal Biomass ....................................................... 34 ComoLake ...................................................................................................... 36 lx

Geologic and Fauna! Evidence for Climate Change Stratigraphy, Black Mats and Small Mammals

...............................

38

......................................

39

Effective Temperature During the Folsom Period in the San Luis Valley Moisture, Nitrogen, and Forage Production

45

.................................................................

46

....................................................................................

49

Primary Productivity in the San Luis Valley

.........................................

Folsom Site Distributions in the Upper Rio Grande Basin

53

.....................................................................

56

...........................................................................................................

56

Site Location and Setting History of Investigations

......................................................................................

56

.......................................................................................

63

Discovery and Initial Testing

.................................................................

Continued Investigation, 1983-1996 ExcavationTechniques Geology

......................................................

.........................................................................................

.................................................................................................................

Bone Weathering Patterns CarnivoreFeeding

63 66 68 70

.....................................................................................

76

.................................................................................................

77

Factors Enhancing Preservation of Activity Areas

...............................................

4. LOCAL AND REGIONAL LITHIC RAW MATERIAL VARIATION Introduction

50

...................................

3. SITE SETTING AND INVESTIGATION Introduction

41

.........................................................

Nitrogen Enriched Snowrnelt GrazingIntensity

...........

......................

...........................................................................................................

Raw Material Variation at Cattle Guard

...............................................................

Methods and Background Information

...................................................

Black Forest Silicified Wood and Trout Creek Jasper

x

81 81 84 84

............................

88

...............................

93

..................................................

98

Cumbres, Mosca, Hornfels and Morrison Quartzite Chuska, Alibates, and Edwards Cherts

79

Summary of Lithic Procurement at Cattle Guard Raw Material Abundance and Condition

..........................................................

......................................................................

Mobility Patterns, Raw Material Use and Retooling

111

...............................................................

116

....................................................................................................

116

Who Used What.

.

.

And Where?

..................................................................

Gift Exchange and the Long-distance Transport of Materials

......................................

Historic and Recent Land Use in the Study Region by Native People Protohistoric Alliance and Exchange in the Study Region ConcludingRemarks

...........................................

141

Production Technology Bifaces

.....................

...................................................................................................................

....................................................................................

Description of the Analytical Sample

.............................................................

..................................................................................

...........................................................................................................................

Preforms

136 138

.....................................................................................................

StoneTools and Flaking Debris

120

.........................

5. STONE TOOL TECHNOLOGY AT STEWART’S CATTLE GUARD SITE Introduction

101

....................................................

Inter-Regional Patterns of Raw Material Use Introduction

99

.........................................................................................................

Stage 1-3: Blank Acquisition and Initial Shaping

.............................

Stage 4-7: Thinning of First Face and Channel Flake Removal

.......

144 148 148 149 149 15 1 157 158 159 162

Stage 7-9: Thinning of Second Face and Channel Flake Removal

168

Stage 10: Post Fluting Retouch

170

Stage 11: Margin Polishing ChannelFlakes ProjectilePoints Design

........................................................

..............................................................

..............................................................................................

...............................................................................................

...............................................................................................

Breakage Patterns: Tips

xi

....................................................................

173 174 177 179 183

..................................................................

191

........................................................................

193

..............................................................................................

195

Breakage Patterns: Bases Incidenceof Burning UltrathinBifaces

Production Technology

....................................................................

............................................

197

........................................................................

204

.....................................................................................

212

........................................................................................................................

218

Description of Cattle Guard Ultrathins Design and Function MeatStorage Unifaces

195

.................................................................

219

....................................................................................................

225

Flake Knives and/or Side Scrapers Endscrapers

Hide Processing: Seasonal Rhythms and Tool Kits Gravers and Perforators

..................................................................................

Flake Tools Modified by Use Only

................................................................

Occurrence of Red Coloration on Black Forest Artifacts ConcludingRemarks

............................................

....................................................................................................

6. ACTIVITY PATTERNING AND SITE FORMATION Introduction

........................................

........................................................

..................................................................................................................

249 25 1 255 257 262 262

....................................................................................................

265

.......................................................................................................

265

ChapterOrganization Fauna! Assemblage

235

...................................................................................

273

..........................................................................................

276

Kill and Initial Butchering Area Pattern Recognition Studies

..........................................................

276

..............................................................................

285

Piece-Plotting and Density Contouring K-Means Cluster Analysis

...................................................................

286

.................................................................

294

.........................................................................

309

K-Means Bone Clusters K-Means Lithic Clusters Preliminary Use-Wear Results

xl’

Concluding Remarks 7. SUMMARY AND CONCLUSIONS

. ............................................................................

APPENDIX A: RAW MATERIAL DESCRIPTIONS AND DATA TABLES

331 349

...................................................................

349

.............................................................................................

349

Measurements and Fragmentation, All Tools

Additional Variables, Flake Tools

....................................................................................

Additional Variables, PrefonTis and Folsom Points

BIBLIOGRAPHY

325

.................................................

APPENDIX B: ANALYSIS VARIABLES AND ATTRIBUTES

General Variables, All Tools

............................

318

..........................................................

............................................................................................................................

xiii

352 353

355

TABLES

Table

Page

1 Dental Ages and Estimated Season of Death at Folsom Bison Kill and/or ProcessingSites . ...................................................................................................................

3

2.

Comparative Precipitation Data from the Upper Rio Grande Basin .

3.

Modern Climatic Data for Summitville and Hermit Lakes, Upper Rio Grande Basin .

.........

26

4.

Uncalibrated AMS Radiocarbon Dates from Black Mountain Lake Core BML 97-2 .

.........

28

5.

Effective Temperature Estimates for the Folsom Period, San Luis Valley .

......................

42

6.

Relative Frequencies of Animal Species Reported from Twenty-three Folsom Sites .

7.

Raw Material Utilization in Folsom and Post-Folsom Components at Cattle Guard .

8.

Inventory of Excavated Area and Recovery by Field Season, 1981-1996 .

9.

Key to Folsom Sites in Figure 24

..............................

...........

61

...........

65

........................

.......................................................................

10.

Raw Material Frequencies at Five Folsom Sites in the Upper Rio Grande .

11.

Percentage Frequency of Raw Material Types in Total Lithic Sample By Selected Artifact Category .

22

68 89

.......................

93

........................................................................

103

12.

Mean of Approximate Transport Distances at Different Percentage Compositions .

13.

Four Forager Mobility Models Compared with Raw Material Transport Distances Reflected At Selected Folsom sites .

134

14.

Block Sample by General Artifact Category .

.........................................................

149

15.

Total Sample by General Artifact Category .

.........................................................

151

16.

Dimensions of Two Possible Cores .

17.

Original Form of Blanks Used to Make Flake Tools .

18.

Minimum Number of Bifaces in Total Sample by General Tool Class And Raw Material Type .

158

Type of Preform Breakage by Manufacturing Stage .

170

............

..............................................................................

...................................................................

.........................................................

..............................................................................

19.

xiv

................................................

124

154 157

Table

Page

20.

Marginal Retouch Density for Points, Preforms and Channel Flakes .

21.

Fluted Preform Descriptive Statistics .

22.

Channel Flakes, Preforms, and Folsom Points in Total Sample by Raw MaterialType .

175

23.

Channel Flake Measurements by Fragmentation Type .

175

24.

Comparison of Channel Flake Size at Selected Folsom Sites .

25.

Unfluted Folsom Point Descriptive Statistics .

26.

Fluted Folsom Point Descriptive Statistics .

27

Variation in Tip Length .

............................

172

.................................................................

174

..........................................................................................

.

...........................................................

.....................................

.........................................................

...........................................................

.................................................................................

177 179 179 183

28.

Incidence of Burning Among Stone Tools and Fragments .

29,

Ultrathin Bifaces and Ultrathin Preform Fragment .

30.

Comparison of Nitrogen Isotope Data for Bison at Cooper Site and Pecos Ruin .

31.

Comparison of Stone Tool Function at Cattle Guard and Bob Tail Wolf Sites .

32.

End Scraper Descriptive Statistics .

33.

Relative Occurrence of End Scrapers at Three Sites .

34.

Percentage Frequency of Raw Material Types for Gravers and Perforators inTotal Lithic Sample .

251

35.

The Occurrence of Reddening on CG Artifacts .

......................................................................

256

36.

Minimum number of Selected Bison Elements .

......................................................

37.

Sample of Mandibular Molars used to Estimate CG Seasonality .

38.

Dental ages and Estimated Season of Death at Folsom Sites .

39.

Representation of Bison Bone Articulations by Carcass Segment .

40.

Projectile Point Breakage and Material Types in the Kill/Initial Butchery Area .

41.

Comparison of Skeletal Segment Representation in K-means Bone Clusters .

42.

Comparison of Selected Tool Classes and Material Types in K-Means Lithic Clusters .

43.

Sample of Unifacial and Bifacial Artifacts Studied by Use-Wear Specialists .

.....................................................

.................................................................

194 195

...................

215

......................

219

.....................................................................

225

...............................................................

..........................................................................................................

xv

...........................................

......................................

..........................................

....................

........................

........

........................

244

268 268 270 271 273 292 307 309

Table

Page

44.

Bison Carcass Density and Number of Tools Recovered in Seven Folsom Bison Kills .

45.

Minimum Number of Selected Tools Broken, Discarded, and/or Lost at Cattle Guard and Estimates of Minimum Number of Adult Males and Females Based on Them .

.........

314

............

320

..............................

332

46.

Raw Material Type Descriptions, Stewart’s Cattle Guard Analysis .

47.

Reference Information for Primary Lithic Sources Depicted in Figure 32

48.

Representation of Key Raw Materials at Selected Folsom Sites .

49.

Mean Distances from Possible Primary Source to Archeological Sites for Selected Raw Materials in Percentages Greater than Twenty .

........................

..................................

.............................................

50.

Mean Distances from Possible Primary Source to Archeological Sites for Exotic Raw Materials in Percentages of Six .

.................................................................

.

......

336 338

340

340

51.

Artifact Frequencies for Sample Included in Stanford 1991 and This Raw Material Study.

341

52.

Refit Groups by Artifact and Breakage Type .

341

53.

Frequency of Unifacial Tools by Material Type, Size, and Presence of Cortex .

54.

Frequency of Retouched Tools versus Non-retouched Tools by Material Type .

55.

Functional Classification of Stewart’s Cattle Guard Flake Tools .

56.

Dimensions and Use-Wear Traces on Flake Knives and/or Side Scrapers .

57.

Basic Steps to Dry-scrape, Brain-tan Bison Hides .

........................................................

xvi

.....................

342

.....................

342

............................................

343

.............................

344

..................................................

345

ILLUSTRATIONS Page

Figure 1. Rio Grande Headwater Region in the San Juan Mountains at 3096 m Elevation . ...............

16

San Luis Valley Floor 2300 m Elevation with Sangre de Cristo Mountains RisingUpwards of 4200 m

17

3.

Map of San Luis Valley, Colorado Showing Hydrologic and Physiographic Features

19

4.

Schematic Cross-Section of the Central San Luis Valley Showing Geology, Hydrology and Recent Vegetation Zones .

20

Physiographic Map of South-central Colorado Showing the Locations of Palynological Studies .

25

Geochronology and Magnetic Susceptibility for a Portion of Black Mountain LakeCore 97-2 .

27

7.

Preliminary Pollen Diagram From Black Mountain Lake, Colorado . ......................................

29

8.

Head Lake, Colorado: Palynomorph Percentage Diagram . ........................................

35

9.

A Climatic Model of Temperature and Precipitation for Alamosa, Colorado . ...................

43

10.

Map Showing Density Distribution of Mean Annual Precipitation and Runoff

48

11.

Views to the Northeast of: a) Stewart’s Cattle Guard Site and b) Surrounding Area .

12.

View Southeast of Grazing Corridor Between the Wetlands and the Sangre de Cristo Mountains .

2.

..............................................................................

..........................................................

5.

.................................................................................

6.

.........................................................................................

.

................

..........

..........................................................................

13.

Distribution of Folsom Localities in the Upper Rio Grande Basin .

14.

Log Scale Showing Relative Frequency of Folsom and Post-Folsom Artifacts .

15.

Grid System and Sequence of Excavation at Stewart’s Cattle Guard .

16.

...............................

................

57

58 60 64

.............................

67

Bison Bone Fragments and Associated Haminerstone/anvil in One-meter Square G1l2-A2.

..........................................................................

71

17.

Topographic Map of Stewart’s Cattle Guard Site .

...................................................

72

18.

Aerial Photograph of Stewart’s Cattle Guard Site .

...................................................

73

xvii

Figure 19.

Page Corner of Excavation Unit Showing Bioturbation in Lower Sand Unit Containing Folsom Component .

75

20.

Differential Weathering of Bone Surfaces .

78

21.

Graph Comparing the Relative Sizes of Tools and Flaking Debris .

22.

Physiographic Map Showing Geographic Features Mentioned in the Text .

23.

Possible Source Areas for Lithic Raw Materials at the Cattle Guard Site .

24.

Landform Map Showing Location of Key Folsom Locales .

25.

Range in End Scraper Weight and Length by Raw Material Type .

26.

Log Scale Showing Raw Material Representation from Block Sample .

27.

Frequency by Material Type of Core and Biface Reduction Flakes Largerthan 25 rum .

106

28.

Variation in Length of Black Forest End Scrapers .

107

29.

Variation in Length of Complete Black Forest Projectile Points .

30.

Variation in Maximum Length of Flake Tools by Material Type .

31.

Frequency of Retouched and Use-Modified Flake Tools by Material Typein Total Sample .

110

Map Showing Distribution of Selected Lithic Source Areas in the Rocky Mountains and Plains of North America .

122

Geographic Distribution of Raw Materials at Percentage Frequencies Greater than Twenty .

125

Raw Material Frequencies at the Cooper, Cedar Creek, Lipscomb and WaughSites

127

Raw Material Frequencies at Reddin, Zapata, Cattle Guard, Linger, La Manga and Sudberry .

129

Geographic Distribution of Raw Materials at Percentage Frequencies ofSix or less .

132

37.

Location of Block Sample, N26 to 80, E101 to 136

150

38

Cores and Wedge .

.....................................................................

...............................................................................

........................................

..........................

.......................

........................................

...............................

..........................

....................................................................................

..................................................................

..............................................................................

36.

..........................................................................................

.

104 106

109

.............................................................................................

35.

90

.................................

..................................................................................

34.

87

108

...................................................

33.

85

..................................

.................................................................................

32.

80

.................................................

........................................................................................

39.

Bifacial Core from Blackwater Draw, Mitchell Locality .

40.

Tools Made From Flake Blanks Possibly Removed From Bifacial Cores . xviii

...........................................

..............................

153 155 156

Figure

Page

41.

Manufacturing and Use Sequence for Folsom Preforms and Projectile Points .

42.

Complete Unfluted Folsom Points .

43.

Complete Fluted Preforms .

44

Preform Fragments .

.

....................................................................

..............................................................................

..................................................................................................................

45.

Bifaces Broken in Manufacture .

46

Fluted Points.

.

..................

........................................................................

.......................................................................................................

47.

Unmodified Flakes .

48.

Variation in Length of Complete Points and Preforms by Material Type .

49.

Folsom Projectile Point Tips and Bases. (Adapted from Jodry 1987:Figure 4.2)

50.

Folsom Point Breakage Patterns at Stewart’s Cattle Guard Site .

51.

Graphs of Percentage Frequency of Point Breakage Types .

52.

Oriented Photomicrograph of Hafting Wear Traces on Fluted Point Base .

53.

Oriented Photomicrographs of Hafting and Impact Wear Traces on Refitted Fluted Point .

.......................................................................................

........................

.......................

55.

Weaponry Retooling at Cattle Guard in Comparison with Other Folsom Sites .

56.

Refitted Ultrathin Biface (0104-Al-16 and 19)

57.

Oriented Photomicrograph of Refitted Ultrathin Biface (O104-A1-16 and 19)

58.

Distribution of Ultrathin Bifaces and Associated Artifacts .

59.

Ultrathin Biface Fragments: a, Cumbres chert (S 100); b, Morrison Quartzite (O106-A1-14).

.............................

.................

......................................................

..................

.........................................

...............................................................

62.

Views of Meat on Drying Racks in Kiowa Camps .

63.

Flake Knives: a) S134-A2-6 and b) N108-A1-32.

64.

Flake Knives: a) Gl12-A2-12, c)A112-A2-6, d) Al 14-A2-6 and Side Scraper: b) Z 16-Al-i. xix

169 171 176 178

184

Oriented Photomicrograph of Hafting Abrasion on Fluted Point Tip .

Oriented Photomicrograph of Ultrathin Biface (E106-A2-2) .

166

...................................

54.

61.

163

181

...............................................................................

Oriented Photomicrograph of Ultrathin Biface (S 100)

161

....................

........................................

60.

159

..............................................

.......................................

.................................................................

...................................................................

.................................................................................................

185 188

189 190 192 198 199 200

202 203 205 208 221

222

Figure 65.

Page Oriented Photornicrograph of Hard Material Cutting Wear Traces on Flake Knife (Gl 12-A2-12) .

...................................................................................................

223

66.

Distribution of Flake Knives and Side Scrapers in the Block Sample .

67.

Percentage Frequency of End Scrapers by Material Type .

68.

Variation in Discarded Endscrapers .

69.

Early Stage End Scraper Made of Heat Altered Cumbres Chert (Hi 18-A2-l).

70.

End Scrapers and Utilized Flake Recovered in Southwestern Work Area .

71.

Oriented Photomicrograph of Hard Material Wear Traces on End Scraper Used as an Adz (N108-Ai-l).

231

72.

Four End Scrapers with Hard Material Use-Wear Traces .

232

73.

Oriented Photomicrograph of Soft Material Wear Traces on an EndScraper (0104-Al-i).

233

The Distribution of End Scrapers (Triangles) and Use-wear Traces Consistent with Hide-Working (Circles) .

234

Oriented Photomicrograph of Hard Material Wear Traces. End Scraper Used on Bone or Antler (Ti20-Ai-l).

236

....................................

......................................................

........................................................................................

............................. .

230

.......................................................

.................................................................................

............................................................................................

76.

226 228

.....................................................................................................

75.

225

.......................

..........................................................................

74.

224

Hide Working Tools: Top, a) Beamer for Deer-sized Animals, b) Bone Chisel Flesher, c) Hafted and Unhafted Staking Tool Used to Stretch and Soften, d) Hand-Held Scraper e-I) Hand-held Stakers; Bottom, a) Bone Chisel Flesher, b) Bison Tibia Chisel Flesher From the Folsom Level at the Agate Basin Site, c) Modern Bison Tibia for Scale .

240

End Scrapers Hafted in Elk Antler Handles, National Museum of Natural History, Smithsonian Institution .

242

Graph Comparing Size of Discarded End Scrapers in Southwest Work Area With a Combined Sample From All Other Site Areas .

246

Gravers and Perforators: a) F 12-A2-2, b) El 12-A24, c) E106-A2-i, d) El 12-A2-1, e) C 14-A2-2, f) Dl 14-A2-5-1983, g) Dl i2-A2-5-1981, h) C 12-A2-7, i) SL 10 64

249

80.

Distribution of Gravers and Perforators .

252

81.

Flake Tools: a) Hii2-A2-7, b) W118-Al-1/Y114-Al-6, c) D1l6-A2-i, d) Yl12-Ai-2, e) Al 18-A2-2, 1) El 12-A2-15, g)Yii2-Al-2, h)Bl 14-A2-5, i)D108-A2-1.

253

Distribution of Flake Tools Modified by Use Only and Retouched Flake Knives andSide Scrapers .

254

.............

77.

.........................................................................................................

78.

.........................................................

79.

.....................................................................................................

..................................................................................

........................................................................................................................

82.

..................................................................................................................

xx

Page

Figure 83.

Distribution of General Areas of Activity .

........................................................................

263

84.

Distribution of the Bison Bone .

..............................................................................

266

85.

Distribution of Left Astragali used to Calculate the Minimum Number of Bison

86.

Distribution of Bison Bone Articulations .

87.

Distribution of Impacted, and Green Fractured Bone .

88.

Distribution of Bison Bone and Folsom Points in the Kill/Initial Butchering Area .

89.

Folsom Point Fragments Conjoined Between Kill/initial Butchery and Camp Areas

277

90.

Kill Area Projectile Points (K- Means Bone Cluster 2): a) S130-Al-1, b) T124-Al-1, c) S126-A1-1.

278

Kill Area Projectile Points (K- Means Bone Cluster 2, b-c): a) W 120-Al -1, b) P122-Al-3, c)Q122-A1-1.

279

Kill Area Projectile Points (K- Means Bone Cluster 6): a) W132-A1-1, b) W134-A1-1, c) V128-A1-2.

280

93.

Distribution of Bison Bone and Bone Cracking Equipment .

...................................................

282

94.

Density Contour Distribution of Flaking Debris .

.....................................................................

283

95.

Distribution of Six K-means Clusters, Lithic Tool Data .

96.

K-Means Bone Cluster Graphs: Top) Difference in Log 10 % SSE; Bottom) Log %SSE .

........

97.

K-Means Lithic Cluster Graphs: Top) Difference in log 10 % SSE; Bottom) Log %SSE .

........

288

98.

Bone Specimens Assigned to Six K-means Clusters .

...............................................................

289

99.

Lithic Specimens Assigned to Six K-means Clusters .

.............................................................

290

100.

Distribution of K-Means Bone Clusters and Piece Plotted Bone .

101.

Distribution of Refitted Hammerstone/anvils, Impacted and Green Fractured Bone Relative to K-Means Bone Clusters .

.................

................................................................................

................

........................................................................................................................

92.

272 274

.........................................

........................................................................................................................

91.

267

........................................................................................................................

............................................

275

284 287

............................................

287

....................................................

291

102.

Distribution of Stone Tools and Fragments Relative to K-Means Lithic Clusters .

103.

Distribution of K-Means Lithic Clusters and Clusters of Bone and/or Lithic Materials Identified Using Density Contour Methods .

296

Tools from Southwestern Work Area (K-Means Lithic Cluster K-i; Density Cluster 7; K-Means Bone Cluster K-3) .

298

..................

............................................................................

104.

.................................................................................................

xxi

295

Page

Figure 105. Folsom projectile points: a) Cumbres tip from kill/initial butchering area; b and c) Trout Creek points from the southwestern work area.

299

106. Tools From Residential Camp (K-Means Lithic Clusters K-2 and K-5) . ................................

300

107. Tools From Residential Camp (K-Means Lithic Clusters K-3) . ..............................................

301

108. Tools From Residential Camp (K-Means Lithic Clusters K-6, Density Cluster 2)

304

.................

109. Comparisons of Bifaces versus Unifaces, Weaponry versus End Scrapers and Ultrathins, and Hammerstone/anvils versus Utilized Cobbles in K-Means Lithic Clusters . .....................

305

110. Comparisons of Lithic Raw Material Types in K-means Lithic Clusters . ...............................

306

111. Distribution of Raw Materials in the Block Sample . ................................................................

310

112. Distribution of Refits in the Block Sample in Relation to K-Means Lithic Clusters . ..............

311

113. Distribution of Use-Wear Samples Studied by Ahler, Kay, and Tomenchuk . .........................

312

114. Distribution of Tools with Hard Material Use-Wear Traces . ...................................................

315

115. Distribution of Tools with Butchering and Meat Cutting Use-Wear Traces . ...........................

316

116. Distribution of Tools with Abrasive and Soft Material Use-Wear Traces . ..............................

317

117. View of Women Fleshing Buffalo Hides, 1878 Camp Near Mouth of Yellowstone River. ......................................................................................................................................

319

CHAPTER 1 THE RESEARCH PROBLEM

Introduction An abiding interest in Paleoindian stone tool technology and the hunting of now-extinct Pleistocene animals was set in motion in New Mexico in 1927 when a Folsom projectile point was uncovered amidst the ribs of ancient buffalo

(Bison bison antiquus).

Since that discovery a succession of

theoretical and methodological frameworks for investigating hunter-gatherer archaeology (e.g., B ettinger 1987, 1991; Jochirn 1976; 1981; Kelly 1995; Kroll and Price 1991; Thomas 1986; Walde and Willows 1991), and faunal (e.g., Binford 1981; Hudson 1993; Lyman 1994; Speth 1983; Stiner 1994; Todd 1987) and lithic assemblages (e.g., Binford 1979; Hayden 1979; Nelson 1991; Torrence 1989c) have been employed. Yet in many respects the behavioral dynamics which link Folsom people and technology to other aspects of their foraging life, such as social organization, food procurement and processing, and patterns of mobility, remain poorly understood. This dissertation seeks a greater understanding of these dynamics by examining how strategies for tool manufacture, use, and replacement and patterns of social interaction and residential mobility may have been interrelated in the context of cooperative bison hunting. Two explicit assumptions underlie this endeavor: 1) primary subsistence practices strongly influence patterns of social and technological organization, and 2) ecological circumstances hold important keys to the structure and variation of social and subsistence activities across the landscape, particularly for people with a foraging way of life. In keeping with this perspective the role of bison procurement and processing is singled out for investigation. Bison strongly dominate the faunal samples from excavated Folsom sites. This is due in part to the archaeological visibility of bison bone accumulations and the relative ability of dense skeletal elements to survive destructive taphonomic processes. This study aims to show, however, that the predominance of bison remains accurately reflects the relative importance of this species in Folsom

N

economic and social life. I agree with Jochim (1976:53) in reasoning that the greater the security and desirability of a subsistence resource the greater will be its ’pull’ on mobility and settlement patterns, and therefore its likely effect on social variables and technology. It is postulated that bison constituted just such a strong "pull" following the aftermath of Rancholabrean faunal extinctions and coincident with the expansion of bison habitat Another reason for choosing this particular research problem is a perceived need, and an excellent opportunity, to address organizational aspects of Folsom technology and social relations using site-specific data from Stewart’s Cattle Guard (CG) site. CG contains the archeological remains of a processing camp established where 49bison were procured in a single hunting event, or possibly a few closely timed hunting episodes. Cultural debris from butchery, bone marrow recovery, hide processing, and weapon tip manufacture and repair were recovered from extensive, contiguous block excavations. This is an appropriate assemblage with which to investigate the organization of activities accompanying mass bison processing and weaponry retooling. Seasonal mortality data from a small but informative sample of CG bison dentition indicates a late summer-early fall occupation(s) and provides the impetus for investigating the wider occurrence and ramifications of bison hunting during this season (Table 1) (see Todd et al. 1990:820-24). Until recently, dentally determined seasonality assessments for Paleoindian bison kill and processing sites primarily represented cool season kills made between late fall and early spring (Todd 1991:223, Figure 11.2). Most of these sites were late Paleoindian in age (post-dated 10,000 BP) and clustered geographically on the northwestern Plains (Casper, Finley, Homer, Carter Kerr-McGee, Agate Basin) and to a lesser degree, on the southern Plains (Lubbock FA6-1 1, Perry Ranch, Plainview). Conversely, early Paleoindian data were limited to a single Folsom (Agate Basin) and two Goshen (Mill Iron and Sheaman) sites from the northwestern Plains and the Murray Springs Clovis site in Arizona. Seasonality information was sorely needed for the early Paleoindian period from the Southern Plains and Southern Rocky Mountain regions. Reinvestigations of the Folsom, Lipscomb, and Lake Theo faunal assemblages by Todd and Hofman (Hofman 1995; Hofman et al. 1989; Todd et al. 1990, 1992, 1996) resulted in identification of a

warm season component to Folsom bison hunting strategies on the Southern Plains. Further, their analyses identified a larger number of animals than was previously recognized in these kills (Todd et al 1992:162). A model of late summer-early fall bison hunting during Folsom times received compelling support with the subsequent investigation of the Cooper site in Oklahoma (Bement 1999). Here three separate bison kill events took place in the same arroyo. All involved greater than twenty animals and all took place during the late summer-early fall season. These findings from assemblages excavated in the 1920s, 1930-40s, 1970s, and 1990s, respectively, demonstrate the long-term value of museum collections and the importance of (and continuing need for) site-specific empirical data in the construction of regional models.

Table 1. Dental Ages and Estimated Season of Death at Folsom Bison Kill and/or Processing Sites. Site IviNI NISP’ Dental Ages Season References Agate Basin

9

30

-N + 0.9

Cattle Guard

49

6

-N + 0.3 to 0.4

Linger

5

1d

N + 0.3 to 0.4

Folsom

54

33

N + 0.4 to 0.5

Late winter to early spring Late summer to early fall Late summer to early fall Early fall

Lake Theo

12-14

7

N + 0.4 to 0.5

Early fall

Lipscomb

55

18

N + 0.2 to 0.5

Frison 1982c:258 Zeimens 1982:227 This study This study Todd et al. 1996: Table 8.15 Harrison and Smith 1975:16; Todd et al. 1996: Table 8.15 Todd et al. 1990:141-144; Todd et al. 1996: Table 8.15 Bement 1999:127

Late summer to early fall Cooper 29 N + 0.3 Late summer upper kill to early fall Cooper 29 N + 0.3 Late summer Bement 1999:127 middle kill to early fall Cooper 20 N + 0.3 Late summer Bement 1999:127 lower kill to early fall aNumber of individual specimens used to determine seasonality. bSingle calf mandible and two near-term fetal bison. ’Five lower M3 with differential cusp wear and one partially erupted, unworn lower M 2 (Group 2 ENS 1]; Group 3 [N=4]; Group 4 [N1]). dSingle mandibular M’

There are now eight Folsom components representing a seasonal pattern of bison hunting in the late summer or early fall on the Southern Plains (Lipscomb, Folsom, Lake Theo, and Cooper sites) and in the Southern Rocky Mountains (CG and Linger sites). Like the original discoveries at the Folsom type site, CG comes into clearer focus when viewed within a broader pattern of large-scale, cooperative bison hunts. It is toward a greater understanding of this seasonal subsistence activity that this dissertation is directed.

4

The viability of bison hunting as an economic mainstay required reliable access to foraging areas sufficient in size to overlap herd movements as they varied in time and space. This ensured commitment to high rates of residential mobility, at least in some seasons. It is generally acknowledged that mobility effects the structure of lithic assemblages in fundamental ways, however the dynamics involved in particular circumstances require further study. The aim here is to explore how 1) planning depth, 2) high rates of residential mobility, and 3) time constraints imposed by the need to process meat, marrow and hides, en masse, in a relatively brief period of time may have influenced technological and social strategies. Technological strategies involved the procurement of toolstone, and the design, production, use, and repair of portable, effective gear. Social strategies entailed the development and maintenance of kin, alliance, and information networks and quite likely gendered divisions of labor. An economic focus that depends heavily on pedestrian hunting of mobile prey such as bison is inherently risky. The uncertainty involved in such an enterprise arguably places a premium on developing multiple strategies to offset unpredictable shortfalls. Maintenance of far reaching social alliances and information networks, monitoring and utilizing alternate foods, and adopting a flexible stance with regard to residential mobility are examples of such strategies. Some researchers emphasize the role of social alliance in Paleoindian adaptations (Hayden 1982:117-120; MacDonald 1999:148; Wilmsen 1973:25-26). Others view Folsom groups as relatively isolated populations who lacked an intimate familiarity with the landscapes and resources over which they traveled. With regard to the latter, Kelly and Todd (1988:23 1, 239) postulated that frequent shifts in territorial range coupled with low population densities precluded both social alliances and detailed knowledge of habitat, especially plants. Similarly, Abler and Geib (1999:26) recently proposed that Folsom groups possessed a limited knowledge of lithic resources. This study envisions that Folsom individuals and groups were well acquainted with resources (both social and natural) across large regions. In keeping with perspectives presented by Hayden (1982) and Weissner (1977), it is suggested here that Folsom people engaged extensively in direct and down-theline communication within and between regions as an adaptive form of social insurance, or risk pooling. Organizational strategies are predicted to have been favored during Folsom times that: 1) encouraged groups to share access to key resources, 2) facilitated the ready movement of people between geographic

5

areas and socialgroups, and 3) increased the probability that aid between groups would be rendered and reciprocated as needed. Cooperative efforts within and between local groups are herein viewed as a common and frequent part of Folsom adaptations. In addition to the ramifications for social organization, seasonal subsistence activities that included frequent residential moves and intensive bison kill and processing events established fundamental technological parameters for toolkits. A great deal of planning depth is evident in Folsom strategies for raw material procurement and the subsequent design, production, and maintenance of tools. This dissertation examines technological strategies employed in the particular circumstance of a large bison kill situated away from irmnediate sources of quality toolstone. Almost all lithic provisioning was done ahead of time, non-locally. The following section briefly reviews the theoretical and methodological basis of studies of technological organization and relates this to the research discussed in the following chapters.

Approaches to the Study Lithic technologists have made a conceptual shift from technological studies concerned with describing the form, function, and production techniques of stone tools to explicit attempts to understand what this variability means with regard to hunter-gatherers who juggle social and subsistence options in different situations. In trying to understand why particular assemblages look the way they do, researchers have investigated the role of differences in the seasonal and/or spatial availability of food (Amick 1994a), toolstone (Andrefsky 1994; Ellis 1989; Hofman 1992; Monte-White and Holen 1991), and mates (MacDonald 1998b; Wilmsen 1973; Wobst 1974; 1976). Additionally, the effects of risk (Bleed 1986); mobility requirements (Binford 1979; Kelly 1988, 1995; Shott 1986a, 1989), time stress and energy costs (Bleed 1986; Torrence 1983), and engendered patterns of tool manufacture and use (Boismier 1991:211; Gero 1990; Jodry 1998) have been considered. Nelson’s (1991) informative review of these issues examines the complex relationships that exist among these variables and their combined effects on archeological assemblages. Kelly (1994:134) proposed that regional scale studies will prove to be "more real and meaningful" and "less precise, but more accurate" than single-site studies, because they effectively ’swamp" sources of variability and allow large-scale patterns to emerge. In keeping with this prediction, regional studies have

rel

proven to be of considerable value in documenting variability in Folsom hunter mobility across the southern Plains (Amick 1994a, 1994b; Hofman 1991, 1992), Basin and Range (Amick 1994a, 1994b; Judge 1973), and Southern Rocky Mountains (Jodry 1999; Stanford 1991). However, due to interpretive constraints regional analysis of surface assemblages by necessity focus on projectile points, preforms and channel flakes (Amick 1994a:99-100; Bamforth 1988:166-167; Hofman 1991:336) for which minimal information is available regarding the behavioral dynamics responsible for site formation. As a result, it is difficult to assess the roles played by social frameworks and task applications in the overall organization of technology (see Boismier 1991:189-199; Gamble 1991:7-11). It is assumed here that Folsom points and preforms better represent the activities of male hunters and/or flintknappers than the activities of females, young children and non-hunting males. Therefore variability in the activities, mobility, and patterns of lithic procurement for large segments of Folsom populations are not directly addressed in most regional studies. We currently understand little regarding the ways in which variation in weaponry relates to variability in other portions of lithic and faunal assemblages. Nor are we able to assess much better the meaning of differences in tool composition among surface assemblages from different sites. Shott notes the following regarding diachronic regional comparisons of site assemblages: it is necessary to control the nature of task applications in which technologies are used, in order to observe variability in technological organization. This can only be accomplished by showing that the same activities characterized the systems under study. In other words, if we are to attribute differences between the technologies to the postulated differences in settlement mobility, it is necessary to control for differences in the nature and range of task applications which characterized the respective systems. (Short 1986a: 148) This is also true when comparing assemblages from sites occupied by members of a single culture. The following excerpts give voice to widely recognized concerns regarding the study of technological organization, be it among prehistoric or modern hunter-gatherers. Several ingredients lead to strengthening inferences about technological organization. An accurate reconstruction of the behavioral context of tool use is the foundation of organizational studies. (Amick 1994a:9) . . the location and timing of the use of technology, the manufacturing or otherwise securing of tools, their maintenance, recycling, discard and loss, will also be related to the activities of which they are a part. If the effects of raw material can be held constant then the residual patterning may be related to behavioral variables. (Kelly 1994:133)

7

Context more than form is required to interpret important design characteristics of past material culture. (Greaves 1997:20) understanding tool use requires not only a focus on manufacture, design, and use of tools, but linking technological systems to other behaviors the study of technological organization .. should address both the planned use of tools and the effects of their participation in multiple systems of use and discard. (Greaves 1997:347) ...

Just as there is general agreement that the contexts of tool-related activities are critical to their interpretation, so there is broad appreciation for the difficulties inherent in reconstructing contextual data. It is typically laborious and sometimes impractical to assess the relative contributions which social frameworks and activity orientations make to the nature of lithic assemblage variation and deposition at Paleoindian sites. This is especially true in cases of limited preservation of material culture, dietary remains, and hearth and structural features. Yet at sites where there is a productive opportunity to do so, the efforts can be well rewarded (Bamforth 1991; Bement 1999; Deller and Ellis 1992; Frison 1982a; Hofrnan et al. 1989; Larson 1994; MacDonald, G. 1969; Reber and Frison 1980; Root and Emerson 1994; Stanford 1978; Storck 1997; Wheat 1972). Frison (1967) convincingly illustrated the importance of site-specific data for regional studies some years ago when he examined the maximization of time, effort, and raw materials with regard to tool manufacture and use at the communal bison-processing sites at Piney Creek, Wyoming. He fruitfully combined single-site, regional and ethnographic data to reconstruct aspects of protohistoric Crow economy on the northern Plains. Wheat (1978) successfully used detailed excavation data from two bison procurement and processing sites to compare the ways in which activities and length of occupation effected the character of individual tool assemblages and created variability in patterns of regional settlement. Recently, Dillehay (1997) provided an unusually detailed account of the organization of activities and tool use at the site of Monte Verde in Chile and used this information to model early Paleoindian land use strategies in the Southern Hemisphere. As summarized by Thomas: The challenge for archaeology is quite clear. Adequate theory cannot rest merely on a global, hologeistic approach. . . we must supplement [this approach] with solid empirical case studies. . . [and] develop alternative research approaches to emphasize the range of coping behavior, rather than merely isolating the modal, focal tendencies in huntergatherer exploitative strategies. (Thomas 1983a:17, as quoted in Bamforth 1991:218)

8

Clearly the close examination of single sites as well as regional syntheses of surface and other assemblages are needed (Kelly 1994:134). Indeed they are two methodological sides of the same analytical coin. The goal of this dissertation is to combine case-study information with regional ecological and archeological data to model the ways that cooperative bison hunting and processing, high residential mobility, and patterns of social interaction, raw material procurement, and tool use influence one another. I explicitly acknowledge a theoretical preference for ecological models of subsistence practices. The framework of behavioral ecology provides heuristic insights for modeling bison hunting strategies under different circumstances of aggregation, dispersal, and predictable availability and for relating this information to mobility, technology, and cooperative behavior (Kelly 1995:109).

At the same time, I

recognize that social and spiritual motivations are powerful and decisions made on these grounds are not likely to be as sensitive to testing with such models (Jochirn 1998:23).

Organization of the Dissertation In the following chapters I develop the proposition that the Folsom Cultural Complex represents a suite of interrelated social and technological strategies that buttress an economy focused on bisonprocurement. It is proposed that this subsistence orientation developed in response to a dramatic change in biotic productivity and large animal resource structure during the Younger Dryas Climatic Period (nearly 11,000 to 10,000 14 C yr BP). Because the development of bison hunting economies among Folsom and Goshen Paleoindians was coincident with and arguably strongly tied to changing ecological circumstances, Chapter 2 presents detailed paleoecological information. These data provide a necessary framework for understanding Folsom cultural patterns in relation to variation in the natural world. A net trend toward more effective moisture and cooler temperatures during the Folsom time period are discussed. This picture emerges from studies of the changing distributions of plants and small mammals adapted to cool, moist habitats, and from geologic evidence of former water regimes. Two mechanisms are proposed that may have significantly increased the net level of nitrogen (i.e. fertilizer), and by extension forage quality, in the Southern Rocky Mountain Region during the Younger Dryas interval. These are 1) changes in the nature of alpine snow packs and 2) increases in the size and grazing intensity of bison populations. I suggest that nitrogen-enriched forage in combination with a longer somatic growing

season (Guthrie 1984:269) may have set positive feedback loops in motion that led to improvements in net bison condition and fertility (Frison 1991:9) during early Folsom times. Studies of bison skeletal evidence confirm that Folsom-age animals were larger than their Holocene counterparts. As open sedge and grassland habitats expanded during the terminal Pleistocene, and competitive grazers declined due to Rancholabrean extinctions, bison are presumed to have experienced ecological release and population increase. It is postulated that this change in resource structure favored the development of specialized bison hunting economies on the Plains and in the adjacent intermontane basins of the Rocky Mountains during the early portion of the Younger Dryas climatic period. Following this discussion of regional climatic variation in Chapter 2, I turn in Chapter 3 to a description of the geographic setting, archeological investigations, and site formation processes at CG. The site lies within a natural grazing and travel corridor located between mountains and wetlands. This linear expanse of sedge and grass meadows lured and concentrated bison herds as they moved along the valley’s eastern side, creating ample hunting opportunities and placing people in proximity to a wide array of other resources. The heart of the dissertation, from the perspective of material culture and social and technological organization, is presented in Chapters 4, 5 and 6. Chapter 4 discusses where CG may fit within overlapping networks of regional land use and social interaction. Lithic raw material utilization provides proxy data for the movements of people on both local and inter-regional scales. The fourth chapter begins by discussing the sources and relative frequencies of chipped stone materials at CG. Raw materials are compared in terms of relative "wear and tear" along trajectories of use and resharpening within and across tool classes. These data address the relative length of time different raw materials may have been represented in the composite assemblage prior to its deposition at CG. Next, the wider Folsom landscape is considered in relation to the possible role of kinship and alliance networks as social mechanisms that mitigated economic shortfalls and provided cooperative labor when bison were procured and processed en masse. The relative occurrences of sixteen material types at forty-five Folsom sites are mapped in relation to primary source locations. The distributions of toolstones at individual sites (in >20% of the total lithic assemblage) suggest recurrent land use patterns that connect

IN

people along the Front Range and southern intermontane basins of Colorado with groups in the San Juan Basin and Southern Plains. Likewise a seasonal pattern of movement from the Edwards Plateau to the Panhandle Plains of Texas and Oklahoma is strongly evident. The distribution of weaponry-related artifacts made of exotic raw materials (those comprising

..4f

ARTESIA’I ,OJIFER

l k

)-1

: SANTA FE FORMATION VOLCANIC FLOWS,TUFFS AND BRECCIAS

AD

-/ "-

1/

c:..:

::::.

.. -’

I

v-

-. I

2

::)_\IK

_

-II

-)1c-’

f’/

’\ %I7

...

:::::::

(OLIGECENE)

/

/’,ii-’-((/’’,

2,

/?(/ N"

- /

’J’’

Ii

I’-

Data sources: Dixon 1971, Emery et a! 1971, McCalpin 1983, Powell 1958, Tweto 1979.

Figure 4. Schematic cross-section of the central San Luis Valley showing geology, hydrology and recent vegetation zones.

5 lIen.

21

(3 100 m) in the eastern Baca graben, and thinnest over the mid-valley Alamosa horst (250 m) (McCalpin 1996). Emery et al. (1971:129) estimated that two billion acre-feet of water were stored in the upper 1 829 in of sediments within valley-wide systems of unconfined and artesian aquifers. Crouch (1985) described a decrease of about 880,000 acre-feet from December 1969 to January 1980 due to irrigation and other water diversions. Legal battles over rights to water in the Baca Graben are a hotly contested political and economic reality in the region (Bingham 1996:135-152). Despite modern aridity depth to the water table is generally less than 3.5 m for much of the valley (Emery et al. 1971:130). This shallow water table greatly influences the character of the basin in that minor fluctuations in its level produce dynamic shifts in the relative proportions and composition of its wetlands, and in the amount of standing water created during spring runoff. Valley margins, in particular, are strongly affected by annual changes in water levels due to snowmelt entering the grabens (McGowan and Plazak 1996). Other features of importance are geothermal springs associated with faults at a number of locations and small streams connected to tectonic sources of warm water that do not freeze during the winter, even at temperatures below -5 C. Modern Climate The valley’s distinctive climate is influenced in large measure by its shape, high altitude, and rain shadow position. Steeply surrounded by mountains, the valley acts like a bowl to collect cold air moving down the mountain slopes where it becomes trapped as low as 15 in above the ground surface by overlying warmer air. This cold air lingers until strong winds push it south, past the San Luis Hills into New Mexico. Frigid temperatures (down to -10 C) result from Pacific Fronts that blanket the valley in snow. Most snow, however, falls in the surrounding mountains, leaving the valley floor open for much of the winter. The San Juan Mountains typically get more snow in a single day than the valley center receives during the entire winter (Doesken and McKee 1989:81) (Table 2). Winter temperatures are mitigated by the fact that valley winds are usually light, and the sun normally shines (330 days annually). Basin temperatures are relatively cool and average 6.4 C annually. This is comparable to Williston, North Dakota; Kalispell, Montana; Burlington, Vermont; and Kodiak, Alaska. Over the course of a year, the valley experiences diurnal temperature variations of more than 17C. Mean July temperatures for the period 1947-1987 (at approximate N latitude 37 40’; longitude 106; elevation 7560

22

m) range between daily highs slightly above 26.6C, to nighttime lows of 10 C. The growing season is approximately 142 days in the vicinity of Stewart’s Cattle Guard site. Along the river and valley center the growing season decreases to 107 days, due to cold air drainage (Doesken and McKee 1989:86).

Table 2. Comparative Precipitation Data from the Upper Rio Grande Basin. Station Elevation Setting Period of Mean Annual m Record Precipitation cm Summitville 3459 San Juan Mtns. 1939-1947 91.28

Wettest Year cm 126.69

Driest Year cm 85.62

Wolf Creek Pass

3243

San Juan Mtns.

1957-1987

112.44

172.61

75.23

LaVetaPass

2816

Sangre de Cristo Mtns.

1931-1953

54.61

88.74

25.12

Great Sand Dunes

2475

East Foothills

1950-1987

26.06

43.07

14.85

Crestone

2473

East Foothills

1982-1987

37.01

41.93

30.50

Del Norte

2402

West Foothills

1893-1987

23.67

42.79

12.09

Saguache

2347

West Foothills

1894-1987

21.48

41.19

9.55

Manassa

2341

South Valley

1931-1987

18.69

30.73

7.54

Center

2341

Mid Valley

1941-1987

17.80

28.09

9.49

Alamosa

2295

Mid Valley

1948-1987

18.00

29.33

8.63

Source: Doesken, N. J. and T. B. McKee, The Incredible Climate of the San Luis Valley. Colorado Groundwater Association, 1989, adapted from Table 1, P. 84.

Denver:

Currently, the majority of precipitation on the valley floor occurs during thunderstorms generated by southerly inflow of Gulf moisture during July and August, the only frost-free months. The spatial distributions of temperature and precipitation mirror the strong altitudinal gradient of local topography. The greatest percentage of rain and snow falls on and above the foothills. This, together with the presence of numerous spring-fed mountain streams, keeps the valley margins better watered. Thus, a number of wet meadows exist here. These provide especially favorable grazing areas for modern herds of bison, cattle, and elk (United States Department of the Interior, Bureau of Land Management [USD1, BLM] 1977:11-68). Judging from the distribution of Folsom sites, bison and other animals were regularly hunted in these meadows during the mesic Younger Dryas period.

23

Younger Dryas Climatic Oscillation Following the Wisconsin Glacial Maxima at 18,000 BP, a warming trend led to rapid de-glaciation in many parts of the world. Global warming was interrupted, however, between 11,000 and 10,000

’4

C BP

by a return to colder conditions during the late-glacial Younger Dryas (YD) oscillation. The trigger for this climatic interval is thought to lie in a disruption of thermohaline circulation patterns in the North Atlantic Ocean, related to the melting and calving of the Laurentide and Scandinavian ice sheets (Broecker 1994:421-22, 1990:459-67; Broecker et al. 1988:1-3; Manabe et al. 1995: 165-67). Recent studies of Greenland ice cores (Alley et al. 1993:527-529; Dansgaard etal. 1993:218-220; Mayewski et al. 1993:195; Taylor et al. 1997:825-827), corals (Edwards et al. 1993:962-3), and lake varves (Stromberg 1994:177) indicate a rapid onset and termination of the YD. While this climate change is thought to have been experienced world wide (Alley et al. 1993:527-52), paleoenvironrnents were variably affected from one region to another. For instance, in the Canadian and Colorado Rockies the YD led to alpine glacial advance that appears to be synchronous with the YD event in northwestern Europe, although the magnitude of cooling was significantly less in North America (Reasoner et al. 1994:442; Menounos and Reasoner 1997:45). Conversely, Clark and Gillespie (1997:21) determined that the YD did not result in glacial advance in the Sierra Nevada of California. Along these lines, Grigg and Whitlock (1998:296-97) concluded, from palynological studies in Western Oregon, that the YD event affected North Pacific coastal sites to a greater degree than inland locations. These differences underscore the need to independently assess the effects of the Younger-Dryas in any given study area. In order to model Folsom adaptations we need ever more refined reconstruction of the ecological conditions effecting terminal Pleistocene resource structure. The millennial scale of the YD Cold Episode further dictates the need for high resolution dating methods to register short-term changes (cf. Haynes 1993). For these reasons, I initiated the coring of Black Mountain Lake in the upper subalpine forest of the San Juan Mountains to provide region.

a record of late-glacial vegetation change for the Rio Grande headwater

This study was a collaborative effort by researchers from the Smithsonian Institutions

Paleoindian/Paleoecology Program, the University of Colorado’s Institute of Arctic and Alpine Research, and the Rio Grande National Forest (Reasoner and Jodry 1998a, 1998b). Corroborative insights are

24

provided by lacustrine cores from the San Luis Valley and Sangre de Cristo Mountains, investigated by University of Arizona researchers (Shafer 1989; Davis and Shafer 1991; De Lanois 1993).

Palynological Studies in the Upper Rio Grande Basin Introduction Three lacustrine sediment cores register changes in late Pleistocene vegetation and climate for the Upper Rio Grande Basin. These palynological records transect the San Luis Valley from the upper subalpine forest of the eastern San Juan Mountains to the alpine/forest ecotone of the western Sangre de Cristo Mountains. From west to east they are: 1) Black Mountain Lake in the San Juan Mountains (Reasoner and Jodry 1999), 2) Head Lake (Shafer 1989; Jodry et al. 1989; Davis and Shafer 1991) on the intervening basin floor, and 3) Como Lake in the Sangre de Cristo Mountains (Shafer 1989, Jodry et al. 1989) (Figure 5). Collectively they signal shifts in the positions of alpine and foothill timberlines, and variations in basin lake levels and plant communities. The data from Black Mountain Lake, investigated as part of this dissertation research, is discussed first, followed by summaries of the results of the Head and Como Lake studies.

Black Mountain Lake Black Mountain Lake is situated at 3413 in elevation, approximately three km southeast of the Continental Divide, west of Creede, Colorado. It was cored in conjunction with investigations at the Black Mountain site (5 FIN 55), the highest altitude Folsom hunting camp (3096 m) yet excavated in North America (Jodry et al. 1996:26). The lake’s position, nine km up-valley and 317 in upslope of this campsite, is well placed to record vegetation shifts that may have influenced the activities of Paleoindian people in the area. In particular, its setting some 187 in below alpine timberline made it sensitive to seasonally severe temperature fluctuations which affect the shifting position of the tundra-forest boundary through time (Fall 1997:1315). The lake (less than eight in deep and five hectares in area) currently lies within the upper subalpine spruce-fir forest (Ficea engelrnannii/Abies lasiocarpa). Table 3 provides modern climate data for the area, as recorded at Summitville (3442 in), in a similar subalpine setting, and at Hermit Lakes (2743 m) lying in a cold-air drainage in a montane valley bottom.

4 11

?

ffla 4, Ali

4

I

P

Figure 5. Physiographic map of south-central Colorado showing the locations of palynological studies.

Mel

Table 3. Modern Climatic Data for Summitville and Hermit Lakes, Upper Rio Grande Basin Summitville Hermit Lakes 1939-47 1951-84 Mean Annual Winter Temperature

-8.3 C

-10.5 C

Mean Annual Summer Temperature

8.8 C

12.2 C

Mean Annual Precipitation

99 cm

38 cm

Mean Annual Winter Snowfall

871 cm

198 cm

Source: U.S. Forest Service, Soil Resource and Ecological Inventory of Rio Grande National Forest - West Part, Colorado, 1996, Chapter 5, Table 1.

Two sediment cores, 7.6 cm in diameter and a maximum of 780 cm long, were collected through winter ice with a vibracorer in January 1997 (Reasoner and Jodry 1998b:3). A continuous depositional record spanning the last 10,000 years is represented by massive to faintly laminated, organically enriched deposits in the upper 261cm. The contact between thisgyttja and underlying inorganic, silty-clay sediments is sharp. This distinct boundary marks a strong contrast between the relatively high wet bulk density with high magnetic susceptibility, low organic carbon content and low water content in the lower core (262-740 cm), and the reverse characteristics in the upper core (0-261 cm) (Figure 6). A fir needle (Abies) recovered two cm above this boundary provides an AMS date of 9,930 – 70 BP (CAMS-38702). Analysis by Reasoner was directed at the core interval between 228 and 293 cm, which spans the period from 11,430 – 70 to 8,480 – 80 BP (Reasoner and Jodry 1998a; 1999). Chronological control for the core is based on 19 AMS radiocarbon determinations (Table 4). At six levels in the core between 228 and 293 cm, terrestrial macrofossils were dated. In four cases, a terrestrial macrofossil age was paired with an age determined from extracted humic acids to test the reliability of humic ages for the entire record. The close correspondences between the two sets of data are seen on the right side of Figure 6. With this reliability established, additional humic ages were determined for portions of the core for which plant macrofossils were unavailable. A humic date of 25,060 – 360 (CAMS-47409) from a depth of 775 cm provides a basal estimate for the lower core. The core did not intersect the bottom of the lake, but the resistance of sediments to coring beyond 780 cm may suggest we were close.

d

16

"C Years BP C)

&o

C)

C)

C)

C) C)

CD 140

5260 150

5

_1

15

25(10 7m/g)

180

I

8910–50

:::::70

9370–70 9910–60

9930–70

10,350–50

-

C)

C)

CD

CD CD C) Co

-

C) CO

C) C)

-

CD

C) C)

C)

--

HuicAcid m Age ----------

Terrestrial Macrofossil Age

I

80

CD CD

+ --

I

4980–70

Lt)

C) C)

220

Magnetic Susceptibility

260

10,840–50 11,640–50

11,430–70 300

1i Massive to faintly laminated silty clay

Laminated silty clay

Figure 6. Magnetic susceptibility Black Mountain lake core.

W

Rhythmically laminated silty clay

N..)

28

Table 4. Uncalibrated AMS Radiocarbon Dates from Black Mountain Lake Core BML 97-2, Lab Number Core Depth Material Dated 14 C Date Delta C cm Humic Acids CAMS-48144 37 Humic acids -26.5 1960 – 50 CAMS-47406 59.5 Hurnic acids -24.1 2,090+50 CAMS-38698 160 Picea needle 4,980 – 70 CAMS-38697 160 Humic acids -28.8 5,260 – 50 CAMS-38706 228 Picea needle 8,480 – 80 CAMS-38700 238 Picea needle 8,890 – 70 CAMS-38699 238 Humic acids -28.0 8910+50 CAMS-38705 253 Picea needle 9,370+70 CAMS-38702 259 Abies needle 9,930+70 CAMS-38701 CAMS-45519 CAMS-45520 CAMS-38704 CAMS-38703 CAMS-47407 CAMS-44853 CAMS-44852 CAMS-47408 CAMS-47409

259 274 283 293 293 347.5 440

Humic acids Humic acids Humic acids Conifer twig Humic acids Humic acids Humic acids

581 601 775

Humic acids Humic acids Humic acids

-30.2

9,910 – 60 10,350 – 50 10,840 – 50 11,430 – 70 11,640 – 50 15,660+ 120 16,040 – 70 17,060 – 70

-30.0 -23.5 -23.8 -21.9

16,200+50 25,060+360

-16.8 -23.5

-26.8 -26.7

Palynological interpretations (Reasoner and Jodry 1999) are based on 36 pollen samples, collected at intervals of 2.5 cm and less, from the late-glacial portion of the core. The mean grain count is 660 (range: 488-949). Pollen percentage data are used to reconstruct the latest Pleistocene to early Holocene vegetation history. The down-core variations in the major taxa are shown in Figure 7. Reasoner subdivided the pollen record into four zones on the basis of a strati graph ically constrained cluster analysis (CONISS; Grimm 1988). The cluster analysis draws major dissimilarity partitions at approximately 10,085 BP, 11,130 BP, and 11,620 14C yr BP. The ages of 10,085 and 11,130

14C yr are very close to the

termination and initiation, respectively, of the Younger Dryas climatic oscillation (Alley et al. 1993:527529; Taylor et al. 1997:825-827). High-elevation arboreal taxa decline during the Younger Dryas and increase again about 10,100 14C yr BP. Herbaceous taxa show the opposite trend. As the site is located in high mountainous terrain, these pollen changes likely represent fluctuations in the elevation of alpine timberline.

BLACK MOUNTAIN LAKE, COLORADO: PRELIMINARY PERCENTAGE DIAGRAM LATE-GLACIAL TO EARLY HOLOCENE Trees and Shrubs

Spores

Herbs

0

OA

ell

0)

0 8480–80 8890–70

ci

230 240 250

9870–50 a 9370–110 a 9910–70 9930–60

I

kV

1

. 1f.

260

10,120–50 10,350–50 270 280 10,840–50 14,30–70 290 300 310 2040 20

% gyttja

E

Massive to poorly laminated - silt and clay

Figure 7. Preliminary pollen diagram from Black Mountain Lake, Colorado. Prepared by Mel Reasoner.

CONISS

30

Inferences of vegetation history are based on the following interpretations by Reasoner of relative pollen percentages in Zones Ito 4 (Reasoner and Jodiy 1998b:5-6; 1999).

Zone 1(315 -296.5 cm) Sagebrush-Pine-Grasses Assemblage The top of this zone dates 11,570

14 C yr BP and the base is undated. Total arboreal taxa increase

over Zone I from 22% to 32% of the total pollen in the zone.

Artemesia, (Sagebrush, 34 to 44%) and Pinus

(pine, 18 to 26%) pollen dominate, with Poaceae (grasses), and Picea tubuhjlorae (high spine) represented at percentages from 5 to 10%. Pollen from nonarboreal taxa such as (goosefoot/amaranth),

Chenopodiaceae/Amaranthaceae

Rosaceae (rose family), Ambrosia (ragweed), Cayophyilaceae (pink family),

Cupressaceae, Cyperaceae (sedges) and Salix (willow) occur in percentages from 1 to 5%.

Thalictrum

(Meadow Rue), increases over Zone I from trace amounts to 3%. Also present in trace amounts throughout Zone 1 are Alnus (alder), Polygonaceae (i.e. smartweed),

Umbell(ferae (parsley family), Botrychium

(aquatic algae), Diyopteris-type, Pteridium (ferns), and spores of Selaginella densa (rock selaginella).

Zone 2 (296.5 - 285.5 cm) Pine-Sagebrush-Spruce Assemblage Zone 2 dates from approximately 11,570 to 10,900

14C yr BP. The significant feature of Zone 2 is

the increase of total arboreal taxa from 32% to 45% of the pollen sum. Increases in pollen percentages of Picea, Pinus, Thalictrum, and Pteridium, and declines in Artemisia, Rosaceae, and Tubul?forae high-spine

occur at or near the base of this zone. Abies (fir), Quercus (oak), Cupressaceae, Salix, Chenopodiaceae/ Amaranthaceae, and Ambrosia pollen, as well as spores of Seleginella densa occur consistently throughout

the zone at 2-5%.

Alnus, Caryophyllaceae, Potentilla (cinquefoil), Polygonaceae (i.e. bistort),

Uinbell(ferae, Botrychiuin, and Diyopteris-type palymorphs are present in trace amounts.

Zone 3 (285.5 - 262.5 cm) Sagebrush-Pine-Grasses Assemblage Zone 3 dates from -10,900 to 10,140 14C yr BP and overlaps with the Folsom time period. Total arboreal pollen percentages decline sharply from 45% to remain between 28% and 32% for most of the zone. This change largely reflects decreases in pine, spruce, and fir pollen percentages and increases in sagebrush pollen percentages, suggesting a shift toward steppe conditions around the lake. Throughout the zone, pine and spruce pollen are present in consistent percentages of nearly 25% and 8%, respectively.

31

Oak and fir pollen occurs in steady, but low (20% (Figure 33). Although these tool stones occur as exotics at the CG and Reddin sites, they comprise 28% and 46%, respectively at Linger and Zapata. These latter percentages might be inflated due to small samples sizes (Zapata N=28; Linger N=18). However the fact that Southern Plains materials were recovered at all four Folsom localities excavated to date and are persistently represented among surface assemblages supports a strong inter-regional connection (Jodry 1999). Raw material frequencies at these sites and two others located along the eastern flanks of the Rocky Mountains in northern New Mexico are shown in Figure 35. The percentages of Edwards and Alibates in the latter sites are based on nearly 300 weaponry-related surface artifacts from the sites of Sudberry (n=153) and La Manga (n= 145) (Table 51, Appendix A). Repeated movement and social interaction between the San Luis Valley and the Southern Plains along this route may be suggested. Additionally, exchange may have dispersed some tool stock beyond the areas actually traversed (Sampson 1988:19; Wiessner 1977).

129

V



H

r

1jJ

it

5%

-

Reddin

!-

Cattle Guard

1i

H

-6!

-

-.

46%

Zapata

N;

It

-----

H

v

Linger

7

y

-

30 0)o

N 9! La Manga

Sudberry

r

Figure 35 Raw material frequencies at the Reddin, Zapata, La Manga, Cattle Guard, Linger and Sudberiy sites.

-

130

The apparent dominance of hunting gear over other tool types made of Edwards or Alibates leads one to believe that the San Luis Valley was a peripheral hunting area for people from the Southern Plains. This axis of travel and social contact between north Texas/Oklahoma and northern New Mexico/Colorado continues to this day during hunting season, perhaps for broadly similar reasons. The San Luis Valley and adjacent environs to the south comprise some of the closest Rocky Mountain country accessible to people residing to the southeast. The beauty of this high altitude region and the richness of its game resources are undeniably part of their appeal. The intermontane basin of San Luis seasonally offered a well-watered tract where grassland game species, mountain sheep, foothill plant resources, marshes and waterfowl were all available in a spatially compressed area. Its mosaic habitats included prime bison hunting grounds. The upper Rio Grande basin is simultaneously a western extension of the Plains and an eastern gateway to the mountains. Historically it was something of a crossroads linking the territories of adjacent social groups. Folsom people equipped with tools made of lithic raw materials from different areas evidently came to the CG and Linger sites in the San Luis Valley in the late summer/early fall to hunt bison and probably to do other things for which we have a less distinct record. The known distribution of lithic source areas suggests that some of the people who camped at CG may have been dispersed in different directions prior to that bison kill. Some people may have been moving along the upper Arkansas River and South Park and perhaps in the vicinity of the Front Range. Others families and/or individuals apparently arrived at the summer bison hunt from the southwest. These people may have been closely related families whose ranged in dispersed groups over an extensive residential area and who got together socially during the summer season of abundance. The representation of Chuska chalcedony among a variety of tool types suggests a pattern of residential mobility for some people out of the southwest. The more restricted context of male hunting gear for Edwards and Alibates cherts suggests a logistical pattern of far-reaching travel and/or social contact from the southeast. While the particulars are unknown, the inferred directions of travel strongly suggest that an aggregation of people is represented at CG, perhaps comprised of a group of related kin with the possible presence of a few other visitors.

131

At this point I step back and view the distribution of exotics across the entire Folsom landscape. Figure 36 depicts this distribution for the 16 material types compared in this study plus a few instances of other tool stones noted as exotics in published accounts. Lines connect primary source areas with sites where raw materials are found in percentages of six or less. The mean distance over which these points, preforms and channel flakes were transported is 444 km. The maximum is between 800 and 900 km and includes Chuska at Cedar Creek, Edwards at Lindenrneier, and Knife River flint in southwest Iowa . The latter specimen was recovered near the confluence of the Missouri and Platte Rivers (Billeck 1998:405). Morrow and Morrow (1999:79) report that other Folsom points from Iowa are made of local materials such as Burlington chert and appear to result from a residential occupation of the area rather than "stray points making their way into the region." The additional presence of Niobrara jasper suggests that land use and/or social connections extended from this area toward the Kansas-Nebraska border. To the northwest, Stanford (1991) identified a large proportion of Knife River flint (44%) in the weaponry-related assemblage from the Agate Basin site in eastern Wyoming (Figure 33). This may suggest residential land use extending from the primary source area in North Dakota southwest beyond the Black Hills. Subsequent research in the Hell Gap/Hartville Uplift area by Frison, Sellet and others reportedly identified closer sources of similar-appearing material (MacDonald 1998a:85). As a result of these new findings, estimates of the frequency of Knife River flint at Agate Basin may need to be revised downward. It may occur as an exotic rather than in frequencies >20%. In either case, the direction of movement along a northeast to southwest axis remains the same. Further evidence for this direction of travel consists of exotics unearthed from the lithic workshops at the Bobtail Wolf site in the Knife River primary source area (Root 1993; Root and Emerson 1994). Exotics made of Rainy Buttes Silicified Wood, occurring primarily as unpatterned flake tools, were procured from a small source area located about 100 km to the south (Root and Emerson 1994:185). Likewise exotics of Tongue River silicified sediment are thought to derive from this direction as well. "The presence of these stones suggests that at least some of the groups who occupied the site moved from the south, northward in to the KRF primary source area" (Root and Emerson 1994:185). Biface thinning flakes of phosphoria from the Wyoming-Montana border (nearly 300 km to the southwest) were recovered

132

Figure 36. Geographic distributio;n of raw matrials at percentage frequencies of six or less.

133

near a possible hearth where a flintknapper had been finishing the final thinning and shaping of a biface (MacDonald 1999:153). Recent analysis by MacDonald (1998a) concludes that these material types were contemporaneous in the lower Folsom component of Block 4. They, therefore, provide an indication of the social realm of the inhabitants of this portion of the site. He sums up these results as follows: the region surrounding roughly the Missouri, Little Missouri, Belle Fourche and Powder Rivers was extensively used during the Folsom period and may represent a wellestablished territory... If such a territory existed, it covers much of the same ground that the Hidatsa used in their hunting expeditions. . and is also similar to the territory ascribed to individuals who used the Vore site... during the Late Prehistoric Period. (MacDonald 1999:155) MacDonald and Hewlett (MacDonald 1999:148-153) constructed a three-tier model of Folsom mobility by extrapolating population densities and mating distances from tropical forest and semi-desert foragers to Folsom populations. The three ranges of mobility in this model are briefly summarized here from smallest to largest. Local foraging by individuals and families (i.e., to extend up to 100 km on average.

"micromovement’) is anticipated

"Mesomovement" on the order of 100 to 160 km is predicted to occur

less frequently and include travel to visit family and attend group functions. A mating distance of 80-100 km is thought to fall within this range of travel -- perhaps comparable to an annual range by a minimal band. Abundant tool stones, including exotics, are expected to derive from within this range as exemplified by his analysis at Bobtail Wolf. Based on the mean mating distance of the Aka (119 km) the final level of Folsom mobility was estimated between 160 and 500 km, with a median of 330 km (MacDonald 1999:151). This "macrornovement" consists of the least frequent scale of travel to ’explore exotic sites of potential resources... [it] is similar to the extended and lifetime ranges in Binford’s model [1983] and the lifetime and gift recycling range in Sampson’s model [1988]" (MacDonald 1999:149). Macromovement is the largest sphere of travel predicted. MacDonald (1998a:85) predicts that mobility and/or trade of far-flung non-local lithic tools was quite minimal during the Folsom period. Most mobility, as seen in the distribution of geologic sources, was within 100 km of Bobtail Wolf. . . suggesting fairly well-established territories during the Folsom period. . .the curation of small quantities of stone, in the form of bifaces or perhaps, Folsom points, suggest an extra somatic purpose---religious, social or reproductive---for curation of the stone (e.g., Gould and Saggers 1985). MacDonald and Hewlett (1999) model movement into three distance ranges: 0-100 kin, 100-160 km, and 160-500 km. These may well reflect recurrent stretches of landscape across which Folsom social

134

interaction occurred. However, the postulated time intervals -- daily range, annual range, and extended (lifetime) range -- attributed to these inferred travel distances appear to be too abbreviated to accommodate existing information regarding the distribution of exotics and non-exotics at several Folsom sites. In some assemblages, materials from distances proposed to be indicative of ’macro or ’lifetime" movements may represent, alternately, overlapping annual or seasonal ranges of different individuals and/or families in a single aggregated group. Table 13 compares four forager mobility models with raw material data from what are arguably single component assemblages from Bobtail Wolf, Cooper, Lipscomb, Waugh, and CG.

Table 13. Four Forager Mobility Models Compared with Raw Material Transport Distances Reflected at Selected Folsom Sites. Reference Group Level 4 Level 1 Level 2 Level 3 Binford 1983 Sampson 1988

Nunamiut Kung

Foraging Radius

Annual range

Extended Range

Lifetime Range

Core range

Annual range

Lifetime Range

Gift Recycling Range

Wobst 1974

Paleolithic Model

Nuclear family

Minimum Band

Maximum Band

MacDonald and Hewlett 1999

Folsom Forager Model

Micromovement 0-100 km

Mesomovement 100-160 km

Macromovement 160-500 km

Folsom Locality

Folsom Assemblage

Transport Distance km 0- 100 100 - 160

Transport Distance km 160-500 >500

BobTail WolfBlock 4 Lower Folsom Level

Single Component

Knife River Rainy Buttes Porcellanite Swan River

Phosphoria

Cooper Upper Kill

Single Component

Alibates

Edwards

Cooper Middle Kill

Single Component

Alibates

Edwards

Cooper Lower Kill

Single Component

Alibates

Edwards Niobrara

Lipscomb

Single Component

Alibates

Edwards Niobrara

Waugh

Single Component

Alibates

Edwards

Cattle Guard

Single Component

Black Forest Hornfels

-

-

-

Porcellanite

Trout Creek Cumbres

Black Forest Chuska, Alibates Sources: Bement, 1999: 141-142; Hofman 1991:335-355; MacDonald 1999:341, Table 3;

Edwards

This study.

135

Macrornovements between 160 to 500 km are predicted by MacDonald and Hewlett’s model to be representative of exploratory mobility occurring infrequently, perhaps by males traveling in the outer reaches of their lifetime ranges. However, at each of five bison kills with compelling evidence of single occupations, tools are made of raw material types procured at just these distances. The data from Cooper suggests that three bison kill events occurred at intervals of three to five years (Bement 1999:172). While different mobility patterns may well characterize life on the Northern versus Southern Plains, it is doubtful that this disparity would approach the difference between an annual or semi-annual range in the south and a lifetime range in the north. The cool, high-altitude climate of the San Luis Valley is in many ways more comparable to the Northern than the Southern Plains, yet raw material transport distances are comparable between CG and Southern Plains kills. Differences in site function between lithic workshops on the one hand and early fall bison kills on the other may be contributing to very different views of the scale of meso- and macro-movement as inferred from raw material types. While not discounting the possibility of ideological reasons for the use and transport of exotic raw materials, mobility patterns tied to subsistence concerns appear to be contributing strongly to the widespread occurrence of exotic materials at some Folsom sites. Davis et al. (1985:45) report "a large bifacial core of Hartville uplift chert and a single flake of Alibates’ from the Indian Creek site in Montana, and the additional possibility of Knife River flint there as well (Davis and Greiser 1992:265). If the flake is truly Alibates, it exemplifies transport in excess of 1,300 km and implies contact in one form or another across nearly the entire Folsom landscape. Based as it is on a single flake, it is noted here but it is not depicted in Figure 36. In combination the distributions of exotics and non-exotics reveal extensive, overlapping areas that link smaller patterns of interconnected land use and social contact between regions. Movement up and down the eastern flanks of the Rocky Mountain Front Range is indicated -- perhaps an inter-regional Pleistocene counterpart to Interstate 25 today. A diversified array of exotics was discarded at Lindenmeier, situated near the northern terminus/beginning of the Front Range. Does this represent expected variation within a very large sample size and/or was this a favored stopping place along a well-used inter-regional trail network? Years ago Cassell (1941:451) proposed a corridor of Folsom movement along the eastern

136

mountain front where the favorable ecotone between mountains and Plains was endowed with springs, good grass and shelter, a draw for bison and hunting groups alike. As Benedict (1992a:3) points out, in "few regions of North America are vegetation zones as compressed as in the Colorado Front Range. Distances of as little as 25 to 30 1cm separate alpine tundra from shortgrass plains, making it possible for aboriginal people to visit a broad range of ecosystems with little effort, and to sample, at different seasons, a diversity of resources." High-quality tool stone and mitigated winter temperatures enhance the region’s appeal. It is understandable that the largest human populations in Colorado currently reside along the Front Range. In speaking with m Arapaho Elder Mark Soldierwolf in Fort Collins a couple of years ago about historic use of the Front Range trail, I asked what the trail was called among the Arapaho. He replied (from the perspective of northern Wyoming) that it was it known as "the road south." One presumes that to people from the Southern Plains and Rockies it was "the road north. Face-to-face interaction and down-the-line exchange of information apparently linked people together on both intra- and inter-regional scales during the terminal Pleistocene. The following section discusses the socioeconomic importance of gift exchange among hunter-gatherers and its possible contribution to the transport of lithic materials across regional landscapes. Exchange in raw materials is not postulated here as an important source of stone, per se, but as an important means of bolstering social relations among people living in neighboring areas.

Gift Exchange and the Long-distance Transport of Materials Information and material objects move fluidly among modern hunter-gatherers sharing kinship and other social ties. A common mechanism for the movement of durable items of the sort recovered archeologically is informal gift exchange. Gift exchange is part and parcel of "the elementary good manners of sharing fully with members of [a] local camp" by people on extended visits (Lee 1979:457). Gift exchange, food sharing and mutual access to resources are means of pooling subsistence risk (Gould 1980:156-158; Hayden 1982:117-118; Jochim 1981:137-138, 1998:196). It is by such socioeconomic avenues that the wellbeing of one’s relations is secured through the "storage of social obligations" (Wiessner 1977:13) rather than through the storage of material goods (see also Wenzel

1995:55-57).

137

Pertinent here is a system of long distance gift exchange, known as

hxaro, which connects !Kung

San in social networks of mutual reciprocity (Wiessner 1977): At the heart of !Kung San reciprocity and social networks is an exchange relationship called hxaro. This relationship involves balanced but nonequivalent delayed exchange of gifts, whose continuous flow gives both partners information about the status of an underlying relationship - a bond of friendship accompanied by mutual reciprocity and access to resources. On the average, !Kung have sixteen hxaro relationships, each of which links them to broader hxaro networks ... the person who has gives to the one who is in need, need being relative to the means of both. Hxaro relationships are geared to unpredictability, and returns are measured by their utility to the receiver rather than by a fixed quantity. The giver of assistance has no desire to be paid back immediately, therefore the relationship can not be considered even and is not open to cancellation. Rather, the aim is to store the debt until the situation of have and have not is reversed, allowing hxaro relationships effectively to cover unpredictable losses. However to cover unpredictable losses, the terms of hxaro relationships are so loose that the status of the relationship, particularly for partners living far apart, can become ambiguous. The continual balanced flow of gifts lets each partner know that the relationship is still intact. (Wiessner 1986:105) Hxaro goods travel for hundreds of kilometers across the Kalahari from one person to another in down-the-line fashion (Lee 1979:365). Based on her ethnographic field study of the social dynamics of this gift exchange system, Wiessner concluded that some form of gift giving is to be expected among nearly all foraging groups: As long as hunter-gatherers maintain widespread ties with kin and non-kin, the distribution of goods within the population should very clearly resemble that attained through hxaro because, whether or not gift exchange creates relationship, it almost universally accompanies those formed in other ways. (Wiessner 1977: 363). Of particular interest to this study is the fact that arrows are among the objects widely exchanged among Kung, G/Wi, !Xo and Nharo hunters (Wiessner 1983:255, 260). !Kung individuals are known to exchange these items with people living 60-100 km away in areas with complementary resources: Of 236 !Kung arrows recorded, 57% were made by the hunter himself, 26% were received from exchange partners living 1-20 km away (including camp members), 3% were received from partners living 20-60 km away, 13% from partners living 60-100 km away, and 1% from partners living 100-200 km away. (Wiessner 1983:261) Among the San, an arrow-maker either receives a large portion of the meat procured with his arrow or directs the distribution of this meat, conferring upon the act of arrow exchange a direct role in meat sharing. "Arrows are among the most widely traded technologies in many societies" (Greaves 1997:22 and references therein). A reference to arrow exchange among bison hunters in North America is recorded ethnographically among the Kiowa of the Southern Plains.

138

They usually made enough arrows for themselves and a little more for gift exchange and betting on horses. No high degree of specialization. Most men made their own arrows although quite a few were inferior arrow makers. In gift exchanges between tribes the other tribes were always anxious to secure Kiowa arrows. (La Barre 193 5:Mishkin, VIII19:13) According to ethnologist Steve Beckerman (personal communication 1999), hunters among the Bari of southern Colombia (tropical forest horticulturalists who depend on hunting and fishing as their principal source of protein, tBeckerman 1983) also exchange arrows. This is generally done with trading partners and other members of a social network outside of the immediate (i.e. co-resident) family. A Ban man explained that when he killed an animal with a gifted arrow he thought of the friend who gave him that arrow. The dynamic interplay between reciprocity, social ties and subsistence is highlighted in this example where, by enabling a hunter to kill game, the gift keeps on giving and continues to reinforce the bonds of friendship. As noted above, gifts such as these are often exchanged during extended visits. Extended visits within and between regions are common and, among other things, serve important roles in balancing fluctuations in subsistence resources and in providing opportunities to secure mates. In the study area, the regional ecosystems of southern Colorado and northern New Mexico are characterized by the juxtaposition of montane, plains, and basin and range plant and animal communities. Broadly speaking, seasonal fluctuations in their availability include a progressive ripening of some plants from southern to northern latitudes, from south-facing to north-facing slopes, and from lower to higher elevations. Likewise, elk, bighorn sheep and bison move to and from higher elevations during summer and fall. Superimposed on seasonal rhythms are vagaries of rainfall, drought and the like which effect the yearto-year and place-to-place availability of staple species.

Historic and Recent Land Use in the Study Region by Native People Variation in resource richness and availability provided strong impetus for movements between harvesting areas by eighteenth and nineteenth century Moache and Capote bands of the Southern Ute Tribe whose traditional land use areas included the San Luis Valley (Aldan Naranjo, personal communication 1998). Spiritual Leader and Southern Ute Tribal Elder Everett Burch (1998) explained that since buffalo were abundant in many areas of southern Colorado and northern New Mexico, Ute People moved primarily to obtain other resources that they needed, meanwhile hunting bison in those areas. The different

139

availability of ripening plants, migratory birds, wood, stone, and fiber for tools and cordage (baskets?), as well as opportunities to socialize with other people, are expected to have influenced Folsom mobility patterns as well. Mobility patterns and travel routes are further shaped by features of local geography. The manner in which landscape use was influenced historically by such perennial features as mountain topography may hold keys to understanding population movements by other people in the same area. The Capota and Moache bands ranged throughout the study region historically and today retain a portion of their homeland in the country south of Durango, Colorado. Traditionally, some members of these bands ranged in the vicinity of the Colorado Front Range, where they wintered in protected foothill areas and often hunted during the summers in South Park and adjacent environs. Prominent mountain peaks marked homeland boundaries. Among these landmarks were twin mountain peaks (now known as the Spanish Peaks) northwest of Trinidad, Colorado, that rise upwards of 3800 in and 4100 m, respectively. Aldan Naranjo (1998), a knowledgeable Southern Ute tribal historian, notes that the Capota traveled east onto the Plains as far as they could go and still see these mountains, and then one days ride beyond. This delineated their hunting area on the Plains. Other mountain peaks identified boundaries to the north and west. People of the Capota band also traveled to the San Luis Valley, south to Taos, New Mexico, and eastward to the Texas Panhandle (Aldan Naranjo, personal communication 1998). Members of the Moache band also lived in south central Colorado and northern New Mexico and frequented the San Juan Basin and the country near Abiquiu, New Mexico (Figure 22). Frequent visiting and kinship relations existed between these two bands and with other Ute bands to the north and west. Capota and Moache overlapped in the San Luis Valley. The land traversed by these two bands and a neighboring band to the west (Weeminuche) is summarized by (Opler 1940:123) as follows: Three centuries ago the [Southern] Ute ranged over the mountains of southwestern Colorado bordering the Continental Divide, and over the more level land of states adjoining their rugged base: over southeastern Utah, northeastern Arizona, northern New Mexico, a corner of Oklahoma, and even as far east, on rare occasions, as the Panhandle of Texas. This wide expanse of territory is a land of sharp contrasts and dramatic changes of scenery. In the northern part of the Ute range lies the heart of the Southern Rockies, a tangled crisscross of mountain chains, narrow passes, rock-pockets, and swift streams. Farther south, near the border of New Mexico, are rolling hills that level out in

140

semi-arid mesa and desert country on the west, and flatten down into.., plains and prairies on the southeast The Kapota, occup[ied] the region around the Sangre de Cristo and San Juan ranges in warmer seasons. Still farther to the east lived the Mowatsi, who occupied the territory between the Sangre de Cristo and Culebra ranges on the west over as far east as the present sites of Denver, Colorado Springs, Pueblo, Raton, and Trinidad. (Opler 1940:127) Young (1997:16) shows the maximum hunting range of the Capote and Mouache (eastern) bands as extending to western Kansas, Oklahoma, and a corner of the Texas Panhandle, and south to Santa Fe, New Mexico. Although as Naranjo points out, they normally ranged within a day’s ride of being able to view the tallest peaks along the Front Range. Moache and Capota lands collectively encompass most of the geographic areas represented by lithic materials at Cattle Guard, although the former groups had horses to facilitate their mobility. Other native groups traveled to the San Luis ’Valley for economic and spiritual purposes. Pueblo peoples from northern New Mexico traditionally came here to gather turquoise and feathers and to hunt bison, although the vast meadows of the San Luis Valley were recognized and defended as Ute country. The earliest known written record of bison in the San Luis Valley is in the journal of Don Diego de Vargas for July 1694 (Vargas 1694). He tells of Spanish efforts to secure fresh meat from a herd of 500 animals in the southern valley. The Ute who arrived to protect their rights to these hunting grounds soon challenged them (Colville 1996:219-223). It is of note here that after the Spanish spooked this herd on the first day of hunting, they were largely unsuccessful in killing many more bison on subsequent days. The San Luis Valley is part of the cultural landscape of the Navaho People as well. Blanca Peak, a majestic mountain which rises above 4200 in in the Sangre de Cristo Range (adjacent to the CG, Linger and Zapata Folsom sites), is Sisnaajini, the sacred mountain to the east (Kelley and Francis 1994:52, 82, 174). It is one of four sacred mountains that enclose the traditional homeland in the four cardinal directions. "Traditional Navajos travel to the mountain where soil, plants, and spring water are collected and used in important ceremonies. The protection of gathering rights on the mountain is of major concern to Navajo people" (Spero 1999b:260), as it is to healers among other groups to this day. In sum, the upper Rio Grande basin was situated historically near a point of overlap in the land use areas of two Southern Ute bands. In turn it further overlapped the hunting ranges of tribal groups living to

141

the south and east. During historic (and some prehistoric periods) it might accurately be viewed as a place of cultural contact. Hints that this might also have been the case during early Paleoindian times lie in the local recovery of a crescent and a Haskett point associated with groups in the Great Basin and Dalton points associated with eastern Paleoindian groups (Jodry 1999). In the next section I describe the exchange of Alibates knives and other bison processing tools in northern New Mexico among Puebloan people and neighboring bison hunters. Pertinent to this study is the archeological record of the inter-regional exchange of Alibates and obsidian raw materials that resulted from these social alliances. Folsom people hunted bison with Alibates and Edwards tools thousands of years earlier in this same area (Figure 24, #49).

Protohistoric Alliance and Exchange in the Study Region Between approximately AD 1250 and 1700, nomadic bison hunters who ranged over northern New Mexico, southern Colorado, and western Texas and Oklahoma maintained socioeconomic alliances with sedentary villagers who farmed in northern New Mexico (Spielmann 1983, 1991). Mutualistic exchange of subsistence resources between these culturally distinct populations included bison meat and robes (Creel 1991:40-42), beveled knives of Al/bates, hide-working tools, and mussel shells from the Plains for corn, ceramics, Jemez (Lintz 1991:95) and other obsidian, and turquoise from the Southwest (Spielmann 1983:259-267). Spielmann (1983:268) suggested that this trade developed during an initial 200-year period of encounters and gift exchanges that took place when Pueblo and Plains people met while both groups were out bison hunting. This gift exchange is thought to have resulted in the steady, low-level deposition of nonlocal, durable items at archeological sites. Noteworthy here is the fact that Alibates chert from the Texas Panhandle and obsidian from New Mexico were among the archeologically visible items that changed hands. The exchange of utilitarian items (including bison processing tools made of both bone and stone) increased during the subsequent 250 years. In this latter period at least 58 beveled, Alibates knives found their way to Pecos Pueblo (Kidder 1932:31, 39, as reported in Spielmann 1983:Table 1). I will discuss these knives further in Chapter 5 when I describe similar knives (ultrathins) made by Folsom people. Of immediate interest here is Spielmann’s (1983:269) conclusion that the Plains-Pueblo trade had its origins in

142

gift exchange that cemented social alliances between Pueblo and Plains groups that gave the former access to the latter’s hunting territories when bison migrated out of Pueblo country. As noted by Lintz (1991:98), "the most remarkable characteristic of the list of ... [non-perishable] trade goods.. .is its redundancy with material classes already present." That is, stone (obsidian) was exchanged with people who already had sufficient quantities of stone (Alibates), likewise with pottery and pipes. The value of the items exchanged apparently lies beyond the objects themselves -- to the social connections that were strengthened by the interaction itself (Jochim 1981:188). Lintz (1991:106) notes that Alibates was probably exchanged as flake blanks and perhaps large bifacial cores and such finished artifacts as beveled knives, snub-nosed end scrapers, drills, and projectile points." It is postulated here that periodic fluctuations in the availability of bison and other key resources (i.e., mates) may have provided a strong motivation for Folsom people to maintain social connections with groups in other areas as a form of social insurance against unexpected shortfalls in their own country. Notwithstanding the net trend toward greater animal biomass during the Younger Dryas interval discussed previously (Chapter 2), periodic regional variation that effected resource distributions are to be expected during the lengthy Folsom time interval. Places rich in game and plant foods one year may have provided relatively poor hunting and collecting the next. An example of just such an abrupt, unpredictable resource fluctuation occurred in the study area coincident with a nineteenth century peak in the abundance of bison and other large game animals (Benedict 1999:8-9). A particularly severe storm in the spring of 1844 deposited up to 3.4 m of snow (as measured on level ground) near Colorado Springs. The snow melted slightly, then froze solid. The resultant inability to access forage decimated game animal populations from "the Laramie Plains south to at least the Arkansas River, and from the Front Range eastward 100 km into the Piedmont and Plains" (Benedict 1999:8-9). The storm destroyed all of the bison, and most of the elk, deer, and antelope in the Pueblo area. . . By the summer of 1845, when Colonel Stephen W. Kearny’s dragoons marched from Fort Laramie (on the North Platte River in what is now southeast Wyoming) to Bent’s Fort (on the Arkansas River in southeast Colorado), the landscape that had fed Sage so abundantly [in the summer of 1843] had become an empty larder. (Benedict 1999:9)

143

The extensive size of the region effected, roughly 35,000 km’, gives some scale of the geographic areas that might be effected by short-term ecological disasters, and the corresponding distances over which people might need to travel to escape the resulting subsistence crisis. In short, ecological variability necessitates strategies for dealing with lean times. Regardless of whether subsistence resources, particularly bison, might have been readily available in most years, in many areas, people necessarily adapt to the range, not the mean, of anticipated conditions. The risk of unexpected shortfalls, such as wiped out otherwise abundant game in the 1840s along the Front Range, arguably favored the investment by Folsom groups in social mechanisms that enhanced inter-regional cooperation among groups and that facilitated fluid movement between areas as circumstances dictated: . . in regions of generalized unpredictability, but where specific shortages tend to affect only one or a few groups at any one time, i.e., spatial variation in the timing of risks, one would expect that the groups would be characterized by a variety of mechanisms that facilitate the redistribution of people. Briefly, such mechanisms should include spatially extensive (rather than intensive) social relationships. . (Jochim 1981:193) Generosity and sharing are key social mechanisms for pooling subsistence risks (Cashdan 1985; Hayden 1981; Savishinsky 1994:99). Sharing beyond the level of the household and local group often involves mutual access to resources as well as exchanges of food, luxury goods, and utilitarian items between visitors and hosts. I expect that these socioeconomic strategies were among the suite of options variably exercised by Folsom groups and that both direct and down-the-line exchange of tool blanks, preforms and points contributed to the wide distribution of exotic lithic materials recovered archaeologically. Other factors potentially effecting regional tool stone transport are symbolic connections between individuals and specific source areas and/or certain stone attributes such as color and pattern. For instance, Western Australian Aborigines prefer particular stones from the region where they are born and which are associated with their totemic descent. Thus, a man may have a sense of kinship with some of these localities, and he will value the stone material from them as a part of his own being. Stone materials thus acquired are not sacred in any strict sense but are nonetheless valued highly enough to be transported over long distances by the owners. (Gould et al. 1971:161-62)

144

In another example, Table Mountain jasper in Middle Park, Colorado reportedly ’contained floccules and filaments of red, brown, and yellow chert that reminded the Arapahos of the lungs of freshly butchered animals" (Benedict 1992a:13, from Toll 1962; see also Loring 1997:209-210). Stanford (1999:291) recently suggested that Folsom people preferred purple to white varieties of Flattop chalcedony, although both are thought to have similar flaking qualities. At a quarry on a high butte in northern Colorado, purple chalcedony occurs in seams underlying those of the white material. Despite the fact that the former was less accessible, Folsom people seem to have been motivated to obtain it as evidenced by its greater representation in Folsom collections from the nearby Lindenmeier, Powars and Hahn sites. Later Paleoindian peoples in the same region made use of both varieties. Ellis (1989:

156-16’ )

has argued that Great Lakes Paleoindians selectively used distinctive raw materials from localized sources as emblematic signals that communicated group identity. Folsom knappers were attentive to the aesthetic qualities of lithic materials and spiritual aspects of hunting may have influenced material choice for projectile point manufacture. Frison and Bradley (1999:65) recently expressed similar views with regard to Clovis people. In describing ritual relations

between Mistassini Cree hunters and their prey, Tanner (1979:141) noted that ". . . the decoration of utilitarian objects used in the hunt is for the purpose of showing respect to the animal about to be killed, and to ensure the object performs its function properly." Nelson (1978:286) observed that the "aesthetic impulse lies somewhere outside, or between, practical knowledge and religious concepts" among the Koyukon Athabascans of interior Alaska. Some combination of functional, aesthetic, and spiritual concerns may have influenced raw material selection for hunting and other gear during Folsom times.

Concluding Remarks Extensive interaction networks apparently stretched out in overlapping segments that linked Folsom groups across their entire range. Two principle lines of evidence support this.

First is the

remarkable coherence in Folsom fluted point design and manufacture techniques that persisted for hundreds of years from Canada to Mexico and Iowa to Utah. This alone argues against small, isolated populations. Wiessner’s (1977, 1983) studies of shared reciprocity and projectile point style in the Kalahari suggest that

145

"groups who emphasize risk-pooling tend to de-emphasize group boundaries via stylistic means", as previously noted by Ellis (1989:159) in regard to Paleoindian groups in the northeastern United States. Second, raw material distributions that link Folsom sites with distant source areas crisscross the Folsom landscape at an inter-regional scale. Exotic raw materials from sources 400 to 900 km away were recovered at 46% of the sites in this study (26 of 56). Similarly, raw materials occurring in comparatively high frequencies (>20%) appear to originate from areas approximately 200 km away at 41% of these locales (23 of 56). This suggests high residential and logistical mobility and the likelihood of contact with people from other areas both by design and inadvertently. The strong positive correlation for hunter-gatherers between relative contribution of hunting to the diet and the size of land-use areas (Kelly 1983:297, Figure

5) predicts the existence of large hunting areas

during Folsom times. Among the largest ranges cited are those ethnographically recorded for the Crow who moved over 61,880 km 2 per year (hunting contributes nearly 80% to their diet) and the Nunamiut who covered over 63,700 km 2 per year (hunting provides 87% of their diet) (Kelly 1983:280, 297). Bison and caribou are the principal herd animals hunted by these groups, respectively. Although "inferring annual range size using data on distance between sites and lithic sources is a difficult undertaking" (Ellis 1989:160), recent estimates suggest that members of Folsom bands ranged over areas in excess of 100,000 km2 (Amick 1996:415; MacDonald 1999:156). Some models of Folsom mobility assume that population density was so low as to preclude the need to arrange for access to any area in which a particular group wished to travel (e.g., Kelly and Todd 1988:239). This study assumes that Folsom groups were not colonizing populations and did not live either below the threshold of, or in the absence of, systems of land tenure and foraging rights. Rather it is thought that Folsom people were surrounded by neighboring cultural groups (both Folsom and non-Folsom) who were loosely associated with socially recognized, albeit extensive and partially over-lapping, use-areas. It is further suggested that some form of land tenure was in place whereby people did not simply move as they pleased without any regard to the hunting and gathering areas of neighboring groups (e.g. Wiessner 1977:12; Yengoyan 1976:125). If so, then "for the movement of people to occur, not only must social channels exist, but they must be reaffirmed and reinforced" (Jochim 1991:193). As Kelly (1995:192) put it:

146

Resources may not be there for the taking, but they are apparently there for the asking. The giving of permission is the giving of a gift-- and it puts the receiver in debt... social relationships are the primary social-security system of hunter-gatherers: the more social relationships, the greater the likelihood that a group or individual will be able to fall back on someone in times of need" (Kelly 1995:192). This chapter has shown that gift exchange of lithic implements and/or tool stock is among the social strategies used by some small-scale societies to maintain reciprocal ties among people living in neighboring and distant areas. Social relations cemented through gift exchange form a safeguard in times when critical resources fail. Cashadan (1985:471) maintains that reciprocity networks that serve as insurance against hard times are most likely to prevail in situations where the economic fortunes of the people involved are independent of one another. Basically this postulates that not all the people in a kinship or other network of mutual support will be in "the same boat’ at the same time. This implies that the Folsom groups most likely to be involved in gift-exchange, risk-pooling, and shared access to foraging areas will be those occupying lands with complementary resource structures. This may include regions with perennial differences in topography, plant communities, lithic availability, and/or seasonality. Additionally it might involve areas with temporary disparities in such factors as forage condition, game availability, or gender imbalances that impact mate selection. In sum this chapter argues that: 1) the need for access to large territories and

2) the risks inherent

in depending heavily on mobile herd animals as an economic mainstay selected for social mechanisms that fostered cooperation and mutual assistance among Folsom groups to insure their long-term survival. In the face of economic uncertainty, later prehistoric groups further hedged against shortfalls through food storage and/or animal domestication. Terminal Pleistocene hunter-gatherers may have increased their options through cementing kinship and other ties with neighboring groups. At a local scale, a comparison of the presence/absence of raw materials at CG and their representation across flaking debris and tool types suggests that people arrived at the site from at least two directions: north/northeast and south/southwest. Additionally, some of these individuals had hunting gear made of stone from the Southern Plains and Edwards Plateau. Similar patterns of raw material representation at the nearby sites of Linger, Zapata and Reddin further identify social interaction and/or residential mobility connecting the San Luis Valley, the upper Arkansas Basin and Front Range, and the

147

San Juan Basin and Taos Plateau. These may represent areas regularly traveled by members of one or more bands. Additional extensions of at least hunting ranges appear to link this region with areas farther to the southeast. The following chapter looks closely at technological aspects of the stone tool assemblage at CG.

CHAPTER 5 STONE TOOL TECHNOLOGY AT STEWART’S CATTLE GUARD SITE Introduction The dual goals of this and the following chapter are to describe variation in the CG stone tool assemblage and to explore its meaning in behavioral terms. Chapter 5 summarizes technological attributes of stone tool design, manufacture, and use. These data are combined with information about tool distributions in Chapter 6 to investigate how the site was organized. My intent is to reconstruct, as far as possible, the kinds of activities conducted at Cattle Guard and to infer how their predictability, organization and spatial distribution effected the nature of the recovered stone tool assemblage. My approach to descriptive analysis largely follows that outlined by Ahler (1994) for the study of Folsom materials from Lake Ilo, North Dakota (Root 1993, 1994b; William 1994; William and Shifrin 1995). This system was developed for analysis of Plains Village collections from South Dakota (Ahler

1975) and later was expanded to accommodate a greater range of prehistoric materials from the northern Plains (Ahler 1994:1). Stone tool attributes are classified independently along five dimensions. These include: 1) production technology (how a tool was made), 2) tool function (how a tool was used or was intended to be used), 3) systemic context (a tool’s use and recycling history and its potential for further use at the time of discard), 4) raw material, and 5) style ("morphological variation imparted to the tool as a consequence of the social context in which the manufacturing process occurred" (Ahler 1994:27). The inter-relationships among these attributes are combined with faunal and other contextual data to form the basis for interpreting site activities and assemblage variability at CG. Attributes recorded during lithic analysis are presented in Appendix B. Artifact descriptions are organized by category (bifaces and unifaces) and tool class (projectile points, preforms, ultrathin bifaces, end scrapers). To facilitate comparative studies, projectile point measurements follow conventions established in the Lindenmeier report (Wilmsen and Roberts 148

149

1978:103-106, Figure 99). In like fashion preforms were analyzed with reference to the manufacturing sequence developed for the Hanson assemblage (Frison and Bradley 1980:52) and applied later to materials from the Agate Basin (Bradley 1982:186-189) and Bobtail Wolf (Root 1994a:154-155) sites. A coding system intended to address projectile point haft design through an examination of breakage patterns was developed for this study. The following sections present detailed description and discussion of the CG chipped stone assemblage, emphasizing tools used in bison procurement and processing.

Stone Tools and Flaking Debris Description of the Analytical Samples An important distinction is made in this study between the block sample and the total lithic sample (Figure 37). The block sample consists of tools and fragments (i.e., surface and excavated) recovered from a contiguous area measuring 796 m 2, bounded by coordinates North 26 to 80 and East 101.5 to 136. This sample includes nearly all flaking debris from a contiguous subset of this block (620 m

2) extending from

N26 to 74 and E120 to 136. The block sample is used for intrasite spatial analyses (Chapter 6) and for comparisons of the relative frequencies of tool classes. Table 14 summarizes the block sample by general artifact category.

Table 14. Block Sample by General Artifact Category a General Artifact Category Flaking debris

a

General Artifact Category

17,367

-

Pigment

54

Weaponry-related artifacts

219

34.3

Unifaces

366

57.3

Channel flake Projectile point Preform

128 64 27 26 14 1 1 10

Small edge fragment Flake tool, retouched Flake tool, use-modified only 4.1

3.4

Anvil Hammers tone/anvilb Other utilized cobble

175 96 95 22 1 18 3

Ground Sandstone Implements

5

0.8

Hammerstone/anvil flakes’

171

Sandstone abrader Knapper hammerstone/abrader

4 I

Other Bifaces

Ultrathin biface Ultrathin preform Small, thick, bifacial core Other biface/fragments

Utilized Cobbles

Total = 18,230 Note: Includes N26 to N80, E101 to 136 (excluding test units Ml 16-120 and Ni 18-Al).

--

aCalculated on 638 tools and fragments (excluding flaking debris, hammerstone/anvil flakes and pigment). blncludes fragments larger than 50 mm in maximum dimension. ’Includes flakes and fragments smaller than 50 mm in maximum dimension.

150

i

1

i

SCALE

100 104 108 112 116 120 124 128132 136 140144148152156

Figure 37, Location of block and total samples.

&

151

The total sample includes the block sample plus all other Folsom projectile points, preforms, other bifaces, end scrapers, gravers, abraders, pieces of red pigment and a spoke shave recovered from excavated (N26-100 /E10l-180) and surface (N26-114/E60-232) contexts (1981 to 1996 field seasons). Technological analyses of projectile points, preforms, channel flakes, and end scrapers draw on the larger sample sizes represented in the total sample. Additionally, the total sample was used in refit studies and spatial analyses of points, preforms, channel flakes, and end scrapers. The total sample consists of 18,643 lithic specimens (Table 15). All ultrathin bifaces recovered to date were found in the block sample portion of the site. Tabl e 15. Total Sample Included in the Study by General Artifact Catego ry . a General Artifact Category General Artifact Category Flaking debris 17,367 Pigment -

Weaponry-related artifacts

536

Channel flake Projectile point Preform

276 211 49 27 14 1 1 11

2.7

Ground Sandstone Implements

8

0.8

Sandstone abrader Knapper hammerstone/abrader

7 1

Other Bifaces

Ultrathin biface Ultrathin preform Small, thick, bifacial core Other biface/fragment

54

Unifaces Small edge fragment Flake tool, retouched Flake tool, use-modified only

a

74

-

394

39.7

2.7

Anvil Hammers tone/anvilb Other utilized cobble

179 117 98 27 1 23 3

Hammerstone/anvil flakes’

210

-

Utilized Cobbles

Total = 18,643 Note: Includes N26-N1 14 and E60 to E232

’Calculated on 992 tools and fragments (excluding flaking debris, hammerstone/anvil flakes and pigment). blncludes fragments larger than 50 mm in maximum dimension. ’Includes flakes and fragments smaller than 50 mm in maximum dimension. Both the total and block samples derive from what is interpreted here as a single component late summer/early fall bison kill and associated processing and residential camp.

The primary difference

between the two samples is that the block sample includes all tools (un(facial and otherwise) from the designated block while the total sample does not include all unfacial tools from the entire site.

Production Technology The advanced stages of lithic manufacture are best represented at Cattle Guard in keeping with the site’s geographic position at some distance from raw material sources. There are a few exceptions. A blocky piece of Hornfels (tested core?) was brought to the site where it was reduced in a discreet

152

flintknapping area. Twenty-three primary and secondary percussion flakes were among over two hundred pieces of Hornfels flaking debris. Refit analysis indicates that several large flake blanks were removed from this knapping area. Represented among the debris were late-stage biface shaping and thinning flakes, possibly representing the manufacture of a Homfels ultrathin biface (previously discussed in Chapter 4). Two cores were recovered from the site. Both appear to represent opportunistic flake production from thick pieces of Black Forest silicified wood. The first specimen (0104-A1-2) is a small, unpatterned biface with a width/thickness ratio of 2.2 (Figure

38, a). It is fractured on both ends of its long axis and

along one lateral margin. Two of these fractures are coincident with impurities in the raw material. This bifacial core concluded its use-life as a source of very small flakes (50 mm mi in length) that typically exhibit retouch along the working edge. A few are modified by use only.

The salient attribute defining this group is a long, straight to convex working edge on one or both lateral margins (Figures 63 and 64). These edges are typically 55 to 70 mm long for flake knives. Maximum length varies between 34.48 and 121.5 mm (mean is 84.3 mm), maximum width between 21.9 and 86.2 mm (mean is 46.6 mm), and maximum thickness between 5.85 and 15.9 mm (mean is 10.5 mm). Working edge angles and inferences derived from use-wear analysis differentiate flake knives from side scrapers. Edge angles of the former vary between approximately 25 and 65 degrees (mean is nearly 45 degrees), and the latter range between 50 and 80 degrees (mean is nearly 60 degrees). Frison (1979: 261) demonstrated that bison can be very handily skinned and dismembered using flake knives quickly fashioned by percussion (see also Huckell 1979:184). These tools predominate in Folsom (Bement 1999; Figures 27a, 40g) and later period bison kill and processing areas (Frison 1991:128). Opening the hide requires a sharp cutting tool and the keen edges of an umnodified flake are well suited to this task. A combination of low and high power use-wear data indicates that some flake knives were made elsewhere and transported to CG where they were used (and resharpened) as butchering tools (Figure 63, b; Figure 64 a, c and d). Whether others were produced expediently on-site is uncertain (Table 56, Appendix A). One butchering tool is made on a large, flat flake of Chuska chert (Figure 64, a). The convex cutting edge exhibits narrow resharpening flakes evidently removed by pressure flaking. The cross-section of the retouched working edge appears faintly rounded with isolated areas of step-faceting. Abler (1997) infers transport and handling wear from moderate to light smoothing on arrises that does not extend over the retouched working edge. At higher magnification Kay (1996a) documents striae, microplating, and recrystallization judged consistent with extensive use in carcass dismemberment where hard material wear traces resulted from cutting meat, tendons and probably bone (Figure 65). Frison (1979:260) notes that the cooperative effort of someone ’holding the animal in proper position throughout the butchering, although not mandatory, can speed the process immensely." All flake knives and side scrapers were recovered in the residential camp and southwestern work area where activities appear to have concentrated following initial butchery (Figure 66).

221

Figure 63. Flake knives: a) S134-A2-6 and b) N108-A1-32.

222

Figure 64. Flake knives: a) G112-A2-12; c) A112-A2-6; d) A114-A2-6 and Side scraper b) Z116-A1-1.

223

E



Figure 65. Oriented photomicrograph of hard material cutting wear traces on flake knife (G112-A2-12). Adapted from Marvin Kay 1998, Figure 18.

224

82-

NI

100

104

108 112

116

120

124 128 132

Figure 66. Distribution of flake knives and side scrapers in the block sample.

136

225

End Scrapers There are 69 artifacts within this general class in the

total sample including 46 nearly complete

tools and one medial, one lateral, two proximal, and 19 distal fragments. The block sample includes 42 end scrapers and eleven fragments. Figure 67 compares the percentage frequency of raw materials used in end scraper manufacture (total sample). The implications of variation in end scraper length and width by raw material type (Figure 25) are discussed in Chapter 4. Descriptive statistics are provided in Table 32.

Table32. End Scraper Descriptive Statistics. Variable Number Range Length 54 19.2-68.7 Width 50 20.8-43.2 Thickness 59 4.3-11.2 Distal Thickness 60 1.8 - 10.2 Weight in grams 56 2.8-27.0 Note: all measurements in mm.

Man 33.2 29.9 7.3 5.7 8.1

.

tandard Deviation 9.1 5.2 1.7 1.6 4.9

SampleVariance 83.8 26.8 2.9 2.4 24.4

End scrapers are identified by a transverse working edge that is nearly always opposite the bulb of percussion and by a longitudinally elongated flake axis that gets shorter with resharpening. Similar to those from the Agate Basin site, CG end scrapers exhibit a ’wide range of morphological characteristics" (Frison 1982:45). Much of this variation is attributed to the effects of attrition and differences in flake blank selection (Figure 68). Most CG endscrapers (48%, n=33) are made on percussion flakes from

226

Figure 68. Variation in discarded end scrapers. Specimen numbers: a) C 12-A2- I, b) Eli O-A2-4, c) E108-A2-10, d) Dl 12A2-1981-1, e) Dl 10-A2-10, Al 12-A2-12, g) Al 18-A2-1, h) Di 18-A2-1, i) Di 12-A2-15, j) C 14-A2-6, k) Al 12-A2-5, 1) Di iO-A2-17.

227

non-bifacial cores. Thirteen appear to be biface thinning flakes (nearly 19%), and one is made on a polyhedral blade. Blank type is uncertain for 22 specimens (32%). Figure 69 shows a curated and possibly heat-treated, early stage end scraper (HI 18-A2-1). It is made of mottled tan Cumbres chert, which is known to turn red or pink when heated. The dorsal face on the proximal end exhibits a reddened and heat altered veneer (hatched in Figure 69). Invasive flaking removed portions of this veneer and exposed progressively lighter pink hues beneath. This retouch seems to have been done at CG as it overlaps transport wear indicated by dorsal ridge rounding and smoothing, and irregular marginal flaking. The bulb was removed by flaking. Reddening appears restricted to the proximal end and the chert in this area is slightly more reflective and may be modified in texture. Whether the heating was intentional or inadvertent is uncertain but the presence of heat alteration seems clear. The transverse edge of this tool is not shaped by retouch but its use to scrape soft material is suggested by irregular step flaking and edge rounding. The slightly convex working edge has an angle of nearly 65.

Maximum tool dimensions (mm) are 68.67 (length); 42.67 (width); 8.94 (thickness).

Gallagher’s (1977:410) observations among Ethiopian tanners indicate that the ideal dimensions for making an obsidian end scraper are 60 to 70 rum long, 40 to 50 mm wide, and about 20 mm thick. These tools are fashioned in five minutes or less and are used exclusively to scrape cowhides. By way of comparison, a North American tanner recommends an end scraper blade (made of metal) that is 80 to 160 mm long, 25 to 50 mm wide, and 7 to 8 mm thick. He notes further that a convex working edge is important because it allows you to apply pressure while scraping without corners cutting into the hide’ (Riggs 1980:37). As the working edge loses its convexity through resharpening, its suitability for skin dressing decreases. The early stage Cumbres end scraper conforms well to the dimensions sought by modem tanners and may imply that a woman geared up with tool stock prior to coming to CG. As discussed in Chapter 4, Cumbres end scrapers exhibit less attrition (on average) than end scrapers of other materials. The possibility that some scraper blanks may have been deliberately heat-treated is intriguing but conjectural. A second side scraper with a pronounced greasy luster and the pink color characteristic of many heat crazed and pot lidded Cumbres artifacts at CG may also be heat altered. Additional evidence is needed to assess whether or not Cumbres material was deliberately heat treated during Folsom times.

port wear in high irrises

Red Flake surface from heat alteration. Unflaked since heating. I

Pink due to heat alteration, flaked at CG after heating. Percussion flaked since transport and heating.

229

Another early stage end scraper is made on a large outre passe flake of Black Forest silicified wood (P102-A1-3; Figure 70,a). The distal margin retains the flat, weathered exterior of the parent material and has an edge angle of 80 to 85. This transverse working edge is modified by use only and an adjoining lateral margin is retouched. Kay (1998) inferred (from use-wear evidence) that this tool was a hafted end scraper used on soft materials. It is 91.72 mm long, 39.48 mm wide, and 13.65 mm thick; weight is 36.3 grams. Several expended end scrapers were recovered in the same area, three of which are shown in Figure 70 (0104-Al-1, b; P106-Al-19, c; and M104-Al-61, e). The smallest seems to have been recycled to plane wood after use as a hide scraper (Abler 1997). A utilized flake made of translucent, silicified palm wood (Q106-Al-2, Figure 70, d) has microwear traces suggesting its hand-held use to scrape soft materials. A professional tanner reports using a similar hand-held flake of obsidian (see Figure 76, d) to scrape and soften the outside edges of a hide, in and around stake or lashing holes. Because these areas are difficult to reach with a hafted scraper while the hide is staked to the ground or stretched in a frame, they are often scraped and stretched separately later (David Christensen, personal communication 1997, see Janes 1983 :Figure 24). Refit and use-wear data concur that end scrapers were used at CG as fabricating implements on soft (nearly 40%) and hard (nearly 33%) materials (see also Brink 1978; Hayden 1979, 1986b:68-69; Siegel 1984:48-49). The distal corners of six end scrapers apparently served as slotting tools and four end scrapers are identified as adzes. One of the latter made of Trout Creek Jasper bears high magnification traces of hard material microwear (N 108-Al-I; Figure 71). Four additional end scrapers identified by Kay with hard material wear are illustrated in Figure 72. Two of these refit resharpening flakes and provide compelling evidence of their use on-site (Figure 72; E110-A1-4, a and T120-Al-1, d). Tool edge rounding (400X) equated with contact against soft material is documented in Figure 73. This Black Forest end scraper (0104-Al-1) refits a resharpening flake found nearby. The distribution of end scrapers (triangles) is presented in Figure 74. Circles indicate tools with use-wear evidence consistent with hide working. Seven tools refit resharpening flakes and five conjoin with broken fragments. End scrapers were recovered throughout the camp but were notably absent

230

Figure 70. End scrapers and utilized flake recovered in southwestern work area.

231

ARn

0.1 mm 200X

2 cm Figure 71. Oriented photomicrograph of hard material wear traces on endscraper used as an adz (N108-A1-1). Illustration adapted from Kay 1998:Figure 12.

232

Figure 72. Four end scrapers with hard material use-wear traces: a) E110-A2-4; b) 0104-Al-10; c) 0104A1-6; d) T120-A1-1.

233

Figure 73. Oriented photomicrograph of soft material wear traces on an endscraper (0104-Al-1). Illustration adapted fromKay 1998:Figure 7.

234

82 N

80

M

78 L

A

76

K

74 72 70

A

H

68 0

66

F

64 E

End scrapers with hide-working use-wear

62 D

60

A

54

A End scrapers with other or indeterminate use-wear

A A

B

56

A

Hide Processing

C

58

0 Other flake tools with hide-working use-wear

A

AREA

z AREA I

A

52 Y

50 x 48 W

46

Kill and Initial Butchering Area

V

44 U

42

701

T

40 S

38

Kill and Initial Butchering Area

R

36

Q

34

Work Area with Hide Processing

tN Scale = 2 meters Unexcavated

II___

100

104

108

112

116

120

124

128

132

136

Figure 74. The distribution of end scrapers (triangles) and use-wear traces consistent with hide working (circles).

235

from the kill/initial butchering areas in the southeast part of the excavation. An end scraper (T120-A1-1) recovered due west of the kill appears to have been used to scrape bone or antler (Figure 75). Most of the end scrapers were recovered in a work area situated roughly ten to twenty-four meters southwest of the kill. This work area comprises only 6% (88 m 2) of the total excavation (1434 m 2). Yet it yielded 42% of all Folsom end scrapers (26 of 69), including 50% of those judged to have been hide working tools based on soft material use-wear traces (14 of 28) and 52% of those used on hard materials (12 of 23).

Hide Processing: Seasonal Rhythms and Tools Kits Use-wear and refit data, ethnohistoric and etlmoarchaeological observations, and information from modern hide tanners form a basis for inferring hide processing activities from material remains at CG. Of 43 end scrapers studied by use-wear specialists, 65% bear soft material traces consistent with hide working. Seven other types of flake tools have similar wear. The distribution of these artifacts and the resharpening flakes that refit them indicate that skin dressing was prevalent in at least two camp locations (Figure 74). Table 57 in Appendix A summarizes the steps involved in dry-scrape, brain tanning of bison hides. This was the predominant tanning method employed traditionally by nearly all tribes with strong bison economies on the North American Plains. Schultz’s (1992:336) review of 41 ethnographic accounts (Arapaho, Ankara, Assiniboine, Blackfoot, Blood, Cheyenne, Comanche, Crow, Hidatsa, Kiowa, Omaha, Osage, Pawnee, Ponca, Shoshone, and Teton) with "adequate descriptions of hide processing tools" suggests that these tools were "morphologically and functionally uniform in all descriptions." While differences in tanning technology existed within and between tribes, these differences were "primarily in the amount of effort dedicated to a particular stage. There is enough regularity in the pan-Plains technologies to generalize a method of processing bison hides" (Schultz 1992:336). The structure and size of animal skins determines to a large extent the type of processing techniques and tools required to make rawhide (relatively stiff, untanned leather, hair removed), buckskin (soft, porous, tanned leather, hair removed), and robes (hair retained, opposite side tanned) (see Richards 1997:13-19). Bison hides, while thinner than those of domestic cattle (Frison 1991:235), are relatively thick compared with deer, antelope and mountain sheep. Consequently the former require more robust fleshing tools and carry different ramifications for end scraper attrition rates.

236

Figure 75. Oriented photomicrograph of hard material wear traces on endscraper used on bone or antler (T120-A1-1). Illustration adapted from Kay 1998: Figure 16.

237

Bison hides vary in size, thickness and wool characteristics by age, sex and season, and from one portion of a hide to another. Bull hides are typically larger and thicker with more pronounced humps than skins of females of the same age. Bull hides were favored for some items such as shields, but cow and calf hides were preferentially used for lodge covers, many containers, robes, and clothing. Many societies distinguished between animals killed during the summer when they had lost their winter coats and animals killed in the fall when the new winter coat was at its best. . . The summer bison hunts of the Plains equestrian hunters had good-quality hides as a major goal. (Driver 1990:19) The summer hide of the buffalo was called teshna’ha, meaning ’hide without hair." From teshna’ha clothing, moccasins, and tent covers were made, as these hides were easily tanned on both sides. The hides taken in winter were called meha; these were used for robes and bedding and were tanned on one side only. The hide of an old bull was preferred for bedding. (Fletcher and La Flesche 1911:272) Historically, lodge covers were replaced (on average) every year or two, usually during seasons when animal hair was not prime. Buffalo hides that we wanted to use for making tent covers, were taken in the spring when the buffaloes shed their hair and their skins are thin. The skin tent cover which we then made would be used all that summer; and the next winter, perhaps, we would begin to cut it up for moccasins. The following spring, again, we could take more buffalo hides and make another tent cover. Not all families renewed a tent so often. Some families used a tent two years, and some even a much longer time; but many families used a tent cover but a single season. It was a very usual thing for the women of a family to make a new tent cover, in the spring. (Wilson, G. L. 1987:118) Kiowa on the Southern Plains tended to replace lodge covers during the summer or fall, We made our tipis in the fall. We used them all winter. In summertime they were wearing out and we were ready for new ones. We used the old ones for linings inside the new ones, to keep out the rain. We just used them one year. (Marriot 1936-6:2) The Assiniboine on the northern Plains considered four-year-old cow hides ideal for this purpose in November (Kennedy 1961:66). Other seasonal considerations that played an important role in scheduling bison hunts in addition to hide condition were local weather patterns (Driver 1990:15-21; Frison 1996:210). Dry weather greatly facilitates jerking meat and tanning hides. The Kiowa scheduled hunts for bison hides used to make tipi covers and clothing during the summer in part because of "the good weather to work outside on them" (La Barre 1935:VI-2; 7/16/35; Bagyanoi). Rain and high humidity prolong the time required to process hides and inopportune rainstorms can spoil meat during drying. This may have ramifications for Paleoindian

238

studies if models of the onset and intense nature of summer monsoon circulation between approximately 10,000 to 9000 years ago are valid (Fall 1997:1306; Friedman et al. 1988:350-353; Thompson et al. 1993:495). During Cody times a predictable pattern of late summer thunderstorms in the Southwest may have altered the scheduling of communal meat/hide hunts relative to pre-monsoon Folsom times. Of six Cody bison kills for which seasonality is determined, only Olsen-Chubbuck occurs during the late summerearly fall season overlapping the period of monsoon rains. Others took place during late spring-summer (Scottsbluff); fall (Hudson-Meng); late fall-early winter (Homer I and II and Finley); or early winter (Carter/Kerr McGee) (Todd 1991:Table 11.1). Fletcher and La Flesche (1911: 342-345) report that, "If a rain set in just after a hunt, quantities of meat and pelts were apt to spoil, owing to the difficulty of preserving them in a warm, moist atmosphere." The use of bison hides in the construction of Folsom dwellings is conjectural but seems likely, particularly during seasons when snow and freezing temperatures curtailed travel in the Southern Rockies and Northern Plains. In pre-horse days, Plains pedestrian bison hunters used dogs to transport lodge covers and other belongings. Medium-sized dogs could transport a load of 22.7 grams

(50 pounds) which was

equivalent to the weight of a single tent cover (Wheat 1972:120 and references therein). The seasonal implications for the Folsom interval are that the manufacture of rawhide and buckskin was the likely goal of hide working during the late summer and early fall.

Both types of leather

were used for a variety of purposes including ropes, moccasin soles and containers (rawhide), and tent covers, containers, and some clothing (buckskin). Leather shirts, dresses, and leggings last many years and evidently did not have the replacement rates (for adults) of other items such as lodge covers, containers and moccasins. Early ethnographic examples of Plains clothing indicate the frequent use of elk, deer and mountain sheep hides rather than bison. Buckskin made from mountain sheep, in particular, is lighterweight yet strong, better proportioned to the human body, and not as hot as that made from bison (David Christensen, personal communication 1997). For these reasons, clothing manufacture arguably was not the primary motivation of hide acquisition at CG, with the possible exception of moccasins. Moccasin replacement rates among pedestrian peoples apparently exceeded those of many other clothing items. The sole of the moccasin is cut out of buffalo rawhide, the shape of the sole of the foot... The upper is sewed to the sole with dried, tight-twisted buffalo neck-sinew, with an

239

overhand stitch. . . Men’s moccasins last about two months or so. A man will take four or five pairs on a war party, or even up to ten pairs if the party expects to be gone long, and the man’s wife is not with him . . . women’s moccasins last a little longer than men’s about three months. (La Barre 1935: XIV-l; 8/6/35; Mary Buffalo) Buffalo Bird Woman’s account of preparations for an extended fall buffalo hunt to the Yellowstone River in 1869 provides additional insight regarding the attrition rate of footwear and the essential hide working, sewing, and meat and bone processing tools Hidatsa women transported while traveling. The family planned to travel as lightly as possible, carrying with them only the bare necessities to get through the winter. They intended to live on dried meat, corn, and vegetables they brought with them and on whatever fresh meat the men could kill along the way. Preparation for the long journey required careful planning. The women of Buffalo Bird woman’s family packed sugar and coffee, pillows, buffalo robes, skins for moccasins, and sewing thread. Buffalo Bird Woman brought twelve pairs of moccasins for her husband and herself; since both she and her husband alternated walking and riding, they would have worn out several pairs each by the time the journey was over. Besides material for moccasins, her bag held a buffalo skin with hair on it for making winter moccasins, an elkhorn scraper, a round, flat stone for sharpening the scraper, a buffalo scapula (shoulder bone) for dressing hides, an iron awl, a butcher knife wrapped in a piece of skin, a bunch of sewing sinew as big as her two palms, and a child’s cloth blanket. She tied up the bag and covered it with a buffalo robe which was used at night as a cover. Each family brought a heart skin bag for carrying water and used it as a lunch bag by filling it with the meat they would eat at noon. . .The women also packed a stone hammer, a round stone for pounding dry meat and cracking bones to make bone grease. [and] a skin tent that would house the entire extended family. (Peters 1995:133-34) The remainder of this section discusses archaeological remains at CG in light of the stages of hide processing and the tools used to accomplish these steps. Hides are fleshed as soon as possible after skinning, usually while staked to the ground, or lashed in a frame. Significant differences in the tenacity of subcutaneous tissues exist between bison (and moose) and deer, caribou, mountain sheep, or even elk. For bison, specialized bone chisel fleshers were preferred to remove this layer whereas other tools often are substituted as fleshing implements for deer-sized animals (e.g., beamers). Figure 76 illustrates the beamer (upper photograph, a) and chisel flesher (top photograph, b) used by profession tanner, David Christensen. The same chisel flesher (elk metatarsal) is compared in the lower photograph with a cast of a chisel flesher recovered from the Agate Basin site and believed to be of Folsom origin (Frison and Craig 1982:Figure 2.116). This Paleoindian tool was made from a proximal bison tibia. Another chisel flesher associated with the Folsom level at Agate Basin is made of camel bone (Frison and Craig 1 982:Figure 2.117).

240

Figure 76. Hide working tools: Top, a) Beamer for deer-sized animals, b) Bone chisel flesher, c) Hafted and unhafted staking tool used to stretch and soften, d) Hand-held scraper, e-f) Hand-held stakers; Bottom, a) Bone chisel flesher, b) Bison tibia chisel flesher possibly from the Folsom level at the Agate Basin site, c) Modern bison tibia for scale.

241

Chisel fleshers have a serrated working edge that removes adipose tissue and other membranes (hypodermis) adhering to freshly skinned hides. This is the first step in hide processing (Table 57, Appendix A). No other non-metal tool is quite as effective for fleshing bison (David Christensen, personal communication 1997), and chisel fleshers have been is use since Late Glacial times. Although a hafted end scraper can be substituted, the chisel flesher was overwhelmingly the (non-metal) tool of choice for bison among Plains groups (Schultz 1992:Table 2). Bone fleshers are noted ethnographically among all tribes except the Ponca in Schultz’s review. Although hafted end scrapers were mentioned as fleshing tools in 10 of 41 ethnohistoric accounts, 90% of the tribal groups cited also used bone fleshers and their use of hafted end scrapers was arguably auxiliary (see Hayden 1986b:68). After fleshing, hides are often dried. Due to their thickness bison hides then require further thinning to remove membranes that inhibit the absorption of tanning solutions and thereby prevent softening. This involves a greater amount of scraping than is required for deer or elk. The flesh-side of the hide is scraped and thinned whether one is making rawhide, buckskin, or robes. What happens on the opposite (hair) side of the hide depends on the tanner’s goal. If rawhide is desired, the hair and epidermis are scraped away. If buckskin is needed, scraping penetrates deeper to remove the papillary layer or "grain" and a portion of the underlying dermis (containing the deepest hair follicles; Richards 1997:14). This process is called graining (Richards 1997:15). If robes are the aim, the hair is left intact. During the late summer and early fall when hair is removed routinely to make rawhide and/or buckskin, both sides of the hide are scraped. Robes, which are typically prepared during the late fall and winter, are scraped on one side only. Thus the net amount of scraping may vary by season (discussed further below). Ethnographic examples of composite tools used to grain and thin bison hides on the Plains are illustrated in Figure 77 (see also Metcalf 1970:Figure 1 and 2; Wedel 1970:Figures 1 and 2). Antler handles such as these were heirlooms handed down among female relatives. Among Koyukon Athabaskans it is believed "that a woman’s tanning luck is transmitted with her tools, so the tools of an expert tanner are prized possessions passed down through the generations" (Mautner 1978:160). End scrapers on the other hand, were replaceable parts that experienced high rates of attrition during the

242

r$p

mg

4

Iigurc 77. 1indscrapers halted in elk horn handles, National Museum of Natural I listory. Smithsonian Institution.

243

graining and thinning stage of hide processing. This task demands sharp tool edges that are maintained through frequent resharpening (Riggs 1980:37; Schultz 1992:345). Obsidian end scrapers used for this by Ethiopian cowhide tanners are reduced ’at the rate of over 1 cm per hour’ (Gallagher 1977:411). On the average of every 100 strokes ...the scraper must be resharpened... The problem is not so much that the scraper is dull, but that it develops small jagged or uneven edges or nicks which dig in to the surface of the leather, ruining the smooth surface or even cutting it. . (Gallagher (1977:411) Usually it will require four scrapers to complete one large hide... each scraper is used and resharpened until so little of the piece protrudes from the handle that the proper angle for scraping is not possible. (Gallagher 1977:411) Illustrated examples of exhausted Ethiopian hide scrapers (Gallagher 1977:Figure 13) indicate a discard length of 25 to 45 mm, comparable to that at CG (Figure 28, Chapter 2). Gallagher’s approximated attrition rates (i.e., one cm per hour; four scrapers per hide) are considered limiting estimates in this study. Obsidian may wear more quickly than materials used at CG and the cowhides involved may have been thicker (Frison 1991:235, although, Bison bison antiquus hides were larger). Experimental hide working indicates an end scraper resharpening frequency of 100 times per bison hide (Schultz 1992:345). "Completely new bit edges were pressure flaked. . . in approximately 30 seconds" (Schultz 1992:345). Use and frequent resharpening generates an abundance of microflakes (Figure 47, d), some tranchet-like bit removals (Figure 68, a; see Shafer 1970:481-84, 1983:Figure 2c and Figure 4, Wilmsen and Roberts 1978:98), and broken tool edges (Figure 68, i). In turn, accelerated attrition leads to higher discard rates for expended end scrapers. It is postulated here that the graining/thinning stage is responsible for the greatest amount of lithic debris produced during Folsom hide processing.

Holding size and number

of hides constant, end scraper attrition may have been greatest at those sites where buckskin was being produced in large quantities because both sides of the hide were scraped as opposed to a single side. A late summer-early fall peak in hide scraper attrition rates is predicted during the Folsom time period due to processing method as well as the number (and sides) of hides involved.

If the production of buckskin

(suede on both sides) was more frequent in warm seasons than cold, then a higher number of resharpening flakes and/or discarded end scrapers might be expected on warm season sites.

As an initial test of these

ideas the discard/loss rate for end scrapers at CG (warm season) is compared with that for the Folsom assemblage at Agate Basin (cool season). Forty-four percent of the retouched flake tools in the CG block

244

sample are end scrapers (42 of 96), while 24 percent of the Folsom retouched flake tools at Agate Basin are endscrapers (14 of 57; Bradley 1982:199-201). These data are consistent with the hypothesis. For further comparison, 17% of the retouched flake tools (6 of 35) recovered at the spring-early summer camp locality at the Mill Iron Goshen site are end scrapers (Bradley and Frison 1996:Table 4.2). Use-wear information indicates that one of these tools was used on dry hide (No. 1319), another on wet hide (No. 1293) (Akoshirna and Frison 1996:77-78). Data from additional camps of known seasonality are needed to further assess this hypothesis concerning seasonal variation in end scraper attrition rates. Table 33 summarizes these data and compares the number of end scrapers, projectile points, number of bison, and area excavated at each site. The ratio of end scrapers to bison is similar at Cattle Guard and Agate Basin but both differ from Mill Iron. All three sites diverge in terms of the ratio of points to bison. Gross differences in sample sizes may undermine the comparability of these data. Likewise, the degree to which the activity-related samples excavated at each site are comparable is as crucial as it is uncertain. Camp debris and faunal remains representing the processing of a single kill event (or closely timed kills) seem to be represented in each case. Yet the degree to which the individual slices taken from each site permit accurate generalizations regarding the range of seasonal activities conducted there is by no means clear. The degree to which this holds true effects the validity of intersite comparisons accordingly.

Table 33. Relative Occurrence of End Scrapers at Three Sites. %a Site Area NMI End End Scrapers Points per m2 Bison Scrapers per Bison Bison

Seasonality

Cattle Guard

1434

49

69

44%

1.4: 1

2.1 : I

Summer- Fall

Agate Basin Folsom Level Mill Iron

247

8

14

24%

1.4 : 1

0.3 : 1

Winter- Spring

> 140

29

6

17%

0.2 : 1

1.1 : I

Spring- Summer

apercentage that end scrapers contribute to total sample of retouched flake tools per site.

Further discussion of this issue is beyond the scope of this study and will be addressed elsewhere. The primary aim here is to investigate seasonal task orientations shaping the CG assemblage itself, specifically those related to cooperative bison procurement and processing. The preliminary comparisons of the CG, Agate Basin and Mill Iron assemblages cited above are offered simply as grist for the mill.

245

Returning to a consideration of the CG assemblage, end scraper discard size exhibits an interesting spatial dichotomy. The end scrapers in the work area were discarded in a more expended state (on average) that those discarded elsewhere (Figure 78). Higher end scraper attrition rates and a greater proportion of resharpening flakes from steeply beveled unifaces appear to characterize the former relative to all other excavated areas. It may be that a greater number of hides were fleshed and grained/thinned in the former location. Freshly skinned hides are heavy, but the manufacture of rawhide makes them significantly lighter. Analogous to a flintknapper eliminating bulk and weight at a lithic quarry and then further reducing bifaces in subsequent camps, a hide worker likewise had the option of reducing the bulk and weight of skins close to the carcasses. That is, the initial stages of hide processing (i.e., fleshing and graining /thinning) required to make rawhide could be undertaken close to the kill. This stabilizes the hides against decay and makes them much lighter. At this point they could be transported elsewhere for tanning. When the tribe was on the annual hunt a certain part of the work of dressing the skins had to be done at once in order to preserve the pelts for future use... [after graining/thinning] the skin was taken home for tanning. Often a family would have a number of skins to prepare in this way when on the hunt and the women would be kept busy day and night if the hunters were successful. Not only did the skins have to be attended to at once in order to save them but the meat had to be jerked immediately, otherwise it would spoil and be attacked by insects. When the people reached home the tanning was done at the convenience of the women. (Fletcher and La Flesche 1911: 343-344) At least four or five days are required to complete the tanning process required for buckskin (La Bane 1935:1-1; 7/3/35; Mary Buffalo), but only a day or two is required to make rawhide. Containers made from various portions of the bison (e.g., intestines and stomach linings, hides) were used to transport dried meat and other items away from bison kills. Historically, rawhide containers were used to pack pemmican and dried meat and kept their contents "dry even when traveling in the hardest rain" (Ewers 1945:16). Bison rawhide is relatively heavy and canine-aided transport might be expected if many rawhide containers were utilized. In order to manufacture buckskin, tanning solutions (made from brains, fat, and other substances) must be applied and further processing undertaken to stretch, soften, and optionally waterproof the leather (Table 57 Appendix A). Additional tools are needed. Notable among these are hafted and/or hand-held staking tools. The former consists of a single, straight piece of wood appropriately shaped at the working

00 rel

end, or a composite tool comprised of a stone bit hafted in a long, straight handle (Figure 77, c). Functional requirements favor a working end that is slightly flared, round-cornered, convex and wedge-shaped (Riggs 1980:69). This tool is used to apply friction while aggressively and nearly continuously stretching a damp hide while it dries.

11

1 15-20

20-25

25-30

30-35

35-40

40-45

45-50

50-55

55-60

60-65

65-70

Discard length in 5 mm increments

El Southwestern work area El Elsewhere

Figure 78. Graph comparing size of discarded end scrapers in southwest work area with a combined sample from all other site areas.

The goal of this stage is to stretch and thereby soften the brain-lubricated fibers. At the beginning of this stage excess water may need to be stripped away in squeegee-like’ fashion. As the work proceeds and the hide begins to dry, the fibers are stretched with increasing force. Tool sharpness is neither required nor desirable during this stretching/softening stage. ’It is important to keep your staker end smooth; a wayward splinter or rough spot can gouge in to the skin" (Riggs 1980:76). This stage is often done cooperatively with two individuals working together on a single hide. For half a day one or two women will scrape and scrape a skin until whatever is loose of the mixture is scraped off, and the hide is thoroughly softened... In tanning hides relatives appear mostly to help. (La Bane 1935: I-I; 7/3/35: Mary Buffalo) In addition to hafted stakers, unhafted softening tools are also used. Figure 76 (d, e) show two such tools used by a modern tanner. A Folsom stretching/softening tool is identified in the CG assemblage. It is a large side scraper of Black Forest wood (0106-Al-12; Figure 104, m). It has a normalized,

247

retouched lateral margin. At high magnification Kay (1998) identified striae at an oblique angle to the edge that "V’ out from the direction of use. He infers that the edge was utilized to scrape medium soft to hard material. Use-wear overprints a flake scar from a retouch flake refitted to this edge. The tool and refitted flake were recovered in the southwestern work area and indicate that the tool was used, sharpened, and discarded where found. Polish and smoothing visible on dorsal arrises, but absent on the retouched edge, are interpreted as evidence of transport wear (Abler 1997). If these surface modifications resulted from wind or sand abrasion they should effect both retouched and non-retouched areas alike. Under low magnification, Ahier noted areas of step/blunting suggesting that retouch was by percussion. This is further suggested by the size of the refitted retouch flake. This side scraper may have been hafted, as suggested by a subtle change in edge configuration and the removal of a small scalar flake possibly due to torque against a handle. This edge damage is located just proximal to, and to either side of, the working edge. Similar edge modification is seen on several end scrapers in this assemblage. A black residue adhering to flake scars on the postulated hafted portion of the tool may represent mastic. The absence of heat-related discoloration due to friction on the dorsal and ventral faces (seen on a number of other hand-held tools and discussed at the end of this chapter) may indicate that the faces of this tool were bound within a protective cover during use. David Christensen (1997), an accomplished hide tanner and a museum consultant regarding Plains ethnographic clothing and tipi construction, identified this tool as ideally suited for use during the stretching and softening stage of hide processing. The edge configuration and angle (50 to 55) of this side scraper are very similar to tools he uses for this task (see Figure 76, upper photograph, c, f). If this functional interpretation is accurate, it may suggest that some hides were finished into buckskin at CG. Side scrapers with this edge configuration and soft material microwear traces were not as frequently discarded at CG, as were end scrapers. Alternate explanations may account for this. It may reflect lower rates of attrition for tools used to stretch and soften versus end scrapers used to grain and thin. It might suggest that wooden staking tools were used for which no material remains are evident. It may be that some hide working was staged between sites and net more rawhide was produced at CG than buckskin (Abler and Jodry 1997). That is, the processing of some hides stopped after the graining/thinning stage was

248

completed (i.e. the stage at which hides are stabilized, flexible enough to fold, and as light as hide working will make them). Abler noted that end scrapers at CG are strikingly different from nearly all Plains Village age end scrapers he had previously examined. Three observations are key in this regard (Ahier and Jodry 1997). Far fewer CG end scrapers exhibit wear features consistent with scraping abrasive material (3%; Type 16) than Plains Village scrapers (20%). The maximum intensity of use-wear (Type 06 and Type 16) is much lower in the CG assemblage with no occurrences of wear traces visible without magnification, while 20% of the Plains Village sample exhibits such wear. Finally, many more CG end scrapers (nearly 72%) were discarded in a relatively unworn condition (as detected at low magnification) than is the case for Plains Village scrapers (11%). From these data it appears that: 1) Plains Village scrapers were not being maintained by resharpening with the same intensity immediately prior to discard as CG scrapers, and 2) the final use of the former was for tasks other than graining/thinning where sharp edges are critical. It seems that end scraper use in alternate hide working tasks was characteristic of the Plains Village sample relative to the CG Folsom material. Staging in hide-working during both periods may be represented with earlier stages better represented at CG (Ahler and Jodry 1997). Evidence for hard material wear (Type 17) among the CG end scrapers analyzed by Ahler was 3%, while it was 14% among his combined sample of all Plains Village sites. Analyses of the Hanson site assemblage indicate that projectile point production was carried out in stages from one camp to another (Frison and Bradley 1980, Ingbar 1992), and data from CG and other Folsom sites support this. This dissertation provides new information to suggest that female activities (hide processing) were likewise sequenced between camp locations and that late summer-early fall bison kill/processing events led to heavy attrition of both male and female gear. While bone chisel fleshers and end scraper handles represent long-lived implements, end scrapers were replaceable parts that quickly wore out during early-stage hide processing. Hide workers no doubt anticipated these needs and had strategies for ensuring that appropriate stores would be on-hand at large bison kills. At least 69 end scrapers were recovered. If three end scrapers were expended per hide, then fewer than 23 hides may have been grained/thinned since not all end scrapers were used on hides.

249

Gravers and Perforators Thirty flake tools are classified primarily as gravers and perforators in the total sample, 27 in the block sample. John Tomenchuk analyzed seven of these implements for use-wear and provided preliminary identifications. Specimen El 12-A2-4 (Figure 79, b) is thought to have been used to drill soft wood or hide and consists of a thin, distal flake fragment made of Black Forest silicified wood with a long, narrow retouched tip. This functional interpretation is based on engineering estimates of service load and penetration resistance, as well as iridescent points suggesting possible use on hides (Tomenchuk 1999:5). Four compass gravers are identified. The first (Dl 12-A2-5-1983; Figure 79, g) is a double-scribe compass graver, a Paleoindian tool type recently recognized from the Early Paleoindian Fisher site in southern Ontario (Tomenchuk and Stork 1997:508). The CG example has ’a pilot (pivot) spur flanked on both sides by individual scribes equidistant from the pilot’ (Tomenchuk 1997:6). Microwear suggests it may have been used on bone or hard wood (Tomenchuk 1999:4). Two other forms of multiple spurred gravers are noted at CG. He describes these as follows: The first, a variant of the double-scribe compass graver, possesses a pilot spur unilaterally flanked by two scribes (i.e., Al 10-A2-5). The intervening web notches are equal in length. The second form is a triple-scribe compass graver (no. El 12-A2-1). The pilot spur is similarly unilaterally flaked, but in this case, by the spent remnants of three variably spaced scribes. The exceptional condition of this artifact--spent but not broken-revealed several unexpected features of compass graver design and use. The unique configurations of multiple spurred gravers found at the SCG [CG] site underscore the remarkable variability in design of this category of tool and hence its potential value in comparative studies. (Tomenchuk 1997:6) This triple scribe compass graver is made on a decorticate flake of very fine-grained chert of uncertain origin (E112-A2-1; Figure 79, d). Tomenchuk

(1999:5) infers its probable use on bone. The

circumferences scribed by the three gravers spurs are estimated to have been 27.02, 45.24, and 66.6 mm, respectively. The double-scribed compass graver (Al lO-A2-5) is also thought to have been used on bone (Tomenchuk 1999:2). Others examples of Folsom compass gravers are tentatively identified by Tornenchuk and Storck (1997:Table 2) at Agate Basin (Frison and Stanford 1982:Figure 2.19m-r, Figure 2.210, Hanson (Frison and Bradley 1980: Figure 44e), and Lindenmeier (Roberts 1935:Plate 13e, and g).

250

Figure 79. Gravers and perforators: a) F112-A2-2, b) £112-A24, c) E106-A2-1, d) E112A2-1, e) C114-A2-2, f) D114-A25-1983, g) D112-A2-5-1981, h) C112-A2-7, i) SL 10 64.

251

A final Black Forest graver examined for use-wear was recovered on the surface (SL-10-64; Figure 79, i). It exhibits a short shank coring bit and an alternate cutting edge both thought to have been utilized on hard wood (Tomenchuk 1999:5). Table 34 provides percentage frequencies of raw materials represented by gravers and perforators. Spatial distributions of these tools are shown in Figure 80.

Table 34. Percentage Frequency of Raw Material Types for Gravers and Perforators in Total Lithic Sample. Raw Material Percentage of Gravers Raw Material Percentage of Gravers Black Forest

53%

Cumbres

Trout Creek

16.7%

Chuska

Hornfels

3%

Indeterminate

10% 13.3% 3%

Flake Tools Modified by Use Only A total sample of 98 flake tools modified by use only was recovered. Ninety-five are represented in the block sample. Variation in this tool class is similar to that illustrated in the Lindenmeier (Wilmsen and Roberts 1978), Hanson (Frison and Bradley 1980), Agate Basin (Frison and Stanford 1982), Bobtail Wolf (Root and Emerson 1994) and Mitchell Locality (Boldurian 1990, 1991) reports. Figure 81 illustrates a sample of these utilized flakes including such blank types as prismatic blades (Figure 81, f), channel flake fragments (Figure 81, c), biface thinning flakes (Figure 81, b), opposed diving flakes (Figure 81, h), and hinge flakes (Figure 81, i). Eighty-six of these tools appear to have been used on a single edge, 31 on two edges, a dozen on three edges, and two on four edges. The majority of the non-retouched flake tools are made of Black Forest silicified wood (nearly 72%). Trout Creek Jasper and Cumbres chert represent 13% and 12%, respectively. Other material types each contribute less than 5% to the total sample of tools modified by use only. The spatial distribution of these artifacts in the block sample is shown in Figure 82. Material types thought to originate from source areas located south of the site are differentiated (i.e., Cumbres and Chuska). Cumbres is distributed widely across the camp area while Chuska is somewhat more concentrated in the northern half of the block.

252

82N 80M 78L 76K 74; 7270H 6864

F0

o o

F

o

E

62

Single-scribe graver

D

60

Triple-scribe compass graver

1

54 -H

-

fi Double-scribe graver

03

58

52

1 Gravers and perforators

L

AREA 2 AREA I

Yl

50-H xl 48-H WI V!

44U

F-1

T

40H SI

38-

I

36 Q I! 34-

P.

32 01 30H

DO

()

El

0

NI

28H

I

I

S

100 104

108

I

112

I

116

I

I

120

I

I

tN Scale = 2 meters

-

Unexcavated

-

I

I

I

124 128 132 136

Figure 80. Distribution of Gravers and Perforators.

-

253

Figure 81. Flake tools: a) F1112-A2-7, b) Wi 18-Al-1/Y1 14-A1-6, c) Dl 16-A2-1, d) Y112-Al-2, e) Al 18-A2-2, f) El 12-A2-15, g) Y112-A1-2, h) Bi 14-A2-5, i)D108-A2-1.

254

82--80-I

MI

+

78-I 76-s K

7472-

1+

I +++e+

70-

H

68-

+

10

0

66-i

+++

F

64-

+.f+

E

62D

60-

0

t

56-11

Retouched Flake Knife

I

+

+

Al AREA I

1 52 54-

o Retouched Side Scraper +

++

+ Flake Tool Modified by Use Only

AREA I

+ +

t+ / ++

V

42-H TI 40si 38-I 34:

2826L 24100

::

+

tN

[III Scale = 2 meters Unexcavated

104

108 112 116 120 124

128 132 136

Figure 82. Distribution of flake tools modified by use only and retouched flake knives and side scrapers.

255

In the following section, I describe the occurrence of heat alteration attributed to friction that modifies the color of Black Forest artifacts from yellow-brown to red. Minimally, this modification has the potential to inform efforts to distinguish between hafted and non-hafted implements in the CG assemblage.

Occurrence of Red Coloration on Black Forest Artifacts Judge (1973:112) identified "reddened" areas on gravers and other sharply projecting utilized edges on Paleoindian tools from the Central Rio Grande Valley in New Mexico. This reddening occurred on 21 Folsom artifacts (or 2% of the total Folsom sample) but was absent from Clovis and other Paleoindian artifacts. He documented reddened areas on six knives, five gravers, four end scrapers, two point ears, one channel flake, one preform tip, one side scraper and one utilized flake. Most of these tools were made of brown or tan jaspers. Both Crabtree and Judge (1973:113) concluded that the reddening was due to oxidation of iron in the raw material. Crabtree thought this resulted from deliberate heat treatment of tool blanks and suggested "that the reddened tips represent those areas of the raw material which were near the surface of the heating pit, and were thus oxidized to yield the isolated red color’ (Judge 1973:113). Noting the consistent association of red coloration with utilized tips, Judge suggested that the working elements of finished artifacts (rather than tool blanks or preforms) were deliberately heated to harden them for heavy incising. Reddening occurs as a distinctive trait on 63 tools and 13 resharpening flakes from steeply beveled unifaces at CG (Table 35). All but one of these artifacts is made of Black Forest silicified wood, an ironrich, yellow-brown jasper. The exception is a thin biface-thinning flake of Trout Creek jasper with two sharp graver tips that are reddened. During tool analysis the incidence of reddening was recorded separately for platforms, dorsal ridges, working edges, dorsal and ventral faces, and as natural occurrences of red color within the raw material. The following patterns are noted. Reddening is strongly associated with "ground" (i.e., polished

sensu Titmus and Woods 1991)

platforms. Nearly 35 percent of all Black Forest tools which retain platforms (28 of 79) exhibit reddening in direct association with the polished areas. It is significant that the reddening does not extend to portions of the platform that are not prepared by polishing. Similar reddening is present on prominent dorsal ridges

256

of five unifacial tools. It is also coincident with discontinuous patches of use-wear on the edges of four end scrapers that appear to have been used on hard materials such as wood, antler, or bone. In these latter cases, red is visible at low magnification (lOx) as a narrow line along the leading working edge at the juncture of the ventral and dorsal tool margins. Thirteen resharpening flakes (recovered from a 12 rn 2 portion of the southwestern work area) exhibit reddening along the working edge and demonstrate that this modification is removed during the rejuvenation of tool edges. Heat generated by friction and/or abrasion seems best to account for the red coloration present on ground platforms and scraper edges. I began recording this attribute systematically late in the analysis of umnodified flakes, and the thirteen resharpening flakes cited here represent a minimum sample of those present in the total assemblage.

Table 35. The Occurrence of Reddening on CG Artifacts. Graver/ Other End scraper Perforator’ Uniface Number (%) of 405 Tools

16(3.9)

Number (%) of 229 16(6.9) Bik Forest Tools

Ultrathin Preform

Resharp. Flake

Total

(%)C

76 (100)

5 (1.2)

41(10.1)

4(l.7)

41(17.9)

1

13

1

13

75 (98.6) 29(38.1)

Platform

14

Working Edge

4

8

11

Face

1

1

22

24 (31.5)

Dorsal Ridge

1

4

5 (6.5)

Red in Material Naturally

1

8

9 (11.8)

13

45(59.2)

’Includes the only non-Black Forest tool with reddening, a Trout Creek flake with two reddened tips. bThjs represents a sample from three 2x2-meter units in the southwest work area. ’Individual artifacts may exhibit reddening in more than one location. The reddening on eight thin, delicate graver or perforator tips is confined to within 1.5 to 2.5 mm of the distal tool margin. Its cause on these particular tools is uncertain. Experiential use of a thin (1.12 mm) Black Forest flake edge to incise a groove in hard material (dry antler) failed to produce any reddening although the tool edge felt warm to the touch. Direct exposure of this edge to a hot flame for 5 to 10 seconds was sufficient to redden it and to significantly heat the stone tool. If Judges idea of deliberate heat treatment of sharp pointed tools is correct, it is possible that the heat itself was a desirable

257

attribute that facilitated some tasks such as cutting through mastics or incising/drilling certain materials. In my cursory experiments, exposure to direct flames for longer than about 20 seconds caused thin (1-2 mm) Black Forest tool edges to explode. Another suite of red modifications appears on the broad faces (dorsal and/or ventral) of an end scraper, a graver and 22 other unifacial tools. These consist of short, narrow, red lines that appear to be heat streaks produced by high pressure point contact (Ahler 1997, notes regarding tool P102-A1-3). Their orientation appears to be variable on a single tool and between tools. In a few instances microscopically visible pitting (25 to 40x) seems to be associated with these red streaks. Their possible origin(s) in the realms of use- or transport-wear begs further study. Bradley reported that large flakes and bifaces he produced and then transported without padding exhibited "fine, minute polish spots" on their faces that resembled modifications on the Penn Cache Clovis artifacts (Frison and Bradley 1999:81). The red streaks on CG implements hold clues to tool histories that have not yet been unraveled and may help distinguish between hafted and non-hafted artifacts.

Concluding Remarks Chapter 5 described the CG stone tool assemblage and made inferences regarding the manner in which functional requirements of bison procurement and processing are related to aspects of Folsom technology. Predictable animal aggregations during the late summer and early fall when cows were relatively fat and their hides in appropriate condition for the manufacture of rawhide and buckskin enabled Folsom groups to hunt cooperatively and productively during this season. Is appears that the constraints of processing meat, marrow, hides, sinew and the like from a large number of animals in a short time frame drove a small subset of the tool kit in an efficiency optimizing direction. Folsom projectile points, ultrathin bifaces, chisel bone fleshers and end scrapers are seen as effective and specialized implements designed to facilitate these tasks in a way that conserved raw material for people who evidently traveled widely during the summer and fall seasons. The use of specialized tools strongly implies that Folsom groups anticipated making both large and small bison kills on a recurrent and sufficiently predictable basis to warrant the development of time-energy-material-saving implements (Torrence 1982, Hayden and Gargett 1988).

258

Folsom planning depth arguably incorporated expectations that tool attrition rates would vary by season. Large late summer-early fall hunts called for items of personal gear that were made ahead of time such as weapon tips, bifacial knives, and bone and antler portions of scraping tools. Additionally, a wider array of more generalized flake tools and expediently fashioned cobbles used as bone-cracking equipment needed to be on hand but did not require an equal investment in manufacture. The predictable, recurrent nature of bison hunting and the large scale of processing during seasons of high residential mobility are seen here as key elements shaping the organization of Folsom technology. A division of tasks along gender lines (and within genders by age and levels of relative skill) is postulated. During these big summer-fall kills, the demands of meat and hide processing led to higher attrition rates for ultrathin knives and stone scrapers (females-dominated and/or female-specific gear) and weapon tips (male-specific gear). It is postulated by this author that ultrathin bifaces were designed primarily for processing meat and constituted a tool form recognized culturally as a "woman’s knife" and used predominately by females. The advanced flintknapping skills required for the manufacture of Folsom points and ultrathins are expected to have been possessed predominately by males. The ability to produce stone bifaces with width to thickness ratios in excess of 15 to 1 demands a commitment in time to develop. This is equally true of the advanced skills required to manufacture sophisticated items of soft technology such as clothing and tent covers capable of keeping people alive during cold winters. These competencies on which all members of the group depended for survival are suggested to have been complementary, engendered skills in Folsom society. In this model young females spent considerable time developing skills related to hide processing and leather working, plant identification and processing, small game procurement, large game processing, child birth and infant care, religious ritual, and the maintenance of social alliances, while also mastering the basics of flintknapping. It is expected that females made many but not all of their own tools and maintained and repaired most of them. Conversely, young males are expected to have spent considerable time developing skills related to tracking, large and small game hunting, biface manufacture, landscape navigation, religious ritual and the maintenance of social alliances, while also mastering the basics of clothing repair, and food and hide processing. The primary economic unit is envisioned as an adult female and male who together produced the social and ritual relations, materials items, subsistence resources and

259

children that ensured group survival. The relations between individual males and females may have been quite fluid, but the complementary division of socioeconomic responsibilities along lines of gender and age is believed to have bound female and male roles together in ways that were pervasive and enduring. Increasing evidence indicates that Folsom groups were making large bison kills with sufficient regularity to warrant the manufacture of tools geared to this resource and their planning depth with regard to these activities was evidently extensive. They developed a small suite of specialized tools that were easily maintained, conservative of raw material, highly effective for their primary intended tasks, and beautiful as well. These were added to a tool kit otherwise characterized by multipurpose implements of more general utility. The following examples highlight the development of specialized tools to optimize modern economic pursuits in Vancouver: .as the amount of fish processed increases, there is a marked tendency for the number of different types of knives being used in the process to increase. At the extremes, full time fish processors use up to seven different types of knives in their work while it is rare for the occasional sports fisherman to use more than one. In this study, exactly the same range of operations were performed by the occasional sports fisherman and the most intensive fish processing plants. These operations included: heading, tailing, gutting, filleting, and steaking of several different types and sizes of fish in each case. In the most intensive processing operations different knives were sometimes used to perform the same task on different types of fish. Most of us are familiar with similar phenomena in everyday life around us. Full time carpenters have an entire array of different types of hammers and screwdrivers, whereas home carpenters have only a few. The principle can be extended to plumbing, butchering, electrical work, wood cutting, or any other domain in which tools and high processing volumes are important. (Hayden and Gargett 1988:16) Some Paleoindian groups such as Clovis (Kay 1996b:315), Cody (Wheat 1979:Figure 41), and Dalton (Goodyear 1974:32) employed hafted points that were switch-hitters -- that is, they could be employed as projectile points and as knives. In contrast, Folsom points were arguably designed exclusively for killing with no intentions or provisions for use in butchery. These are two very different technological approaches with advantages and disadvantages that may be understood within varying parameters of planning depth regarding the pursuit, capture and processing of game, and the frequency and anticipated range of conditions under which these would occur. Very thin, alternately beveled knives appear among several, but not all, bison hunting groups. I would suggest that times and places where these knives appear in the archeological record of bison hunters is a clue to the frequency of very large kills made under circumstances when processing time is constrained

P1411

by season/climate or labor force. Additionally and/or alternately, specialized tools may point to varying cultural practices whereby the potential number of adult processors in a given group is constrained by engendered divisions of certain processing tasks. The designs of chisel fleshers and some hafted end scrapers are functionally specific for the primary tasks of fleshing and graining/thinning hides, respectively. The latter have been ubiquitous items in the tool kits of many hunting groups since the Upper Paleolithic when they begin to account for 10 to 40% of most assemblages (Bordes 1968, cited in Hayden and Gargett 1988:16). End scrapers are not ubiquitous because they were generic tools, but rather because they were ideal, specialized tools for key, frequent tasks that resulted in high tool attrition rates. One of the logical outcomes of this emphasis on specialized tools is that bands producing them should produce activity-specific types of assemblages where specialized economic resources are being exploited intensively. (Hayden and Gargett 1988:16) The archaeological material at CG is interpreted here as an assemblage that is activity-specific with regard to late summer-early fall bison procurement and processing. This does not preclude a range of other activities undertaken while people lived in this camp, rather that the overall cast of the assemblage is strongly shaped by the engendered tasks accompanying the mass processing of bison during this season. In contrast, I propose that the Folsom assemblage at Agate Basin differs in important ways. Its seasonal emphasis (winter-early spring) with regard to hide working suggests a shift to dressing skins for the manufacture of robes and clothing. This is suggested by the recovery of a high incidence of needles (one complete and 16 fragments; Frison and Craig 1982:165-66) and debris from their manufacture as well as by the representation of a range of prime pelt producers (including red fox, wolf, coyote, elk, rabbits, and thirteen lined ground squirrels (Walker 1982:284-297; see Soffer 1985:340). A single needle midsection was recovered at CG and no culturally modified bones from small mammals were found despite the recovery of unmodified bones from small mammals. Differences in the proportional representation of end scrapers and stone projectile points between the two sites were noted previously in this chapter. The following chapter describes the spatial distributions of cultural material at CG and discusses the ramifications of these for the functional and social organization of activities.

261

Note to Users

Page 261 missing in number only; Text complete and follows on page 262.

CHAPTER 6 ACTIVITY PATTERNING AND SITE FORMATION Introduction The highly informative Folsom component at the Agate Basin site provides a detailed view of activities conducted in a late winter-early spring camp associated with a bison kill (Frison 1982a). Stewart’s Cattle Guard provides a complementary view from a processing camp associated with a late summer-early fall kill. The procurement of upwards of fifty animals during the latter season seems to be represented at the sites of Lipscomb (Hofinan etal. 1989; Schultz 1943; Todd et al. 1992), Folsom (Todd et al. 1996:173); and Cooper (Bement 1999:172) in addition to CG. Planned operations of this scale involved ’an integrated system of travel, preparation, and logistics preceding kills and the intricacies of butchering, processing, and distribution following kills’ (Brumbach and Jarvenpa 1997:429). Kinship and other alliances are thought to have provided a social framework that facilitated wideranging mobility and cooperative approaches to bison hunting and processing. The presence at CG of two suites of raw materials from sources located north and south of the site, respectively, may be evidence that people came together at this location for a communal hunt. It is suggested that following successful procurement of no less than 49 bison, people set up camp beside the kill for the duration of processing, and perhaps longer to enable people to socialize. This chapter discusses the spatial distribution of cultural material highlighting activities conducted in three functionally related settings. The first was a scene of initial butchery in the vicinity of the bison carcasses (Figure 6. 1, a). The second appears to be a special use area roughly twenty meters away where hide working and a variety of other tasks took place (Figure 6. 1, b). The third area is a residential camp where portions of carcasses were extensively processed, weaponry repaired and manufactured, and other activities conducted in household activity areas near hearths and possibly shelters (Figure 6. 1, c).

262

263

1

0

Bison ?rocesing East

0 GEND I Block sample Unexcavated I Not Included in Activity Area Analysis in this study

SCALE tN 0 = 2 meters

100 104 108 112 116 120 124 128 132 136 140 144 148 152 156 160 164 168 172 176 180

Figure 83. Distribution of general areas of activity at Stewarts Cattle Guard site.

264

Analysis in this chapter is guided by the hypothesis that Folsom groups divided some procurement and processing tasks along gender lines during large-scale hunts. Several researchers have summarized circumstances favoring gender and age-based divisions of labor among hunter-gatherers (e.g., Hayden 1981:402-407; Jochirn 1981:128-133; Kelly 1995:262-270; Watanabe 1968:76-77). A compelling implication of these data is that engendered divisions of labor are pronounced among groups that regularly hunt big game (Hayden 1981:405; Watanabe 1968:77). Among Plains and Subarctic peoples, females traditionally took charge of skin dressing, the majority of thin-cutting of meat for drying, as well as mass processing of bones for the recovery of marrow and/or grease. This was especially true when kills were processed in bush camps’ away from villages (e.g., Brumback and Jarvenpa 1997:431; Fletcher and La Flesche 1911:274, 342; Janes 1983:71, 99-100). At large bison kills, I am assuming that Folsom females were primarily responsible for fleshing, graining/thinning and softening hides, and for filleting much of the meat that was dried. (see also Amick 1999:171-72; Davis and Ureiser 1992:266; and Judge 1973:207). This does not necessarily preclude a less pronounced division of labor in other circumstances, as when a family butchered small kills on their own during times when a band was dispersed. The manufacture, use, and maintenance of Folsom points are assumed to have been male-specific activities. Etimohistoric and ethnoarcheological accounts strongly indicate that the dispatching of large game is "usually men’s work" (Kelly 1995:267) and may represent one of the few gender-specific tasks for males among hunter-gatherers groups (Watanabe 1968:77). Distinctive debris from point production and discard has long provided a readily identifiable record that enables researchers to investigate the mobility patterns of hunters and/or flintknappers (discussed in Chapter 4). In contrast, the material records of females, non-hunting males, and children have been more elusive. The importance of women’s processing and "transformation" skills, or the conversion of animal carcasses to edible meat, clothing, and other usable products, is too easily overlooked by researchers. . . In archaeological interpretations, the focus of attention is often on the kill. (Issac 1995:3 in Brumbach and Jarvenpa 1997:417) A concerted effort is made in this study to identify archeological remains of hide processing as a means of gaining insight into female patterns of organized labor, lithic procurement, and mobility. An attempt to identify preforms produced by novice flintknappers (possibly adolescent males) was discussed in Chapter 5.

265

Chapter Organization Data presentation in this chapter begins by summarizing the bison assemblage in terms its distribution, degree of articulation, inferred cultural breakage and the dental evidence for seasonal determination. Weathering patterns were described in Chapter 3 along with observations regarding carnivore scavenging (Jodry and Stanford 1992:Figure 4.8). Next, the distribution of lithic and faunal material in the kill/initial butchering area is discussed. Cross mends of projectile point fragments between the kill and the camp strongly support an inference of contemporaneous use of these areas Following this attention turns to analyses of spatial patterning of cultural materials in the block sample (Figure 37). Both visual and quantitative methods are used, and the resulting information is illustrated in a series of plan maps. Spatial analysis at CG is ongoing and additional studies are planned. For the purposes of this dissertation the archeological signatures of hide working, marrow recovery, and weaponry repair and replacement are investigated. Technological analyses of tools and flaking debris buttressed by refit and use-wear studies provide a basis for interpreting the archeological remains of these activities. Implications for planning depth with regard to labor force requirements and anticipated rates of tool attrition and replacement (especially for end scrapers, ultrathin bifaces, Folsom points and unifacial flake knives) at mass kill and processing sites are addressed in the concluding section.

Faunal Assemblage Nearly 3,500 bison bones and fragments were recovered from the Folsom component during twelve field seasons between 1981 and 1996 (Figure 84). A minimum of 49 animals is determined from left astragali (Figure 85). The minimum number of elements (MNE) for a sample of 1,454 bison bones from the block sample is presented in Table 36. These estimates are based on the most abundant, nonoverlapping portion or landmark represented per element/side. The identification codes (element/portion/segment/landmark) used during faunal analysis follow Todd (1995) and were recorded in part on forms he designed. The block sample contains 1,695 bones/fragments that minimally represent portions of 30 bison and a single wolf or large dog. Jodry (1987) and Jodry and Stanford (1992) provide a detailed description

1

A

W

fl

U

Q 5AL101 Stewarrs Cattle Guard Site

o

024

N

Meters

El

El

N

..

.

[’

El

G F E

’4

.

. .’ J.

El A Z

AREA AREA

EM W V U T S R z v-

-

: .,.

,.

El

.

C

o

N M

::’ .. 102

106

,,........ 110

114

118

122

N M 126

130

134

Figure 84. Horizontal distribution of the Bison Bones.

267

1

oo

oft1 T1

92i

13

88~

c

D li

13

84~ 8c

~ 13

r

72 68

1

I J0

0

64F1

Minimum Number of Animals = 49 S

Left Astragalus

0

56 AREA 52 AEA I AOX

r

S.,

44

r

IVI ’S S

0

SCALE

Q

IN = 2 meters

32 j

II

I

I

I

I

I

I

100 104 108 112 . 116 120 124 128 132 136 140 144 148 152 156 160 164 168 172 176 180

/

Figure 85. Distribution of left astragali used to calculate the minimum number of bison.

268

Table 36. Minimum Number of Bison Elements (MNE). Element Left Right Element Left Scapula 2 1 Femur 4 Humerus 6 6 Tibia 8 Radius 2 3 Metatarsal 9

Right 3 4 4

Element Innominate Patella

Left 1 2

Right 2

Note: Preliminary data from 1454 bone fragments in the block sample, with the exception of the astragalus which represents all such elements from the site (n= 3,512).

of a subset of these bones (n=393) recovered from a 246 in’ areaexcavated in 1981 and 1983 (Figure 15). While a full reporting of the faunal data is beyond the scope here, selected attributes of this assemblage are discussed and compared between site areas. The methods used to determine seasonality of bison kills by examining patterns of tooth eruption and wear were developed by Frison and Reber (1970; see also Reher and Frison 1980). Although some tooth rows were recovered at CG, either the individual is too old to assess age or the occlusal surfaces are preserved insufficiently. Seasonality is estimated from patterns of differential wear on individual cusps of isolated second and third molars from six different individuals (Table 37). While small, the sample is internally consistent in indicating late summer-early fall bison mortality. The age groupings listed below were identified by the author in consultation with George Frison and with reference to detailed descriptions and illustrations of tooth eruption and wear patterns at other Paleoindian sites (Reher 1974:115-116; Todd and Hofman 1987:508-5l8; Todd etal. 1990:816-821; Todd et al. 1996:151-173).

Table 37. Samole of Mandibular Molars used to F.stiniate (ii Spanna1itv Specimen Number Tooth Age Group (Todd et al. 1990) G 16-A2-41 M2 Group 2 R138-A2-2 M3 Group 3 El I 0-A2-5 M3 Group 4 P134-A24 M3 Group 4 R134-A2-26 M3 Group 4 S140-A2-26 M3 Group 4

Season Late summer-early fall Late summer-early fall Late summer-early fall Late summer-early fall Late summer-early fall Late summer-early fall

Comparative data regarding minimum number of bison and their dental ages are presented for nine Folsom components at seven localities in Table 38. The Lipscomb and Folsom sites are both thought to represent single event kills and the number of bison recovered at each is comparable to CG. An arroyo trap

269

was used at Folsom and Cooper, but no evidence of a natural containment feature is known from Lipscomb (Hofirian et al. 1989:171), CG or Linger. The circumstances by which people were able to successfully kill bison at the three latter sites are uncertain. Strategies of stalking and ambush may have been employed. The rolling dune topography of the eastern San Luis Valley can be used to great advantage to bring hunters close to herds. Such is the case today when stalking bison in this environment. CG is situated presently in a parabolic dune, but this feature is of recent origin. Former dunes may have provided more cover for the hunter than containment of the hunted. It may be of note here that Assiniboine groups sometimes hunted bison by surrounding them on foot, essentially forming a human enclosure (Kennedy 1961). The range of Folsom hunting tactics is expected to have been extensive and sophisticated.

Table 38. Dental Ages and Estimated Season of Death at Folsom Bison Kill and/or Processing Sites Site MINI NI Spa Dental Ages Season Reference Agate Basin

Ca.

N+0.9

Late winter early spring Late summer early fall

Frison 1982: 258; Zeimens 1982:227 This study

Late summer early fall

Bement 1999:127

Late summer early fall

Bement 1999:127

Late summer early fall

Bement 1999:127

-

60

Cattle Guard

49

Cooper upper kill

29

Cooper middle kill

29

Cooper lower kill

20

Folsom

54

33

N+0.4 to 0.5

Early fall

Todd et al. 1996:Table 8.15

Lake Theo

12

7

N+0.4 to 0.5

Early fall

Linger

5

1

Harrison and Smith 1975:16; Todd et al. 1996:Table 8.15 This study

Ca.

N+0.3 to 0.4

-

N+0.3

-

-

N+0.3

-

-

N+0.3

-

-

N+0.3 to 0.4

Late summer early fall Lipscomb 55 18 N+0.3 to 0.5 Late summer Todd et al. 1990:Table 2; early fall Todd et al. 1996:Table 8.15 Number of individual specimens used to determine seasonalilty. bSthgle calf mandible and two near-term fetal bison. ’Five M3 with differential cusp wear; one partially erupted, unworn M 2 (Group 2 [n= 11; Group 3 [n1]; Group 4 [n=4]). dSingle M1 Ca.

-

-

As discussed in Chapter 3, skeletal part representation at CG was effected by density-mediated bone destruction resulting from a combination of human action, carnivore scavenging, and weathering. The relative tenacity of element portions and articulations predicts the sample of bones preserved (Burgett

270

1990; Jodry and Stanford 1992:131-132; Kreutzer 1992). Despite attrition a significant amount of information can be ascertained from the surviving portion of the CG assemblage by means of attributebased analyses of spatial patterning and cultural modification. A fairly consistent pattern of subaerial weathering across the site suggests that the faunal assemblage shared a similar weathering history and was deposited on a single surface at, or close to, the same time. Desiccation cracking apparently began prior to burial and continued post-burial due to overburden compression and alternating soil moisture regimes. In describing the natural deterioration of bison carcasses at Wind Cave National Park, Burgett noted that fluids produced during decomposition caused a die back in vegetation that: may extend from 5-50 cm beyond the margins of the rumen mat . . .this area may remain totally devoid of vegetation for as long as two years. Then forbs and grasses begin to recolonize the ’dead zone". . . Plants that grow on these dead zones are generally taller than conspecifics located nearby and are also much greener. In spite of the more luxuriant appearance, grazers do not utilize plants in these areas for some 2 - 4 years after the death has occurred. (Burgett 1990:153) If vegetation die-off was associated with butchered carcasses and rumen mats at CG (and

Bison bison

antiquus avoided grazing such areas), then this might effect the likelihood (and timing) of subsequent bison

kills at the same location. Weathering patterns across the site strongly suggest that the bones experienced similar pre-burial histories. If two to four years separated kill events, it seems unlikely that such consistency in weathering patterns would be the case. As noted by Rapson (1990:200), ’other things being equal, bones deposited during a single occupational episode should exhibit a distinctive weathering pattern when compared to bones deposited on the same surface at other times." In addition to the dominance of a single prey species that exhibits a consistent weathering pattern other lines of evidence are consistent with an interpretation of a single Folsom occupation at CU. These include: 1) complementary patterning in faunal and lithic assemblages in different areas suggesting functional integration of tasks, 2) refits connecting activity areas and 3) the presence of similar lithic material types (including exotics) across the site (see Storck 1997:177 for similar reasoning at the Fisher site). The cultural modification of the bones and their degree of articulation are considered next. In an analytical sample of 1,695 bison bones/fragments, 10% exhibit green bone fractures (n173) and 2.5% retain impact marks (n=42). The vast majority is disarticulated (96% of 3,512). There are a total

271

of 27 articulations across the site involving 74 bones (Figure 86). Fifty-six percent of these were recovered in the kill/initial butchery area, which contains nearly 20% of the total bones excavated. Thus, the percentage of articulations in the kill/initial butchery area is higher than expected given the sample size. Tarsals are involved in 67% of the articulations (Table 39). Those recovered represent 16% of those originally present in 49 bison (see Burgett 1990:163; Lyman 1994:153 for discussion of this calculation). In 6% of the cases tarsals are articulated with a tibia, in 7% with a metatarsal. Actualistic studies of disarticulation due to weathering and carnivore scavenging indicate that tarsal articulations are among the most long-lasting and durable in the bison skeleton (Burgett 1990:171, Figure 4.5). In general, the sequence of non-cultural disarticulation in bison begins with the forelimb, proceeds with the hindlimb, followed by axial segments (Burgett 1990:177: Binford 1981:42-44).

Table 39. Representation of Bison Bone Articulationsby Carcass Segment. %b Articulation # Mean a Articulation

#

_________ Mean a%b

Lower Hindlimb

18

3

66.7

Axial

2

4

7.4

Tarsal Tibia-tarsal-metatarsal Tibia-tarsals Metatarsal-tarsals

5 1 5 7

2.4 4 3 3.7

18.5 3.7 18.5

Lumbar vertbrae Inn ominate- vertebrae

1 1

4 4

3.7 3.7

25.9

Phalanges

2

2

7.4

Lower Forelimb

3

3.5

11.1 Upper Forelimb

2

3

7.4

Metacarpal-carpals Carpals

2 1

3 4

7.4 3.7

2

3

7.4

Humerus-radius-ulna

Total=2 7

aMean number of bones per articulation type. bPercentage each articulation type represents of all articulations recovered

Upper forelimb joints preserved at CG represent 2%, and lower fore limb joints 3%, of the possible articulations from 49 bison. Lower forelimbs comprise 11% of all articulations recovered. Articulated axial elements are limited to a segment of four unfused, lumbar vertebrae and a complete pelvis articulated with lumbar vertebrae and a small part of the sacrum. Each of these articulations comprises 2% of their potential number in a population of 49 animals. Together they comprise 7% of all articulations recovered. Bison lumbar articulations persist longer than those of thoracic or cervical vertebrae (Burgett 1999:167).

272

p

o

0

0 0

o

0

Cranium

0

0

Articulations n=27 0 Lower hind limb Lower fore limb

0 AREA 2 AREA I

0

+ Upper fore limb

4 lumbar vertebrae

Phalanges Axial articulations or skull Cranium.

o 0’

At Pelvis and vertebrae

4

IN = 2 meters

SCALE

0

100 104 108 112 116 120 124 128 132 136 140 144 148 152 156 160 164 168 172 176 180

Figure 86. Distribution of bison bone articulations.

273

The implications for CG are that a somewhat greater number of articulated bones may originally have been discarded than survived the processes of pre-burial weathering and scavenging. However the higher incidence of articulations in the kill relative to other site areas does not appear to be explained by sample size, or histories of weathering and scavenging. This suggests that extensive butchering in the residential part of the camp is strongly contributing to the higher incidence of disarticulated bone found there. Further support for this idea is drawn from the distribution of green fractured and impacted bones, which are more abundant in the residential camp than in the kill/initial butchery area (Figure 87).

Kill and Initial Butchering Area There are two distinct clusters of bison bone in the kill/initial butchery area that are identified visually (Figure 84) and by density contouring and K-means clustering techniques (discussed later in this chapter). The more southerly of the two is characterized by a greater incidence of impacted and greenfractured bone (Figure 87). Both clusters are associated with Folsom points (n=15) and resharpening flakes (n=65). Table 40 provides the breakage types and raw materials for this weaponry subassemblage and Figure 88 shows its distribution. The only other implement found in direct association with the bone was a proximal fragment of an Edwards chert flake tool (refitted in three pieces). Its distribution is seen in Figure 82. This portion of the site was shallowly buried when excavated (15-20 cm or less of overburden) and it is likely that collectors removed some surface artifacts from this area.

Table 40. Projectile Point Breakage and Material Types in the Kill/Initial Butchery Area Breakage Type Number Breakage Type

Number

la Complete point

2

3a Unground tip

2

2a Basal ear

1

3b Ground tip

3

2b

1/4 break

1

3c Tip and midsection

2

2d Complete base

2

4

Midsection

5

Lateral edge

Material Type

Number

Material Type

I Number

Edwards

I

Black Forest Wood

4

Cumbres

5

Morrison Quartzite

2

Trout Creek

2

Uncertain chert

274

N

80-

Bone incompletely analyzed for breakage

M

78L

76

K

Bone incompletely analyzed for breakage

7472

1

0

70-

Bone incompletely analyzed for - brea kage

0

-

0

-

H

68-

NO

[:]

G

66-

Q3

1D

64

10

E Green fractured bone

-ft

D

60C

A

o,

E1

58B 56 54-

Impacted bone

CID

E

62-

0 00

0 0

AREA I

z AREA 1

52-

o

0

Y

50x 48w 46V

44U

42-

D

I

40-

Ejh C3

OR

380 0

R

36Q

0

34-

rel

x OR 19 0

P

32-

o

I

I

30-

tN Scale =2 meters

N

28 26-

Unexcavated

L

24i 100

o

0 0

I

I

104

I

I

I

108

112

I

I

116

I

I

120

I

124

128

132

Figure 87. Distribution of impacted and green fractured bison bone.

136

275

BISON KILL and ’BUTCHERING: AREA ii ...

c

,w. ,. ,

I-

0

4.

;

-

-.,

I-

S

118

-

- - 4.-

AHL

122

Folsom Point

-

2

-

130

126

I

L

134

j Scale =2 meters

Figure 88. Distribution of bison bone and Folsom points in the kill/butchering area.

276

Projectile point fragments conjoined between kill and camp provide compelling evidence that the two areas were contemporaneous (Figure 89). One refit consists of an impacted damaged tip (T124-Al.-1; Figure 90, b) broken from a base (Figure 49, e) recovered nearly nine meters to the northwest. The two pieces reconstruct a complete point (Figure 46). A nearly complete point of Cumbres chert is reconstructed from a small tip fragment that refits a base (E116-A2-4) recovered 14 meters away in the residential camp. This refitted specimen is illustrated in Figure 53. A small lateral fragment of silicified wood (W132-A1; Figure 92, a) is conjoined with another edge fragment (1l16-A2-2) recovered nearly 28 meters away at the edge of a reconstructed household activity area. Finally, a midsection (Xli 8-A2-2) and quarter-base fragment (Cl 10-A2-l) of translucent fossil wood are conjoined across a distance of nearly 10 meters (Figure 49, t). Collectively, these refits link both of the kill area bone clusters to residential portions of the site. Figures 90, 91 and 92 illustrate nine of the fifteen points and fragments recovered in the kill/initial butchery area. Raw material types from geographic areas located both north and south of CG are represented in the kill, including a point and a flake tool fragment of Edwards chert. The following section discusses the spatial methods employed to identify distribution patterns.

Pattern Recognition Studies Piece-Plotting and Density Contouring Four methods were used to define spatial patterns. The first and most direct approach was visual examination of scatter plots depicting different attributes and classes of lithic artifacts and skeletal elements. Most of these distribution maps were generated by the author from an Access 97 relational database linked to a grid-based contouring and surface plotting graphics program (Surfer, version 6.0, Keckler 1997). Additionally, Marcia Bakry (illustrator, Department of Anthropology, Smithsonian Institution) produced a composite illustration of piece-plotted bone by digitally combining measured field drawings from individual two-meter excavation units (Figure 84). This map depicts an extensive, discontinuous scatter of primarily disarticulated and fragmented skeletal elements. Several bone concentrations are evident, as are a few circular areas (nearly two meters in diameter) where bones are conspicuously absent (e.g., transects along lines E and F in Figure 84). The

277

N

78L

76-

K

747270H G F

Fee

56 A

54

AR

z AR

52 Y

50 x 48 V U

42T S

38R

36p

32-

0

30N

28M

26-

.J

100 104

108 112 116 120 124

uAla y a L.L

128 132 136

Figure 89. Folsom point fragments conjoined between kill/initial butchery and camp areas.

278

Figure 90. Kill area projectile points (K-means bone cluster 2; a) S130-A1-1, b) T124-A1-1, c) S126-Al-1.

279

Figure 91. Kill area projectile points (K-Means bone cluster 2, b-c): a) W120-Al-1, b) P122-A1-3, c) Q122-A1-1.

Mrs

Figure 92. Kill area projectile points (K-means bone cluster 6); a) W132-A1-1, b) W134-A1-1, c) V128-A1-2.

281

incidence of marrow cracked bone appears to be concentrated at the perimeters of these circular areas (Figure 87) and cobbles used as bone cracking equipment (Jodry and Stanford 1992:Figures 4.22, 4.23, 4.24, 4.25 and 4.26) also seem distributed outside their circumference (Figure 93). The possibility that these may represent the former locations of structures is intriguing, although alternate hypotheses are that hides were staked to the ground in these areas, or bison carcasses were lying there. One (and perhaps two) circular structures are postulated for the Folsom level at Agate Basin, their exact diameter is not certain. Bison ribs apparently used as stakes were associated with them (Frison 1982:Figure 2.18), as were an interior hearth, flake concentrations, and a variety of artifacts and animal bone. All the choppers were recovered outside these features at Agate Basin (Frison 1982:Figure 2.16). As noted by Enloe and colleagues (1994:106), visual inspection of distribution maps is helpful in providing a general impression of broad spatial differences but the method is hampered by the sheer number of specimens represented in some categories, such as unmodified flakes for instance. Density contour maps were produced using the Topo module of the Surfer mapping program (Keckler 1997) to define concentrations of flaking debris and bison bone. Specimen counts were grouped by one-meter excavation units to provide density data. The highest frequency recorded for a one-meter-square was assigned a value of 100% (determined separately for bone and for lithics) and contour intervals were calculated as 10% increments of this figure. The resulting contour maps depict a general trend toward complementary distributions in the peak densities of flaking debris (Figure 94) and bison bones (Figure 95) in most areas, except the kill where flakes are underrepresented. The majority of the flakes result from tool resharpening, late-stage manufacture and repair of Folsom points, and manufacture of at least two ultrathin bifaces. Flaking debris is most concentrated in the central and southern portions of the camp and in the southwestern work area. Resharpening flakes were sufficiently sparse in the kill/initial butchery area that they do not appear at the lowest contour interval often flakes per meter-square. Bison skeletal material is more widely distributed with two concentrations in the kill/initial butchery area (lower right), another in the southwestern work area, and a particularly large concentration in the central camp that exhibits dual peaks. Significant bone clusters lie outside the block sample in the far

282

I

82-

I

I

I

*

+ ++ *

N

80-

M

78L

+ ++

75-

i +++

K

+

74-

+

. 4- AnviI

++

-

J

t

72-

+

+

+

+

+

++4

70-

4-$

H

68G 66-

--

+

4

++++ +

+

F

Bone cracking equipment

+~t

64-

* +

62-

+++ +-

+

Bison bone and fragments

+ ++, +

D +++ +++

60-

+

C

El

58B

t 56A AREA 2

,+++ 4+

+++ +

..

54-

+

Z AREA 1

*

+

++++ ++ ++ + + 4+ +4 +

4.+44+

52-

+ +

+

V

+4--

50-

++

+ 4.

4f

+

48-

w 46-

++

+ +++

:

4-

4i-

44

tIll]

T

+

R

+

36-

0

2826L 24-

++

+

0

30-

F I-

4.

.

4. +~

38-

32-

I-

.4-

+ 4.

+

S

P

-

+ +*

40-

34-

4.

4-

4+

.4-I-

+4.

42-

1-

+* +

+

+++4f ++ +

*

I-

4.

14.

+++ +

+ +

~

4.

+** +

+

F+

-

.44.

4-

+

+ :*.:*:+:

I +++ I

*

4-*~

1

t

A

.4-

-41

++ + + +

M

U I

Scale

=

2 meters

Unexcavated I

I

I

I

I

I

I

I

100 102 104 106 108 110 112 114 116 118 120 122 124 126 128 130 132 134 136 1

Figure 93. Distribution of bison bone and bone cracking equipment.

283

10

0

9692

0

Ii

00

88

I]

84 8

7 72

0

68 64 6

10% 20% 30% () 40% 50% 60% 70% 80% 90% 100%

10 flakes 39 flakes 58 flakes 78 flakes 97 flakes 117 flakes 136 flakes 156 flakes 175 flakes 195 flakes

Ii

Peak in Flake Concentration (1981 1/4" screen data)

100 104 108 112 116 120 124 128 132 136 140 144 148 152 156 160 164 168 172 176 180

Figure 94. Density distribution of flaking debris. Contours intervals represent 10% increments of the highest density one-meter square.

284

100

0

0

U]

92

o

8

El

84 10%

8 7 72

10% 3 bones 20% 6 bones 30% 9 bones D 40% 12 bones 50% 15 bones 60% 18 bones 70% 21 bones 80% 24 bones 90% 27 bones 100% 30 bones

6

c

64 6 50g.

0 10%

52 0 0

4 0

44

0

401. 36

/

100 104 108 112 116 120 124 128 132 136 140 144 148 152 156 160 164 168 172 176 180

Figure 95. Density distribution of bison bone. Contours intervals represent 10% increments of the highest density one-meter square.

285

north and northeast portions of the site (Figure 95).

Bone weathering patterns and lithic raw material

representation suggest that all these loci comprise portions of a single, large camp. The bone concentration in the NE excavation block bounded by E160 to 180 and N76 to 86 may represent discard near a second location of initial butchery. Flake tools, resharpening flakes and lithic debris from point production were recovered east of the bone in this block (E170 to 180). Preliminary data suggest that the remains in the NE block excavation are similar to those recovered at the juncture of the camp and kill/initial butchering areas located in the southeast portion of the main excavation block. Thirty-eight meters separates the two apparent loci of initial butchery. Bison bone lies buried across the intervening space as evidenced by fragments eroding at the surface. Density contours indicate that the bone concentration extends toward the east/northeast from circa E132 to 134 and N44 to 48.

K-means Cluster Analysis K-means nonhierarchical clustering (Kintigh and Ammerman (1982) was selected as an additional pattern recognition tool in this study on the basis of its demonstrated ability to accurately reconstruct the known locations of household and peripheral activity areas in a sample of !Kung sites (Gregg et al. 1991; Yellen 1979b). Even when distribution patterns at these sites were artificially "disturbed’ by means of computer simulation to approximate archeological situations, K-means methods continued to identify behaviorally meaningful patterns (Gregg et al. 1991:149). A particular strength of this nonhierarchical clustering method seems to be its ability to delineate activity areas by identifying separate clusters of items within a data set (Blankholm 1991:61). This is the objective here. The K-means algorithim is designed to minimize within-cluster variation, which simultaneously maximizes the variation among clusters. Gregg, Kintigh and Whallon summarize the program application in this way, The sum-squared-error (SSE) statistic is the measure of within-cluster variation that the analysis attempts to minimize, and this statistic is used in interpreting the K-means results... As with any clustering routine, there is always a question of what defines a "significant" cluster that warrants further investigation. We used two methods for identifying the significant clustering levels, and both entailed using the SSE statistic. The easiest method is to plot the log 10 %SSE against the clustering stage and to examine the resulting curve for negative inflection points. These points tend to indicate stages that produce more "economical" clustering, and these clusters should be further examined.

286

An independent method, which we also used, entails comparing the difference between the log 10 %SSE statistic from the input data and the log 10 %SSE statistics from the analysis of randomized version of the input data. The plots of the log 10 %SSE from both the normal run and the random runs are plotted against clustering stage. If the normal data plots within the range of the randomized data, then no significant clustering is indicated. Conversely, significant clustering is indicated where there is strong divergence between the real input and the random data. (Gregg etal. 1991:153, emphasis

added). (Gregg et al. 1991:153) The application of K-means clustering to CG data proceeded as recommended. The X, Y coordinates of 1,761 bison bones and 696 lithic specimens recovered in the block sample were processed using a quantitative software program developed by Kintigh (1992). Ralph Chapman, Biometrics Laboratory, National Museum of Natural History assisted in these computations. Two random runs were specified and the average log percent SSE statistic was plotted against number of clusters. The resulting graphs comparing random and CG data sets are seen for faunal specimens in Figure 96 (lower graph) and for lithic artifacts in Figure 97 (lower graph). Both graphs indicate that CG spatial patterning significantly diverges from expectations based on random distributions. That is, significant clustering exists in the CG data. In the next step, the difference in log percent SSE was plotted against clustering stage for CG bones and lithics. Figures 96 and 97 show these results in the upper graphs. The strongest inflection points in both graphs indicate that the most dramatic variance occurs at a six-cluster solution. While minor inflection points are evident at higher cluster intervals, a decision was made to investigate patterning in this dissertation at the strongest departure point (six clusters). Figures 98 and 99 depict CG specimens as assigned by the K-means program to one of six clusters of bone and lithics, respectively.

K-means Bone Clusters To explore the meaning of these data K-means bone clusters are overlaid with a scatterplot of bone in Figure 100. The bone concentrations in the kill/initial processing area are assigned to clusters K-2 and K-6. All faunal specimens in the southwestern work area cluster in K-3. Scattered bones in the south residential camp are combined in cluster K-i, while skeletal elements recovered in the northern end of the block sample are assigned to cluster K-S. Between K-i and K-S is cluster K-4, which contains bones associated with several hammerstone/anvils used in bone cracking. These expedient tools were abandoned

287

0.40 rI)

ci)

0.35 0.30 0.25

-

0.20 -i 0.15 0.10 – 0.05 + 0.00

1 2 3 4 5 6 7 8 8 9 10 1112 13 14 15 Number of Clusters

Bones 2.00 1.60 ci) (A

1.20 0.80

0.40 0.00

1 2 3 4 5 6 7 8 8 9 10 1112 13 14 15 Number of Clusters Bones

---

Random 1

Random 2

Figure 96. K-Means bone cluster graphs: lop) difference in log 10 % SSE; Bottom) log %SSE.

288

0.35 ci)

0.30

+

0.25 0.20

-

0.15 0.10

+

0.05 0.00

i;illt

III

1 2 3 4 5 6 7 8 8 9 10 1112 13 14 15

Number of Clusters

Lithics 2.00 1.60 JD

1.20 0.80 [IKI 0.00

1 2 3 4 56 78 8 9 10 1112 13 14 15

Number of Clusters Uthics

--- - Random 1

Random 2

Figure 97. K-Means lithic cluster graphs: lop) difference in 1og 10 % SSE; Bottom) log %SSE.

82 80 76 74 72 70 68 66 64 62 60 58 56 I 54 I52 0 z 50 48 46 44 42 40 38 36 34 32 30 28 26

E-27

100

104 108

112 116

120

124

128 132 136

East 6 Cluster configuration bones

Figure 98. Bone specimens assigned to six K-means clusters.

140

290

2 LU rII-

CLU Lo C0 4

LL-

44 )

Iq

4

LO M

LU

c\J 1.00

110

120 130 140 East 6 Cluster Configuration Lithics Figure 99. Lithic specimens assigned to six k-means clusters.

291

828078-

1

I

I

I

I

I

I

I

1

I

.

.

K-5.

0

68

60e 5856 54L 52504846 44

Distribution of K-Means Bone Clusters and Piece-Plotted Bone

..

3.

. .

t

.

.

.

AREA AREA I

.

-

.. ..

)

.

...

..,

.

.

.

0 38

3230-

LIII

tN Scale2 meters

Unexcavated 26 I I I I I I 24100 104 108 112 116 120 124 128 132 136 I

I

I

I

I

I

I

Figure 100. Distribution of k-means bone clusters and piece plotted bone.

I

I

when people left CG. The sequencing of refits to these choppers suggests that the bone-cracking activities in K-4 were among the last undertaken at the site. The bone in cluster K-4 was apparently discarded in work areas associated with exterior fires (suggested by the localized recovery of small burned flakes and fragments of calcined bone, see Figure 103), and possibly structures (circular areas devoid of bison bone). This is conjectural, but the idea warrants further investigation. Bone cluster K-4 is linked with both K-i and K-5 by hammerstone/anvil refits (Figure 101). A tool recovered in K-4 fits several flakes dislodged during previous use in K-S. This tool (GI 10-A2-4; Conjoin 18) is shown in the upper photograph of Figure 106 (lower row, far right), the flakes are seen in the lower photograph (upper row, center) of the same illustration. Another hammerstone/anvil recovered in K-4 refits fragments and flakes from K-I. This tool (G112-A2-8; Conjoin 15) is shown in the upper photograph of Figure 106 (lower row, middle tool). Refitted flakes indicate that a third hammerstone/anvil was used as bone-cracking equipment in K-i before its use and discard in K-4. This chopper (Hl 12-A2-6; Conjoin 16) appears in the upper photograph of Figure 106 (lower row, far left). A final group of refitted flakes (Conjoin 17) that evidently were detached from a fourth hammerstone/anvil (that was not recovered) links clusters K-i and K-4. These data strongly suggest that K-4 was among the final activity areas where bones were processed for marrow. This work may have been done cooperatively by individuals residing in the separate household areas where these tools were previously used. Bone clusters K-i, K-4, and K-5 seem to have been deposited during a single occupation. Table 41 compares skeletal element representation by cluster.

Table 41. Comnarison of Skelef2l Senient

Number of Specimens Impact Green fracture Articulations MNI Unidentified Upper forelimb Lower fore limb Upper hind limb Lower hind limb Phalange/Hoof

K-means I Jodr/ #1, 2 & 5 189 14 47 4 4 17 31 ii 18 23 7

7.4 24.9 16.0 13.3 9.0 16.4 5.8 9.5 12.2 3.7

1Anrecenffinn in TC-mpin fnn

K-means 2 Jodry #8 425

K-means 3 Jodry #7 182

4 21 7 5 124 35 18 14 45 14

8 24 3 4 46 11 2 7 22 6

0.9 4.9 28.0 16.7 29.2 8.2 4.2 3.3 10.6 3.3

4.4 13.2 12.0 13.3 25.2 6.0 1.1 3.8 12.1 3.3

flh11QtPr

K-means 4 Jodry #4 486 13 68 3 8 122 38 30 10 54 24

2.7 14.0 12.0 26.7 25.0 7.8 6.2 2.0 11.1 4.9

K-means 6 Jodry #9 280

1 8 9 53 2 18 20 54 30

0.3 32.0 30.0 18.9 0.7 6.4 7.1 19.3 10.7

293

I

82N 80M 78L 76K 747270H 68G 66F 64E 62D 60C 58B 56A 54z 52Y 50x 48w 46V 44U 42T 40S 38R 36Q 34P 32 0 30N 28M 26-

I

I

L

K-S

K-4 CY

*0 1

Anvil

i Impacted bone

J

Green fractured bone

t North of dotted line bone analysis is incomplete

1

*

K-i

Bone cracking tools and large fragments K-Means Bone Clusters

0 00

K-1 0

0

K-6

0

0

K-2 00

0 0 0

I

LJ*

D

0

9

0K-3

I CN Scale =2 meters

L

I

100

I

104

1

108

112

116

II

120

124

Ti - 1 - Th I

128

132

136

Figure 101. Distribution of refitted hammerstone/anvils, impacted and green fractured bone relative to k-means bone clusters.

294

Table 41. Continued. K-means 1 K-means 2 K-means 3 K-means 4 J o dry a #1, 2 & 5 Jodry #8 Jodry #7 Jodry 44 Axial skeleton 20 10.6 16.2 69 31 17.0 83 17.1 Ribs 14 7.4 49 11.5 29 15.9 72 14.8 Mandible 4 2.1 10 2.3 2.3 7 3.8 14 Crania 3 1.6 6 1.4 2 1.1 6 1.2 Long bones plinter 27 14.3 60 14.1 31 17.0 51 11.1 Notes: Bison bone in K-means cluster 5 is incompletely analyzed and thus not included here.

K-means 6 Jodry #9 22 11.6 4 1.4 11 3.9 12 4.2 32 11.4

’Density clusters 1 to 5 are those identified in Jodry 1987:Figure 4.18 and Jodry and Stanford 1992:Figure 4.29. Clusters 6 to 9 are a continuation of this numbering sequence in this study for clusters similarly identified by density contour methods (Figure 103). b515 bones/fragments recovered, 486 included ; 29 unanalyzed.

The bone clusters in the kill/initial butchery area differ (K-2 and K-6). Twenty-five of 26 impacted and/or green fractured bones from this part of the site were recovered in K-2 (Figure 101), as were a substantially greater number of ribs. Portions of at least nine bison are represented by left astragali in K-6, and portions of seven bison can be accounted for by right astragali in K-2. A total of fourteen bison and fifteen projectile points and fragments (MNI=14) are represented in an excavated portion of the kill/initial butchery area encompassing approximately 200 square meters.

K-means Lithic Clusters The relative distributions of K-means lithic clusters and a scatter plot of stone tools and fragments recovered in the block sample are illustrated in Figure 102. Relatively few artifacts were recovered in the CG kill/initial butchering area. This is consistent with kill site assemblages recovered from Folsom and three kills at Cooper. Figure 103 superimposes six K-means lithic clusters with ten bone/artifact concentrations determined using a density contour method. The latter include five density clusters (#1-5) reported from the 1981-83 excavations and four additional clusters (#7 to 10) identified in this study for areas excavated after 1983. K-means clustering data appear to support and expand previous findings (Jodry 1987, 1992; Jodry and Stanford 1992). Jodry (1987: 203-216) inferred five hearth-centered activity areas from concentrations of flaking debris, bison remains, burned lithics and bone, abraded pieces of red pigment, remains of weaponry repair and replacement, and concentrations of flake tools and broken edge fragments. These five concentrations are regularly spaced and separated by nearly 4.5 meters from center point of one to center of the next (Jodry and Stanford 1992:Figure 4.29).

295

82-

1

N-

I

1

I

I

I

80-

I

+

M-

78

I

I

+ +

-

L-

76-

+

K-

74I

-

+

~+4t + ~

4-++ 4+ 4t-i + f4-14-+4-

H-

68+

66-

+

F

4-

4-

4-

64-

~

~+

~+

+

-

60-

++ ~ + +

62D

~ 4-

C-

+

[J

58-

B-

56-

+ + +

A-LREA2

54z52-

4

REA I + +

x-

+

+

+

~

44_I-

+

I I

.

I

-s~

4-M-

+

4-

Distribution of Six K-Means Clusters and Lithic Artifacts (Piece-plot and Screen Recovery)

+ + +

~ 1-

4-

+

+

(IIIi) 44-

I

1-

~

+

~ +

~

+ 4+

++ +4-44-

1-

W

4-

+4-

~

~

4-

48-

+

+

K-6 4-

+

50-

+

+

+ +

$4-++ -+ + + + + + -#4+

+4-

+

+

~

.I~

+

-

1-

+

4-

+ .-- 4-

4-

14-

4-

4-

4-

4-

+ +

+

I- 4-

+

+

+

4-

+

46-

+

+

1(54-

+

++

+

70 -

++ ~ ~ ~ + 4-

4-41+

+

72-

I

4-

4-

44-

El

T-

S

~

~+

+

+

38-

+

36-

4-

Q-

+~

+

+

4-

40-

4-

K-4

7 ,

U-

42-

+

+

+

4-

4+ +4-

+

+

4-

1~

32

+ (*+

0

I

+ –#~i ++

tN Scale = 2 meters

3028-

Unexcavated

26-

L-

24100

i

I

104

I

I

108

112

116

I

120

124

128

132

136

Figure 102. Distribution of stone tools and fragments relative to k-means lithic clusters.

296

82

NI

80H 78-1 76-

KI

7472 70

KIM

IMW

68 66 64-i

K-Means Lithic Clusters

El

62-i Di

Density Contour Clusters Bone and/or Lithics at 40% in One-meter Square (Jodry and Stanford 1992 and this Study)

60 I

58 56 541

FMI

AREA 2 AREA I

52H

x

o.:

x

(9

Burned lithics 3to9 per m2

5 48H vq

46-i 44-I

r

b H

K-4

NI

42-a 40-

E

KTI

.

38 36H 34-H

K-i

32H DI

tN Scale = 2 meters

30H NI

28-1

Unexcavated

100

104 108

112

116

120

124

128

132

136

Figure 103. Distribution of K-means lithic clusters and clusters of bone and/or lithic materials identified using density contour methods.

297

K-means lithic cluster K-i comprises the southwestern work area, which contains bimodal peaks in both bone and flaking debris (#7; Figure 103). Figure 104 illustrates most of the tools from this area. The upper photograph shows end scrapers (a-g, k-I, n-p), a side scraper (in), and other flake tools (h-j) with use wear consistent with hide working (Ahier 1997; Kay 1996, 1998). The lower photograph includes gravers (a-b), discarded Folsom points (c-e), flake tools (h-j, 1-rn), a thick bifacial core (k), two ultrathin knives (f, n), a utilized cobble (o), and small tool edges apparently broken during use or resharpening. It is postulated that the northern and central part of K-i (N28 to 36) consists of an outdoor area where hide working and other maintenance tasks were carried out cooperatively by women (discussed further later in this chapter). At the southernmost end of K-I (and clearly extending into non-excavated terrain in this direction) channel flakes from two preforms were recovered suggesting that men were working on weaponry replacement in this area. Portions of three finished points were discarded in K- 1. Lithic cluster K-2 combines a bone dump (density cluster #10) and a concentration of bone, tools and flaking debris (density cluster #4). Artifacts from density cluster #4 are illustrated in the upper photograph of Figure 106. Included are three harnmerstone/anvils, two flake knives, a graver and other flake tools (b), two preform tips and several channel flakes representing at least four newly-made points (a, left), and a reworked base, two base fragments, and a complete but impact-damaged weapon tip representing four discarded points (a, right). Nine burned flakes and fifteen small pieces of calcined bone provide strong indications that these artifacts were discarded near a former hearth situated in the vicinity of one-meter square 01 14-A2 NW (Figure 103). Hide working tools are underrepresented in this area. Lithic cluster K-3 combines what appears to be an outdoor activity area near a hearth (density cluster #5) with a more spatially restricted concentration of tools and bone that may be associated with a structure (density cluster 91). The former contains chipping debris from Chuska, Cumbres, and Alibates (south of the site) as well as Black Forest and Trout Creek (north of the site). This represents the greatest variety of material types used in projectile point production recovered from a single artifact cluster (Figure 107, a). More than one flintknapper seems to have been working here. Also found were end scrapers, unifacial knives, and other flake tools. Half of an ultrathin biface that exploded from overexposure to heat

298

Figure 104. Tools from southwestern work area (K-means cluster K-i; density cluster 7; K=means bone cluster K-3). Upper: Tools with hide working use wear. Bottom: Gravers (a), Folsom points (c-e), flake tools (h-j, 1-m), bifacial core (k), ultrathins (f, n), utilized cobble (o); and tool edge fragments (g).

299

Figure 105. Folsom projectile points: a) Cumbres tip from kill/initial butchering area; b and c) Trout Creek points from southwestern work area.

300

Figure 106. Tools from residential camp (K-means lithic clusters K-2 and K-5). Top: K-means lithic cluster 2 (density cluster #4; K-means bone cluster K-4): Bottom: K-means lithic cluster K-5 (K-means bone cluster K-5).

301

Figure 107. Tools from residential camp (K-means lithic cluster K-3; K-means bonecluster K-I). Top: North end (density cluster #5; B: south end (density cluster 1).

302

was recovered and, in combination with other burned items, strongly indicates the former presence of a hearth. The bone in this area (density cluster #5) is a selected group of upper fore limb elements distributed on the periphery of the artifact concentration and inferred hearth. These may have been tossed around the edges of an outdoor seating area such as modeled by Binford (1983:153) from Nunamiut hunting camps. The bone associated with density cluster #1 is tightly clustered and another select group of high meat and marrow yielding elements is represented. These include an impacted and cut femur, green fractured humeri and proximal metacarpals, ribs, and an articulated lumbar segment from an immature bison. Within less than two meters are debris from late-stage manufacture of five Folsom points (Figure 107, lower photograph, a), as well as several flake tools and small broken tool edges (b) and fragments from a hammerstone/anvil (c). A postulated structure with a diameter of approximately two and one-half meters would accommodate the distribution of material in density cluster #1. The projectile points and single flake tool from lithic cluster K-4 (the kill/initial butchery area) have been described. The next lithic cluster is K-S. Of note here was the recovery of a platter-like anvil made of cemented sandstone, a local raw material represented by three hand-held hammerstone/anvils. The anvil is the largest tool recovered at the site. It weighs 2007.8 grains and is 196.5 mm long, 193.9 mm wide, and 44.4 mm thick. Portions of at least eleven tabular hammerstone/anvils and three battered cobbles appear to have been used in bone cracking. They have a mean weight of 700 grams (range 66.7 to 1113.2 grams). The tabular tools exhibit peck marks on one or both flat faces resulting from hammerstone-againstanvil contact during use. The author produced nearly identical damage during bison bone cracking experiments using similar equipment. Discarded near the anvil were large pieces of a broken hammerstone/anvil that fractured along cleavage planes in the raw material (Figure 106, lower photograph, upper left). A second hammerstone/anvil (Figure 106, upper photograph, lower right) recovered eleven meters away in K-2 refits several flakes recovered near the anvil and may have served as a replacement for the broken hammerstone. Figure 106 (lower photograph) illustrates other artifacts recovered near the large anvil including a sandstone abrader and pieces of red pigment (d), two complete preforms (a), a flake knife, an endscraper

303

and other flake tools (b). This suite of artifacts is strikingly similar to that uncovered in density cluster #2 (lithic cluster K-6) (Figure 108) and suggests a comparable set of activities were conducted in both areas. Artifacts from density cluster #2 include: channel flakes and a preform tip representing at least six newly fluted points (a), a basin-shaped sandstone abrader with residues of red pigment, pieces of pigment (d), a flintknapping hammerstone (d, right), gravers, end scrapers and other flake tools (b), and two battered gneiss cobbles. No finished points were discarded here. Raw materials that originate north of the site (Black Forest, Trout Creek, and Hornfels) dominate the flaking debris. The artifacts from K-5 (Figure 106, bottom) and density cluster #2 of K-6 (Figure 108) appear to represent materials discarded in the household areas of two different families during their brief stay at CG. Table 42 compares the proportional representation of selected tool classes and material types in Kmeans lithic clusters. Bifaces comprise between 10% and 25% of all clusters except K-4, which includes the kill/initial butchery and surrounding area. Here bifaces represent 51% of 27 artifacts (Figure 109, upper graph). If bifaces directly associated with the two bone clusters are considered separately (excluding scattered artifacts recovered to the west) this percentage is greater than 80%. Male-specific, weaponryrelated artifacts are best represented in lithic clusters K-3 (47% of cluster) and K-4

(55.6 % of cluster) and

least represented in K-i (7.8% of cluster). In contrast, hide working tools and ultrathin bifaces are prominent among the remains of K-i where they contribute 34.3% to the assemblage. Elsewhere they contribute between 2.4% to 9.5% (Figure 109, middle graph). If hide working and thin-meat cutting were tasks done primarily by females, then the concentration of end scrapers with soft material wear and ultrathins with light butchering wear in lithic cluster K-i is consistent with a female work area. The greatest variety of raw materials among end scrapers recovered from a single lithic cluster is found in K-I. Here end scrapers are made of six different tooistones as opposed to two or three elsewhere. This may suggest that women from separate households worked cooperatively in this location. In like fashion, the greatest variety of raw materials among channel flakes were uncovered in density cluster #5 of lithic cluster K-3. Five toolstone types are represented compared with two or three in other clusters. A group of men may have been flintknapping together in this area.

304

Figure 108. Tools from residential camp (K-means lithic cluster 6, density cluster 92).

305

0) U)

0 a)

CL

K-I

K-2

K-3

K-4

K-5

K-6

K-5

K-6

DBifaces DUnifaces

C,

4-.

U)

0 I-

0)

K-I

K-2

K-3

K-4

11 Points, channel fk, preform 0 End scraper, Ultrathin

I-

0)

4-.

U)

0 0) CL

Figure 109. Comparisons of bifaces versus unifaces, weaponry versus end scrapers and ultrathins, and hammerstone /anvils versus utilized cobbles in K-means lithic clusters.

I.-

a) U)

C.) I-

a)

K-I

K-2

K-3

K-4

K-5

K-6

K-Means Lithic Cluster

DAlibates REdwardsChuska DCumbres

I-.

a) U)

C-) I.-

0.

K-I

K-2

K-3

K-4

K-5

I

IM

N

K-6

K-Means Lithic Cluster Black Forest 0 Trout Creek

80 a) ’l-

U)

o 40

iuI__

I-

0)

0. 20

n

ri K-I

K-2

II I.!ii iiiiii :RJii 1 K-3

K-4

K-5

K-6

K-Means Lithic Cluster [D Southern Materials 0 Northern Materials

Figure 110. Comparisons of lithic raw material types in K-means lithic clusters.

Table 42. Comparison of Selected Tool Classes and Material Types in K-Means Lithic Clusters Class K-I % K-2 K-3 % All Bifaces 18 10.8 18 14.6 29 21.0 All Unifaces 121 72.9 68 55.2 60 44.0 Small Uniface Fragment 43 25.9 26.8 33 20 15.0 Point 7 4.2 13 10.5 12 8.8 Preform 1 0.6 3.2 10 4 7.3 Channel Flake 5 3.0 16 13.0 43 31.0 Ultrathin 5 3.0 1 0.8 6 4.4 Flake Knife 3 1.8 2 1.6 2 1.5 End scraper 26 15.7 2 1.6 7 5.1 Graver 8 4.8 4.0 1 0.7 5 Radial Fragment 4 2.4 1 0.8 Abrader 1 0.6 1 0.8 Utilized Cobble 1 0.7 0.6 1 0.8 1 Hammerstone/anvil 1 4.0 0.6 5 Pigment 16 9.6 11 8.9 Flake modified by use only 29 15.4 22 16.0 17.5 19 Core 1 0.6 0 166 137 Sample Size 123 Material Type 4 9 6.6 8 4.8 5 Morrison Quartzite 4 2.9 1 0.6 Alibates 1 0.7 1 0.8 Edwards 13 9.5 2 1.2 6.5 8 Chuska 17 12.0 13.8 10 6.0 17 Cumbres 25.1 31 12 26 21.1 7.8 South of CG 46.0 63 101 47 38.2 Black Forest 60.8 12.0 16 20 1.2 11 8.9 Trout Creek 2 1.2 Hornfels 79 58.0 123 74.1 47.1 58 North of CG

K-4 14 8 3 14

51.9 29.6 11.1 51.9

1

3.7

1 1

3.7 3.7

3

%

11.1

K-5 17 30 8 13 3 6 1 5 1 1 1 7 14 14

% 24.3 42.9 11.4 18.6 4.2 8.5 0 1.4 7.1 1.4 1.4 1.4 0 10.0 20.0 20.0

70

27

K-6 14 78 20 4 6 52 1 5 13 10

8.2 46.0 12.0 2.4 3.5 31.0 0.6 2.9 7.6 5.9

1

0.6

5 14 27 1 170

2.9 8.2 16.0 0.6

%

2

7.4

2

2.8

7

4.1

4

14.8

6 10 7 2

22.2 37 25.9 7.4

9

33.3

1 4 11 16 19 8 1 28

1.4 5.7 15.7 22.8 27.1 11.4 1.4 40.0

8 12 20 83 26 1 110

4.7 7.1 12.0 49.0 15.0 0.6 65

308

Debris from the use and refurbishing of bone-cracking equipment is represented in all lithic clusters, but the tools themselves dominate the assemblage from K-S (Figure 109, lower graph). Marrow recovery was among the last activities that took place in density cluster #4, and it may have been one of the final processing tasks undertaken prior to site abandonment. Etimoarcheological data from Efe campsites indicate that material items lying beside fires and elsewhere

...

had been deposited during the final days or

hours that the camp was occupied, and thus portray with fine-grained resolution activities that were carried out at the end of the campsite’s lifespan’ (Fisher and Strickland 1991:223). Females commonly undertake the processing of bones for marrow (as opposed to cracking bones for a snack or a meal). It is postulated that the concentration of hammerstone/anvils and impacted long bones in density cluster #4 represents the remains of a cooperative female task involving individuals from several households, perhaps including those represented in lithic clusters K-3 and K-6. Figure 110 compares raw material types in K-means lithic clusters. It is interesting that toolstones from southern source areas (i.e., Alibates, Edwards, Chuska, and Cumbres) and those from the north (i.e., Black Forest, Trout Creek) are represented in nearly equal amounts in the lithic assemblage in K-4, which includes the kill/initial butchery area (lower graph). Northern material types dominate elsewhere and are most strongly represented in K-I (74%) and K-6

(65%). In both these clusters there are concentrations of

hide working tools. Southern materials contribute 26%, 25%, and 23%, respectively, to the assemblages in K-2, K-3, and K-5, where bifaces are comparably represented as well (Figure 109, upper graph). Some Cumbres chert and Black Forest silicified wood entered the site in a newly procured state relative to Trout Creek, Chuska, Alibates, and Edwards. The former are represented in K-4 in similar percentages (Black Forest, 26%; Cumbres, 22%), but in the postulated female work area in K-1 Black Forest represents 60% and Cumbres only 6%. Likewise in K-6, Black Forest contributes 83% and Cumbres 12%. Expedient tools and hide working implements are most often made of Black Forest, the most abundant raw material at the site. If people arrived at CG from two different directions as the more recently acquired raw materials might suggest, it appears that these materials were redistributed through exchange and activities conducted near multiple hearths. Materials from the south are particularly well represented in lithic cluster K-3

309

where 80% of the Alibates, 35% of the Chuska, and 14% of the Edwards tools were recovered, in addition to 20% of the Black Forest and 23% of the Cumbres. The Alibates, Chuska, and Edwards in this area represent late-stage Folsom point production. A single episode of site use appears to be represented by the six K-means clusters. Although relative proportions vary from cluster to cluster, the same suite of raw material types is found across the site (Figure 111). Refit data provide an independent line of evidence. Refitted fragments of finished projectile points and hammerstone/anvils are the most common artifact types linking clusters of discarded bone and cultural material. They connect all K-means lithic clusters except K-I (Figure 112). By comparison, resharpening flakes or edge fragments conjoined with flake tools generally are restricted to a single cluster, as exemplified by refits in K-i. One notable exception is a refurbished end scraper from K-6 (Figure 68, i) that conjoins a fragment from its former working edge recovered in lithic cluster K-2. The next section considers the distribution of activities as revealed by preliminary use-wear data.

Preliminary Use-Wear Results A total sample of 394 unifacial tools and fragments were examined microscopically by the author under magnifications ranging from 6,4 to 40X following low-magnification procedures outlined by Ahler (1975, 1979) and using a Leica Wild M3C microscope and high-intensity reflective light. Use-wear specialists (Stan Ahler, Marvin Kay, and John Tornenchuk) studied 160 of these artifacts including Folsom points, ultrathin bifaces, a bifacial core, wedges, gravers, end and side scrapers, flake knives, non-retouched flakes, and a few channel flakes and refitted resharpening flakes. Table 43 enumerates the tools analyzed by each specialist. The respective distributions of these samples are shown in Figure 113.

Table 43. Sample of Unifacial and Bifacial Artifacts Studied by Use-Wear Specialists. Ahler Kay Tomenchuk Total Number of artifacts examined Percent of use-wear sample Percent of total lithic tool sample Number of artifacts examined exclusively by one analyst Number of artifacts examined jointly (Ahler and Kay) Number of artifacts examined jointly (Ahier and Tomenchuk) Total number of specimensanalyzed by use-wear specialists

70 43.7 7.3 20 49 1

96 60 10 47 49

44 27.5 4.6 43

-

1

-

-

-

-

210 -

-

110 49 160

310

82N

1

I

1

I

I

I

I

I

80M 78-

I

I

1

A E

I

+

+w

7674

– ++* *+

)E

72I

L)o +

6 6664-

++

+

D

60C

+ Black Forest *Cumbres

*E44-A+

+ri

+

* Hornfels

*

D M. Quartzite * Edwards

+ > A>M

t AREA 1

z AREA I

52 50x 48-

W +

A * + +E A +–+

A+*

+

*

> Trout Creek *

++ +++ + Et,+ *0 0

LI

*

r::’ ++

w

Q

A Chuska

+

+

*

B

46V 44U 42T 40S 38R 36-

A

A+[>4

58-

0 Alibates

+

+

62

A

+

+– +

F

56-



+9+

70-

– *

-I-

El

+

+

F=



*

+

+ * +* 1 *

+

*

+

I.

P

32-

I fN

28-

Scale = 2 meters

+–

M

26-

Unexcavated

] ++ tM

I.

24100 104

I

I

I

I

108 112

I

I

116

I

I

I

I

I

I1

120 124 128 132 11 136

Figure 111. Distribution of raw materials in the block sample.

311

IM /

78 76 74 72 70

ers

AF AF

5 5 4

S

3 3 2

I

I

I

I

I

100 104 108 112

I

I

116

I

I

I

I

II

I

-T-

120 124 128 132 136

Figure 112. Distribution of refits in the block sample in relation to K-means lithic clusters

312

11

Distribution ofFomenchuk Use-Wear Sample

LJ

[!.

Distribution of Abler Use-Wear Sample

Distribution of Kay Use-Wear Sample

Figure 113. Distribution of use-wear samples studied by Ahier, Kay, and Tomenchuk.

313

The preliminary results presented here summarize a pilot study of use-wear in the CG lithic assemblage. The aims of this initial work were to assess the presence/absence and range of use-wear traces, the extent of wind/sediment polish, and the kinds of behavioral information derived using different approaches. Ahier (low magnification), Kay (high magnification) and Tomenchuk (parametric or mathematical engineering techniques) employ different methods. Likewise their interpretive emphases vary somewhat with regard to the identification of haft and/or transport wear and the mechanical fracture properties of different toolstones. For this reason the author anticipated that their results would be complementary rather than redundant. In large measure this has been the case. Each use-wear specialist has published the fundamental analytical procedures he follows and the equipment he uses (Ahler 1979; 1994; Kay 1996b; Storck and Tomenchuk 1990; Tomenchuk 1997), and readers are referred to these firsthand accounts. All three researchers are in agreement that use-wear traces attributed to tool use and hafting are preserved in the CG assemblage despite some overprinting from wind and sand abrasion. Clear, unmistakable evidence of wind damage occurred on 13 tools and 14 non-tools (or 28% of the sample), and was plainly absent on 49 tools and 15 non-tools (or 66% of the sample) (Kay 1998:5). Each concurred with an earlier assessment that the chipped stone assemblage was buried fairly rapidly (Jodry 1987). This is evidenced by nearly pristine tool edges, which were not subjected to the extensive subaerial weathering exhibited by most of the faunal assemblage (owing to the greater transverse cross sections of most bones, which required longer to bury as discussed in Chapter 3). The authors block sampling strategy was guided by three goals: 1) to gain insight regarding Folsom activities, 2) to compare use-wear traces between flake knives and ultrathin bifaces and between end scrapers in different areas of the site, and 3) to distinguish, where possible, between hand-held and hafted tools. The latter was in keeping with Marvin Kay’s particular research interest in this assemblage. To enhance the potential of identifying the range of activities in different site areas, I opted for block samples of all (or nearly all) implements recovered in four activity areas. With reference to Figure 103 these are: density clusters #3 (K-6) and 95 (K-3) studied primarily by John Tomenchuk (1995, 1999);

314

density cluster #7 (K-i) studied by Marvin Kay and Stan Ahler; and the projectile points from the kill/initial butchery area (K-4) studied by Marvin Kay. Additionally, Abler and Kay studied most flake knives and all ultrathin bifaces recovered from other areas to compare the relative use of these different knife forms in butchering and other tasks. Finally, to determine whether end scraper use differed between the southwestern work area and other loci in the camp, Ahler examined additional end scrapers recovered outside the southwestern work area. An extended discussion of use-wear data will appear in joint authored articles with the use-wear specialists. With their permission, preliminary functional identifications are discussed in Chapter 4 and the distribution of this material is described here. Figures 114, 115 and 116 show scatter plots of tools with use wear traces attributed, respectively, to hard material wear, to butchering and/or meat cutting wear, and to abrasive or soft material wear that appears to result in large measure from hide working. Tools with hard material wear are distributed in all areas of the site (Figure 114). Tasks involving wood, bone and/or antler appear to have been undertaken frequently. Tools used in butchering and/or meat cutting were recovered in the residential camp and southwestern work area (Figure 115), although some stone knives may have been removed from the shallowly buried killlinitial butchery location by collectors. Hofman (1999:126-28) suggests that flake tools and reuseable points are less likely to be lost in kill areas when carcasses are relatively dispersed and initial butchering proceeds in a less constrained space. Table 44 follows Hofman in comparing carcass densities and tool frequencies in seven Folsom kill/initial butchery areas.

Table 44. Bison Carcass Density and Number of Tools Recovered in Seven Folsom Bison Kills. Folsom Site (Kill Areas) Carcass DensitL Flake Tools Complete Points Cooper (three kills) one bison per 6-12mm >12-18mm >18- 25 mm >25- 38mm

Completeness

1 2 3

Complete or nearly so Proximal Distal

Weight

Weight to nearest 0.1 gram using OHAUS GT4000 electronic scale.

Maximum Length

Maximum longitudinal dimension from the impact on the platform to the distal edge measured along the medial axis to the nearest 0.01 mm. (Wilmsen and Roberts 1974:67).

Maximum Width

Maximum lateral dimension measured perpendicular to the medial axis to the nearest 0.01 mm. (Wilmsen and Roberts 1974:67).

Maximum Thickness

Maximum transverse (dorsal to ventral) dimension measured below the bulb, measured to the nearest 0.01 mm. (Wilmsen and Roberts 1974:67).

Edge Angles

Angular dimensions between the distal and ventral surfaces measured at the edges of a specimen with a goniometer to the nearest 5 0 Wilmsen and Roberts 1974:67). A range of variation recorded per unifacial tool edge.

6 7 8 9

4 5 6

>38 - 50 mm >50-76mm >76 -100nun > 100 nun

Medial Indeterminate Lateral

. (

General Variables, All Tools. Acronym Variable Name

Code

Attribute State

MCII

1

bipointed biface (Ahler 1994:40)

4 5 6 15 21 21.1

ovoid biface ovoid, pointed biface rectangular biface Indeterminate biface fragment generalized end scraper form (unspurred) unspurred end scraper, lacking dorsal flaking from lateral margins

Morphological Class

349

Acronym

Variable Name

Code

Attribute State

CORT

Cortex

1

Present in any amount (Ahler 1994:4 1)

BLNK

Original Blank Form

0 1 3

7 9 11

indeterminate (differs from Abler 1994) flake, not further specified (differs from Ahler 1994) subrounded, rounded, sperical cobble or pebble (Abler 1994:37) other nonbipolar flake, with no platform present or with unprepared platform present bifacial thinning flake blade or bladelet indeterminate

6

STG

Stage

I 2 3 4 5 6

unaltered blank (Ahler 1994:40) edged biface initially thinned biface secondarily thinned biface preform ready for hafting or haft development hafted biface

PrGr

Product Group

1 2 3 4 5 6 7 8

acquisition of raw material (Collins 1975:25) core preparation and initial reduction primary trimming secondary thinning use maintenance specialized discard uncertain

REJ

Reason for Rejection

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15

has potential for further work or use (Abler 1994:35) bending fracture or end shock perverse fracture material flaw or poor stone quality outre-passe fracture (overshot) compound hinge-step occurrence impact fracture small size or exhaustion indeterminate heat or thermal fracture lateral break (snap fracture) radial fracture concentric chunk from tool margin channel flake or fragment other (differs from Abler 1994)

RMAT

Raw Material Type

FUNC

Functional Class

See Table A. XXXXXXXXXX 2 3 5 6 8 16

perforator, drill (Abler 1994:33) light duty bilateral cutting tool basal scraper/grinder light duty transverse scraper used on soft material expedient, general purpose cutting tool transverse scraper used on abrasive material

351

Acronym Variable Name FUNC Functional Class Cont.

MULT

Code 17 18 19 20 21 22 23 26 30 45 58 66 67 68 69 71

Attribute State transverse scraper used on hard material denticulated flake or edge modified tool slotting or grooving tool generalized transverse scraping tool core utilized flake used to saw or slice hard material retouched or utilized flake used on variable material punch/wedge/chisel graving or incising tool spokeshave notched flake flake ridge plane used on resistant material Snap break plane used on resistant material point-concentrated wear on radial break or pie-shaped tool hinge edge tool Wood working adz

2 3 4

single function tool (codes #s differ from Ahler 1994:42) two functions in same artifact three functions in same artifact four functions in same artifact

Multiple Functions

RESD

Residue present

CLASS

General Tool Class

Possible residue noted during analysis I 2 3 3.1 3.2 3.3 3.4 3.5 3.7 3.8 3.9 4 4.1 4.2 4.3 4.4 5 5.1 5.2 5.3 6 6.1 6.4 6.5 6.6 6.7

Manuport Core Unmodified flake uniface use/resharpening flake preform ribbon flake biface thinning fk channel flake double channel flake angular debris hammerstone/anvil flakes heat spall Flake tool use modified only use and deliberately shaped edge uncertain if 4.1 or 4.2 flake tool fragment, small piece of a larger tool Non-chipped stone tool abrader composite abrader/hammerstone flintknapping hammerstone Biface Folsom point stage 4 biface, preform ultrathin biface bifacial wedge bifacial core remnant

352

Acronym

Variable Name

TECHNO Technological Class

Code 7 7.1 7.2 7.3 8 8.1 8.2 9 9.1 9.2 9.3 10

Attribute State hammerstone/anvil (bone cracking equipment) hammerstone/anvi I anvil hammerstone (non-flintknapping) Pigment culturally modified unmodified Ground stone and Fire-cracked rock ground stone fire-cracked ground stone fire-cracked rock, unground specimen assigned lithic artifact number in field and deemed not to be a lithic artifact during analysis in lab.

2 3 4 5 6 7 8 9 11 12 13

patterned, medium to large, thin bifacial tool (Ahler 1994:49) unpatterned, small to medium crude bifacial tools patterbeded, steeply beveled flake tools unpatterned other flake tools unpatterned, thick, bifacial core tools unpatterned non bipolar cores and core tools unpatterned bipolar cores and tools unpatterned pecked, ground, or cobble tools radial break tools retouched tabular pieces ultrathin biface

Additional Variables for Flake Tools Code Attribute State

Acronym

Variable Name

DstT

Distal Termination

Scars

Dorsal Scars

Number of dorsal flake scars from previous removals. Platform reducing excluded. (Deller and Ellis 1992:147)

ScarD

Dorsal Scar Orientation

2 3 4 5 6 7

Direction of dorsal scars relative to flake’s longitudinal axis. (Deller and Ellis 1992:147-8) parallel-unidirectional parallel-bidirectional convergent-unidirectional convergent-bidirectional transverse flat uncertain

1 2 3

flat or diffuse (Deller and Ellis 1992:148) moderate, well-defined pronounced and projects markedly

Bulb

Bulb Type

1 2 3 4 5

feathered hinge or step snap fracture recent damage indeterminate

353

Acronym Ripi CurvD

LtEO

Variable Name Ripples

Code Attribute State 1 Undulations present 2 Smooth Placement of greatest 1 distal (Deller and Ellis 1992:148) longitudinal curvature 2 symmetrical, curvature is well-centered and gradual along the flakes length 3 proximal, curvature pronounced toward proximal end

Lateral Edge Orientation

1

parallel

2 3 4

contracting expanding indeterminate

PP

Platform Preparation

1 2 3 4

cortical unprepared Prepared (i.e. isolated/faceted/reduced/ground) indeterminate

CERT

Flake Tool Certainty

0

uncertain flake tool, possibly exhibiting post-depositional edge damage (Ahler 1994:40) certain, definite use or retouch uncertain tabular tool, possible modification through testing or edging

2

Additional Variables, Preforms and Folsom Points Max Length (L)

Maximum longitudinal dimension from the distal tip to a line constructed between the basal tangs of the point. Measured to the nearest 0.01 mm. (Wilmsen and Roberts 1984:102-3, Figure 99).

Basal Break

Length to nearest 0.01 imi -i, along longitudinal axis from distal fracture to proximal edge of preform.

Face A Length of Lateral Grinding Flute Length Flute Width

Length to nearest 0.01 mm Length to nearest 0.01 mm Average width to nearest 0.01 mm

Face B Length of Lateral Grinding Flute Length Flute Width

Length to nearest 0.01 mm Length to nearest 0.01 mm Average width to nearest 0.01

Max. Width (W)

Maximum width between any two points on the lateral edges measured orthogonally to length. This is the maximum dimension along a line at right angles to the axis of fluting. (Wilmsen and Roberts 1984:102-3, Figure 99).

354

Basal Width

The distance between the tips of the two basal tangs. Wilmsen and Roberts 1984:102-3, Figure 99).

Basal Concavity Depth

The maximum length of the basal concavity measured from a line drawn between the ends of the basal tangs to the deepest part of the concavity. (ProxirnalLength [P]; (Wilinsen and Roberts 1984:102-3, Figure 99),

Max. Thickness

The maximum distance between the dorsal and ventral surfaces beyond the bulbar area (Wilmsen 1970:19). If the bulb is flat or absent, the maximum thickness beyond the platform (Deller and Ellis 1992:147).

Flute Thickness

Maximum thickness to nearest 0.01 mm within flute scars taken at midline (Ts).

Ridge Thickness

Maximum thickness to nearest 0.01 mm., both margins at edge of flute scar.

Tip Thickness

Thickness at distal end of preform, measured 2.5 mm proximal to tip.

Channel Flake Scars

The length and average width of channel flake scars on faces a and b. (Wilmsen and Roberts 1984:102-103, Figure 99, [Sla and Slb]).

Retouch Density

Number of lateral retouch flakes per 1 cm, average of multiple measurements taken along each edge.

Number of Faces Fluted Number of Flutes Number of pseudo flutes

0, 1 or 2, (3 =indeterminate) 0, 1,2, 3, 4 or more 0, 1 or 2

Flute Scar A Interrupted by distal flaking

1 2

yes no

3 indeterminate

Flute Scar B Interrupted by distal flaking

1 2

yes no

3 indeterminate

Relative Timing of Preform Breakage

0 1 2 3 4 5 6 7 8

unbroken before fluting during 1st flute after flute/before 2nd flute 2nd during Flute after 2md Flute while shaping during 3rd flute after 3td flute while shaping uncertain

Point Fragmentation

See illustration, Figure 50.

(Proximal Width [Wp],

BIBLIOGRAPHY Adams, J. L. 1988 Use-Wear Analysis on Manos and Hide Processing Stones. Journal of Field Archaeology 15:307315. Agogino, G. A. and A. Parrish 1971 The Fowler-Parrish Site: A Folsom Campsite in Eastern Colorado. Plains Anthropologist 16:111114. Ahlbrandt, T. S., S. Andrews, and D. T. Gwynne 1978 Bioturbation in Eolian Deposits. Journal of Sedirnentaiy Petrology 48:839-848. Abler, S. A. 1975 Pattern and Variety in Extended coalescent Lithic Technology. Unpublished Ph.D. dissertation, University of Missouri, Columbia. University Microfilms, Ann Arbor. 1976 Lithic Resource Utilization Patterns in the Middle Missouri Subarea. In Trends in Middle Missouri Prehistory: A Festschrfl Honoring the Contributions of Donald J. Lehrner, edited by W. R. Wood, pp. 135-iSO. Memoir 13, Plains Anthropologist, Lincoln. 1979 Functional Analysis of Nonobsidian Chipped Stone Artifacts: Terms, Variables, and Quantification. In Lithic Use-Wear Analysis, edited by B. Hayden, pp. 301-328. Academic Press, New York. 1986

The Knife River Flint Quarries: Excavations at Site 32DU508. State Historical Society of North Dakota, Bismarck.

1997 Use-Wear and Functional Analysis of Selected Artifacts from Stewart’s Cattle Guard Site (5AL101), Colorado. Technical Report Submitted to Margaret A. Jodry, Department of Anthropology, PaleoindianlPaleoecology Program, Washington, D. C. Ahler, S. A. (editor) 1992 Phase II Cultural Resource Investigations Associated with ProposedDani Safety Modifications at Lake flo National Refuge, Dunn County, North Dakota (Final Report). University of North Dakota, Grand Forks and Northern Arizona, Flagstaff. Submitted to the U.S. Fish and Wildlife Service, Contract no 14-16-0010-09-0003, Denver. 1994 A Working Manualfor Field and Laboratory Techniques and Methods For The 1992-1996 Lake flo Archaeological Project. Prepared at Quaternary Studies Program, Northern Arizona University. Developed under terms of Modification ito Cooperative Agreement No. 14-48-001093-901 Between the U. S. Fish and Wildlife Service and the University of North Dakota, Grand Forks. Ms. on file, Paleocultural Resource Group, Flagstaff, Arizona. Ahier, S. A. and Phil R. Geib 1999 The Role of Fluting in Folsom Point Design and Culture. Paper Presented at the Folsom Conference II, Austin, Texas.

355

356

Abler, S. A. and M. A. Jodiy 1997 Scraper Use-Wear as an Indicator of Folsom Mobility in High-Altitude Southern Colorado. Paper Presented in a Symposium entitled "View from Above: High Altitude Archaeology and the Mountain Tradition", 55tl Annual Plains Anthropological Conference, November 19-22, 1997, Boulder. Akoshima, K. and G. C. Frison 1996 Lithic Microwear Studies of the Mill Iron Site Tools. In The Mill Iron Site, edited by G. C. Frison, pp. 71-86. University of New Mexico Press, Albuquerque. Alley, R., D. Meese, C. Shuman, A. Gow, K. Taylor, P. Grootes, J. White, M. Ram, E. Waddington, P. Mayewski, G. Zielinski 1993 Abrupt Increase in Greenland Snow Accumulation at the End of the Younger Dryas Event. Nature 362:527-529. Amick, D. S. 1994a Folsom Diet Breadth andLand Use in the American Southwest. Unpublished Ph.D. dissertation, Department of Anthropology, University of New Mexico, Albuquerque. 1994b Technological Organization and the Structure of Inference in Lithic Analysis: An Examination of Folsom Hunting Behavior in the American Southwest, In The Organization ofAmerican Prehistoric Chipped Stone Tool Technologies, edited by P. J. Can, pp. 9-34. International Monographs in Prehistory, Archaeological Series 7, Ann Arbor. 1996 Regional Patterns of Folsom Mobility and Land Use in the American Southwest. Archaeology 27:411-426.

World

1999 Raw Material Variation in Folsom Stone Tool Assemblages. In Folsom Lithic Technology, Explorations in Structure and Variation edited by D. S. Amick, pp. 169-187. International Monographs in Prehistory, Archaeological Series 12. Ann Arbor. Anderson, D. J. 1998 A Geochemical Analysis and Characterization Study of Narbona Pass Chert. Abstracts of the 63)1 Annual Meeting, Society for American Archaeology, Seattle, Washington, March 25-29, 1998.. Andrefsky, W. Jr. 1994 Raw Material Availability and the Organization of Technology.

American Antiquity 59:21-34.

Andrews, S. G. 1981 Sedimentology of Great Sand Dunes, Colorado. In Recent and Ancient Nonmarine Depositional Environments: Models for Exploration, edited by F. G. Ethridge and R. M. Flores, pp. 270-291. Society of Economic Paleontologists and Mineralogists, Special Publication 31:270-291. Armstrong, D. M. 1972 Distribution of Mammals in Colorado. Monograph of the Museum of Natural History, The University of Kansas, Number 3, Lawrence. Attla, C. 1971 How to Tan Your Moose Hide. Alaska Magazine, Volume XXXVII, Number 10, October 1971. Audouze, F. and J. G. Enloe 1997 High Resolution Archaeology at Verberie: Limits and Interpretations. In High Definition Archaeology: Threads Through the Past, edited by J. A. J. Gowlett, pp. 195-207. World Archaeology 29.

357

Bailey, H. 1960 A Method for Determining Warmth and Temperateness of Climate. GeograjIska Annaler 43:1-16. Bamforth, D. 1985 The Technological Organization of Paleoindian Small-Group Bison hunting on the Llano Estacado. Plains Anthropologist 30:243-258. Technological Efficiency and Tool Curation. American Antiquity 51:38-50. Ecology and Human Organization on the Great Plains. Plenum Press. New York.

Flintknapping Skill, Communal Hunting, and Paleoindian Projectile Point Typology. Plains Anthropologist 36:309-322.

Technological Organization and Hunter-Gatherer Land Use: A California Example.

American

Antiquity 56:216-234.

Banks. K. M. and T. A. Del Bene 1987 Record in Beauty: Prehistoric Resources in the Navajo Nation Forest, Defiance Plateau and Chuska Mountains, Arizona and New Mexico. South Dakota Archaeology 2:20-81. Banks, L. D. 1990 From Mountain Peaks to Alligator Stomachs: A Review of Lithic Sources in the Trans-Mississippi South, the Southern Plains, andAdjacent Southwest. Oklahoma Anthropological Society, Memoir

4. Norman. Banks, N. G. 1970 Nature and Origin of Early and Late Cherts in the Leadville Limestone, Colorado. Geological Society ofAmerica Bulletin 81:3033-3048. Barker, J. H. 1993 Always Getting Ready, Upterrlainarluta, Yup’ik Eskimo Subsistence in Southwest Alaska.

University of Washington Press, Seattle. Bartlein, P. J., M. E. Edwards, S. L. Shafer and E. D. Barker, Jr. 1995 Calibration of Radiocarbon Ages and the Interpretation of Paleoenvironmental Records. Quaternary Research 44:417-424. Basso, K. H. 1996 Wisdom Sits in Places, Landscape and Language among the Western Apache.

University of New

Mexico Press, Albuquerque. Baugh, T. G. 1991 Ecology and Exchange: The Dynamics of Plains-Pueblo Interaction. In Colonists: Interaction between the Southwest and the Southern Plains

Farmers, Hunters, and

edited by K. A. Spielmann,

pp. 89-106. The University of Arizona Press, Tucson. Beckerman, S. 1983 Carpe Diem: An Optimal Foraging Approach to Bari Fishing and Hunting. In Adaptive Responses of Native Amazonians, edited by R. Hames and W. Vickers, pp.269-99. Academic Press, New York. 1999 Personal Communication 1999, Department of Anthropology, Pennslyvania State University.

358

Behrensmeyer, A. K. 1978 Taphonomic and Ecologic Information from Bone Weathering. Paleobiology 4:150-162. 1992 Personal Communication, Department of Paleobiology, National Museum of Natural History, Smithsonian Institution, Washington DC. Belitz, L. 1999 Personal Communication. Professional Hide-Tanner, Hot Springs, South Dakota. Museum Consultant Regarding Plains Material Culture. Bement, L. C. 1999 Bison Hunting at Cooper Site, Where Lightning Bolts Drew Thundering Herds. Oklahoma Press, Norman. Benedict, J. B. 1973 Chronology of Cirque Glaciation, Colorado Front Range. 1981

University of

Quaternary Research 3:585-599.

The Fourth of July Valley. Glacial Geology and Archaeology of the Timberline Ecotone.

Research Report 2. Center for Mountain Archeology, Ward, Colorado. 1985 Arapaho Pass, Glacial Geology at the Crest of the Colorado Front Range. Research Report 3. Center for Mountain Archeology, Ward, Colorado 1992a Footprints in the Snow: High altitude Cultural Ecology of the Colorado Front Range, U.S.A. Arctic andAlpine Research 24:1-16. 1992b Along the Great Divide. Paleoindian Archaeology of the High Colorado Front Range. In Ice Age Hunters of the Rockies, D. J. Stanford and J.S. Day, editors, pp.343-359. University Press of Colorado, Boulder. 1999 Effects of Changing Climate on Game-Animal and Human Use of the Colorado High Country (U.S.A.) since 1000 BC. Arctic, Antarctic, andAlpine Research 31:1-15. Berg, R. T. and R. M. Butterfield 1976 New Concepts of Cattle Growth. John Wiley & Sons, New York. Berger, J. and M. Peacock 1988 Variability in Size-weight Relationships of Bison bison. Bettinger, R. 1987 Archaeological Approaches to Hunter-Gatherers.

Journal of Mammology 69:618-624.

Annual Review ofAnthropology

16:121-42.

1991 Hunter-Gatherers: Archaeological Method and Theory. Plenum Press, New York. Billeck, W. T. 1998 Fluted Points Distribution in the Loess Hills of Southwestern Iowa. Plains Anthropologist 43:401 409. Binford, L. R. 1976 Forty-seven Trips: A Case Study in the Character of Some Formation Processes of the Archaeological Record. In Contributions to Anthropology: The Interior Peoples of Northern Alaska, edited by E. S. Hall, pp. 299-351. National Museum of Man Mercury Series. Archaeological Survey of Canada, Paper 49, National Museums of Canada, Ottawa.

359

1979 Organization and Formation Processes: Looking at Curated Technologies. Anthropological Research. 35:255-273.

Journal of

1980 Willow Smoke and Dogs Tails: Hunter-Gatherer Settlement Systems and Archaeological Site Formation. American Antiquity 45:4-20. 1981 Bones: Ancient Men and Modern Myths.

Academic Press, New York.

1983 In Pursuit of the Past. Thames and Hudson, New York. 1989 Debating Archaeology. 1990 Mobility, Housing, and Environment: A Comparative Study. Research 46:119-52.

Journal ofAnthropological

1991 When the Going Gets Tough, the Tough Get Going: Nunamiut Local Groups, Camping Patterns, and Economic Organization. In Ethnoarchaeoiogicai Approaches to Mobile Campsites, HunterGatherer and Pastoralist Case Studies, edited by C. S. Gamble and W. A. Boismier, pp. 25-137. International Monographs in Prehistory, Ann Arbor. Binford, L. R. and J. F. O’Connell 1984 An Alywara Day: The Stone Quarry. Journal ofAnthropological Research 40:406-432. Reprinted in Debating Archaeology, edited by L. R. Binford, pp. 121-146. Academic Press, Orlando. Bingham, Sam 1996 The Last Ranch, A Colorado Community and the Coming Desert.

Pantheon Books, New York.

Black, K. D. 1993 Lithic Sources in the Rocky Mountains of Colorado. Paper presented at the First Biennial Rocky Mountain Anthropology Conference, September 30-October, 2, 1993, Jackson, Wyoming. Blakeslee, D. J. and R. Blasing. 1988 Indian Trails in the Central Plains. Plains Anthropologist 33:17-25. Blankholm, H. P. 1991 Intrasite Spatial Analysis in Theory and Practice. Aarhus University Press, Denmark. Bleed, P. 1986 The Optimal Design of Hunting Weapons: Maintainability or Reliability. American Antiquity 51:737-747. Boismier, W. A. 1991 Site Formation Among Sub-Arctic Peoples: An Ethnohistorical Approach. In Ethnoarchaeological Approaches to Mobile Campsites, Hunter-Gatherer and Pastoralist Case Studies, edited by C. S. Gamble and W. A. Boismier, pp. 189-214. International Monographs in Prehistory, Ann Arbor. Boldurian A. T. 1990 Lithic Technology at the Mitchell Locality of Blackwater Draw: A Stratified Folsoni Site in Eastern New Mexico. Plains Anthropologist, Memoir 24. 1991 Folsom Mobility and Organization of Lithic Technology: A View from Blackwater Draw, New Mexico. Plains Anthropologist 36:281-295.

360

1991 Folsom Mobility and Organization of Lithic Technology: A View from Blackwater Draw, New Mexico. Plains Anthropologist 36:281-295. Boldurian, A., A. G. Agogino, P. H. Shelley, and M. Slaughter 1987 Folsom Biface Manufacture, Retooling, and Site Function at the Mitchell Locality of Blackwater Draw. Plains Anthropologist 32(1 17):299-3 11. Boldurian A. T. and S. M. Hubinsky 1994 Preforms in Folsom Lithic Technology: A View from Blackwater Draw, New Mexico. Plains Anthropologist 39:445-464. Bomiichsen, R., D. Stanford, and J. L. Fastook 1987 Enviromnental Change and Developmental History of Human Adaptive Patterns: The Paleoindian Case In The Geology of North America, edited by W. F. Ruddiman and H. E. Wright, Jr., pp. 403-424. North America and Adjacent Oceans During the Last Deglaciation, vol. K-3, A. R. Pahner, general editor, The Geological Society of America, Boulder. Bordes, F. 1968 The Old Stone Age. McGraw-Hill, New York. Bowman, W. D., T. A. Theodose, and M. C. Fisk 1995 Physiological and Production Responses of Plant Growth Forms to Increases in Limiting Resources in Alpine Tundra: Implications for Differential Community Response to Environmental Change. Oecologia 101:217-227. Bradley, B. A. 1982 Lithic Technology. In The Agate Basin Site: A Record of the Paleoindian Occupation of the Northwestern High Plains, edited by G. C. Frison and D. J. Stanford, pp. 181-212. Academic

Press, New York.

1998 Personal Communication, Cortez, Colorado. Bradley, B. A. and G. C. Frison 1996 Flaked-Stone and Worked-Bone Artifacts from the Mill Iron Site. In The Mill Iron Site, edited by G. C. Frison, pp. 43-69. University of New Mexico Press, Albuquerque. Brain, C. K. 1981 The Hunters or the Hunted? An Introduction to African Cave Taphonomy.

University of Chicago

Press, Chicago. Brink, J. W. 1978 An Experimental Study of Microwear Formation on Endscrapers.

Archaeological Survey of

Canada Paper 83.

Broecker, W. S. 1990 Salinity History of the Northern Atlantic during the Last Deglaciation. 467. 1994 Massive Iceberg Discharges as Triggers for Global Change.

Paleoocenography 5:459-

Nature 372:421-424.

Broecker, W. S., M. Andree, W. Wolfli, H. Oescher, G. Bonani, J. P. Kennett, and D. M. Peteet 1988 The Chronology of the Last Deglaciation: Implication to the Cause of the Younger Dryas Event. Paleoocenography 3:1-19.

361

Broilo, F. 1971 An Investigation of Surface Collected Clovis, Folsom, and Midland Projectivle Points from Blackwater Draw and Adjacent Localities. Unpublished M. A. Thesis, Department of Anthropology, Eastern New Mexico University, Portales. 1973 Early Human Occupation of the Tularosa Basin. In Technical Manual: 1973 Survey of the Tularosa Basin, pp. 225-244. Human Systems Research, Inc. Tularosa, New Mexico. Broster, J. B. 1983 Paleoindian Adaptations to High Altitudes on Cebolleta Meas. In High-Altitude Adaptations in the Southwest, edited by J. C. Winter, pp. 108-114. Cultural Resources Management Report No. 2. U. S. Forest Service, Southwestern Region, Albuquerque. Brumback, H. J. and R. Jarvenpa 1997 Eth.noarchaeology of Subsistence Space and Gender: A Subarctic Dene Case. 62:414-436.

American Antiquity

Brunstein, F. C. and D. K. Yamaguchi 1992 The Oldest Known Rocky Mountain Bristlecone Pines (Pinus aristata Engelm.). Arctic and Alpine Research 24:253-256. Bryan, K. and A. P. Butler 1940 Artifacts Made of the Glassy Andesite of San Antonio Mountain, Rio Arriba County, New Mexico. University of New Mexico Bulletin, No. 349, Anthropological Series, 3(4):27-3 1. Bryson, R. A. 1995 Annual First Approximations for Modeled Precipitation (950831) and Temperature (950904)Alamosa, Colorado. Graphs on file, Paleoindian/Paleoecology Program, Smithsonian Institution, Washington, D.C. Bryson, R. A. and R. U. Bryson 1995 Site-Specific High-Resolution Climate Models of Paleoindian Sites in the Plains: Hell Gap, Agate Basin, Alamosa, and Blackwater Draw. Paper presented at the 53 id Annual Plains Conference, Laramie, Wyoming. Burch, Everett 1998 Videotaped in Ignacio, Colorado, October 1998 by Pegi Jodry and Ted Timrek. Department of Anthropology, Paleoindian/Paleoecology Program, Smithsonian Institution, Washington, D.C. Burgett, G. R. 1990 The Bones of the Beast: Resolving Questions of Faunal Assemblage Formation Processes Through Actualistic Research. Unpublished Ph.D. dissertation, University of New Mexico, Albuquerque. Button, V. T. 1987 The ClosedBasin of Colorado’s San Luis Valley: Bureau of Reclamation Archeological Investigations 1976-86. U.S. Department of Interior, Bureau of Reclamation, Southwest Region, Amarillo, Texas. Callahan, E. 1979 The Basics of Biface Knapping in the Eastern Fluted Point Tradition: A Manual for Flintknappers and Lithic Analysts. Archaeology of Eastern North America 7:1-180.

362

Carrara, P. E., W. N. Mode, M. Ruben, and S. W. Robinson 1984 Deglaciation and Postglacial Timberline in the San Juan Mountains, Colorado. Quaternary Research 21:42-55. Carrara, P. E., D. A. Trimble, and M. Rubin 1991 Holocene treeline fluctuations in the Northern San Juan Mountains, Colorado, U.S.A., As Indicated by Radiocarbon-dated Conifer Wood. Arctic andAlpine Research 23:233-246. Cashadan, E. 1985 Coping with Risk: Reciprocity Among the Basarwa of Northern Botswana. Cassell, R. K. 1941 A Postulated Corridor of Folsom Migration.

Man 20: 454-74.

Papers of the Michigan Academy of Science, Arts

and Letters 26:451-57.

Catlin, G. 1926 Letters and Notes on the North American Indians. 2 volumes. Edinburgh.

Chambellan, C., M. Kadziel, T. J. Lennon, and E. K. Wade 1984 A Cultural Resource Evaluation of Site 5CF84, Salida Ranger District, Pike and San Isabel National Forests, Colorado. Prepared by Western Cultural Resource Management, Inc, Boulder,

Colorado. Prepared Under Contract No. R2-12-83-72 Between Western Cultural Resource Management, Inc and U. S. Forest Service, Salida Ranger District Pike and San Isabel National Forests, Colorado. Charles, M. C. 1995 A Cultural Resources Inventory of the Clauson Timber Sale and the Mosca Road Rehabilitation Projects, Hinsdale County, Colorado. Report submitted to the San Juan National Forest, Columbine District, Durango. Christensen, D. 1997 Professional hide-tanner from Big Timber, Montana. Consultant regarding manufacture of Plains ethnographic tipis and clothing for the Native American Program and the Paleoindianl Paleoecology Program, Department of Anthropology, National Museum of Natural History, Smithsonian Institution. Clark, D. H. and A. R. Gillespie 1997 Timing and Significance of Late-glacial and Holocene Cirque Glaciation in the Sierra Nevada, California. Quaternary International38(29): 21-38. Clark, J. D. and H. Kurashina 1981 A Study of the Work of a Modern Tanner in Ethiopia. In Modern Material Culture, The Archaeology of Us, edited by R. A. Gould and M. B. Schiffer, pp.303-342. Academic Press, New York. Clayton, L., W. B. Bickey, Jr., and W. J. Stone 1970 Knife River Flint. Plains Anthropologist 15:282-290. Coe, M. J., D. H. Cummings, J. Phillipson 1976 Biomass and Production of Large African Herbivores in Relation to Rainfall and Primary Production. Oecologia 22:341-354.

363

Collins, M. B. 1975 Lithic Technology as a Means of Processual Inference. In Lithic Technology, Making and Using Stone Tools, edited by E. Swanson, pp. 14-34. Mouton, The Hague. 1988 The Archeological Sequence at Kincaid, Uvalde County, Texas. Transactions of the 2511 Regional Archeological Symposium for Southeastern New Mexico and Western Texas, pp. 25-34. Southwest Federation of Archaeological Societies, Midland. 1999 Clovis and Folsom Lithic Technology on and Near the Southern Plains: Similar Ends, Different Means. In Folsom Lithic Technology, Explorations in Structure and Variation, edited by D. S. Amick, pp. 12-38. International Monographs in Prehistory, Archaeological Series 12, Ann Arbor. Colville, R. M. 1996 La Vereda, A Trail Through Time.

San Luis Valley Historical Society, Alamosa, Colorado.

Cook, H. J. 1927 New Geological and Paleontological Evidence Bearing on the Antiquity of Mankind in America. Natural History 7(3):240-247. Copprock, D.L., J. K. Detling, J. E. Ellis, and M. I. Dyer 1983 Plant-herbivore Interactions in a North American Mixed-Grass Prairie. I. Effects of Black-Tailed Prairie Dogs on Intraseasonal Aboveground Plant Biomass and Nutrient Dynamics and Plant Species Diversity. Oecologia 56:1-9. Cotter, J. 1937 The occurrence of flints and extinct animals in pluvial deposits near Clovis, New Mexico, Part VI: Report on the excavations at the gravel pit in 1937. Proceedings of the Philadelphia Academy of Natural Sciences 90:113-117. 1938 The occurrence of flints and extinct animals in pluvial deposits near Clovis, New Mexico, Part IV: Report on the excavations at the gravel pit, 1936. Proceedings of the Philadelphia Academy of Natural Sciences 89:1-16. Crabtree, D. B. 1966 A Stoneworker’s Approach to Analyzing and Replicating the Lindenmeier Folsom. Tebiwa 9(1)3-39. Creel, D. 1991 Bison Hides in Late Prehistoric Exchange in the Southern Plains.

American Antiquity 56:40-49.

Crouch, T. M. 1985 Potentiometric Surface, 1980, and Water-Level Changes, 1969-80, in the Unconfined ValleyFill Aquifers of the San Luis Basin, Colorado and New Mexico. U.S. Geological Survey Atlas

HA-683, Denver. Dames and Moore 1985 Final investigative Report Volume 1, Baseline Vegetation and Wetland Resource Investigation, Closed Basin Division, San Luis Valley Project, Colorado. Dames and Moore, Golden, Colorado.

Dansgaard, W., J. W. C. White, and S. J. Johnson 1989 The Abrupt Termination of the Younger Dryas Climate Event.

Nature 339:532-533.

364

Dansgaard, W., S. Johnsen, H. Clausen, D. Dahl-Jensen, N. Gundestrup, C. Jammer, C. Hvldberg, J. Steffensen, A. Sveinbjornsdottir, J. Jouzel, and G. Bond 1993 Evidence for General Instability of Past Climate from a 250-kyr ice-core record. Nature 364:2 18220.

Davis, L. B. 1993 Paleo-Indian Archaeology in the High Plains and Rocky Mountains of Montana. In From Kostenki to Clovis, Upper Paleolithic-Paleo-Indian Adaptations, edited by 0. Soffer and N. D. Praslov, pp. 263-277. Davis, L. B. and S. T. Greiser 1992 Indian Creek Paleoindians: Early Occupation of the Elkhorn Mountains Southeast Flank, WestCentral Montana. In Ice Age Hunters of the Rockies, edited by D. J. Stanford and J. Day, pp. 225283. University Press of Colorado, Niwot. Davis L. B., S. T. Greiser, and N. Toth 1985 Use-Wear Analyses of Paleothdian Unifaces from the Initial Late-Glacial Occupation of the Indian Creek Site. Current Research in the Pleistocene 2:45-46. Davis, 0. K. 1987 Spores of the Dung Fungus Sporormiella: Increased Abundance in Historic Sediments and Before Pleistocene Megafaunal Extinction. Quaternary Research 28:290-294. Davis, 0. K. and L. Agenbroad, P. S. Martin, and J. I. Mead 1984 The Pleistocene Dung Blanket of Bechan Cave, Utah. Museum of Natural History 8:267-282.

Special Publications of the Carnegie

Davis, 0. R. and D. S. Shafer 1991 Sporormiella: An Indication of the Clovis-Folsom Biomass Collapse? Paper presented at the 56t 1 Annual Meeting of the Society for American Archaeology, New Orleans. Davis, W. A. 1985 The Montgomery Folsom Site. Current Research in the Pleistocene 2:11-12. Dawson. J. and W. J. Judge 1969 Paleoindian Sites and Topography in the Middle Rio Grande Valley of New Mexico. Anthropologist 14:149-163. Dawson, J. and D. Stanford 1975 The Linger site: A Reinvestigation.

Plains

Southwest Lore 41(4): 22-28.

De Lanois, J. L. 1993 Climate Change during the Late Holocene from a South Central Colorado Lake. Unpublished Master’s Thesis, Department of Geosciences, University of Arizona, Tucson. De Mazancourt, C., M. Loreau, and L. Abbadie 1998 Grazing Optimization and Nutrient Cycling: When Do Herbivores Enhance Plant Production? Ecology 79:2242-2252 Del Bene, T. 1979 Once Upon a Striation: Current Models of Striation and Polish Formation. Analysis, edited by B. Hayden, pp. 167-177. Academic Press, New York.

InLithic Use-Wear

365

Deller, D. B. 1989 Interpretation of Chert Type Variation in Paleoindian Industries, Southwestern Ontario. In Eastern Paleoindian Lithic Resource Use, edited by C. J. Ellis and J. C Lothrop, pp. 191-220. Westview Press, Boulder. Deller, D. B. and C. J. Ellis 1992 ThedfordIl, A Paleo-Indian Site in the Ausable River Watershed

of Southwestern Ontario.

Memoirs Museum of Anthropology, University of Michigan, Number 24, Ann Arbor. Detling, J. 1999 Personal Communication. Natural Resource Ecology Laboratory, Colorado State University, Fort Collins, Colorado. Dibble, D. S. and D. Lorrain 1968 Bonfire Shelter: A Stratified Bison Kill Site, Vat Verde County, Texas.

Miscellaneous Papers 1.

Texas Memorial Museum, The University of Texas, Austin. Dillehay, T. D. 1997 Monte Verde, A Late Pleistocene Settlement in Chile. Volume 2: The Archaeological Context and Interpretation. Smithsonian Institution Press, Washington DC.

Doesken, N. J. and T. B. McKee 1989 The Incredible Climate of the San Luis Valley. In Water in the Valley, A 1989 Perspective on Water Supplies, Issues, & Solutions in the San Luis Valley, Colorado, edited by E.J. Harmon, pp. 80-98. Colorado Ground-Water Association, Denver. Driver, J. C. 1990 Meat in Due Season: the Timing of Communal Hunts. Hunters of the Recent Past, edited by L. B. Davis and B. 0. Reeves, pp. 11-33. Unwin Hyman, London. Edwards, R. , J. Beck, G. Burr, D. Donahue, J. Chappell, A. Bloom, E. Druffel, and F. Taylor 1993 A Large Drop in Atmospheric 12C/ 12C and Reduced Melting in the Younger Dryas, Documented with 230 Th Ages of Corals. Science 260(5110):962-968. Elias, S. A. 1996 Late Pleistocene and Holocene Seasonal Temperatures Reconstructed from Fossil Beetle Assemblages in the Rocky Mountains. Quaternary Research 46:311-318. Elias, S. A., P. E. Toolin, and A. J. T. Jull 1991 Revised Age of Deglaciation of Lake Emma Based on New Radiocarbon and Macrofossil Analyses. Quaternary Research 36:307-321. Ellis, C. J. 1989 The Explanation of Northeastern Paleoindian Lithic Procurement Patterns. In Eastern Paleoindian Lithic Resource Use, edited by C. J. Ellis and J. C. Lothrop, pp. 139-164. Westview Press, Boulder. Ellis, C. and J. H. Payne 1995 Estimating Failure Rates in Fluting Based on Archaeological Data: Examples from NE North America. Journal of Field Archaeology 22:459-474. Emerson. A. M. 1990 Archaeological Implications of Variability in the Economic Anatomy of Bison bison. Unpublished Ph.D. Dissertation. Washington State University, Pullman.

366

Emery, P. A. 1996 Hydrogeology of the San Luis Valley, Colorado, An Overview-And a Look at the Future. In Geologic Excursions to the Rocky Mountains and Beyond, Field Trip Guidebook for the 1996 Annual Meeting, Geological Society ofAmerica, Denver, Colorado October 28-3 1, Special Publication 44, CD-ROM, edited by R. A. Thompson, M. R. Hudson, and Charles L. Pilimore. Colorado Geological Survey, Denver. Emery, P. A., A. J. Boettcher, R. J. Snipes, and H. J. Mclntrye, Jr. 1971 Hydrology of the San Luis Valley, south-central Colorado. Inventory Atlas HA-381, Denver.

U.S. Geological Survey Hydrologic

Emery, S. and D. Stanford 1982 Preliminary Report on Archaeological Investigations at the Cattle Guard Site, Alamosa County, Colorado. Southwestern Lore 48:10-20. Enloe, J. G. 1993 Ethnoarchaeology of Marrow Cracking: Implications for the Recognition of Prehistoric Subsistence Organization. In . In From Bones to Behavior, Ethnoarchaeological and Experimental Contributions to the Interpretation of Faunal Remains, edited by J. Hudson, pp. 8297. Center for Archaeological Investigations, Southern Illinois University at Carbondale, Occasional Paper No. 21. Enloe, J. G., F. David, and T. S. Hare 1993 Patterns of Faunal Processing at Section 27 of Pincevent: The Use of Spatial Analysis and Journal of Ethnoarchaeological Data in the Interpretation of Archaeological Site Structure. Anthropological Archaeology 13:105-124. Ewers, J. C. 1945 Blackfeet Crafts. Department of the Interior, Bureau of Indian Affairs, Haskell Institute, Lawrence, Kansas. Fall, P. L. 1994 Pollen Accumulation in a Montane Region of Colorado, USA: A Comparison of Moss Poisters, Atmospheric Traps, and Natural Basins. Review of Palaeobotany and Palynology 72:169-197. 1997 Timberline Fluctuations and Late Quaternary Paleoclimates in the Southern Rocky Mountains, Colorado. GSA Bulletin 109(10):1306-1320. Fawcett, W. B. 1987 Communal Hunts, Human Aggregations, Social Variation, and Climatic Change: Bison Utilization by Prehistoric Inhabitants of the Great Plains. Unpublished Ph.D. dissertation, University of Massachusetts, Amherst. University Microfilms, Ann Arbor. Feiring, C. R. 1990 Late Quaternary Geology and Geoarchaeology of the Upper Trinity River Drainage Basin, Texas. Guidebookfor Geological Society ofAmerica Fieldtrip 11. Dallas Geological Society, Dallas. 1995 Subsistence and Mobility Patterns at the Aubrey Clovis Site, Texas. In Abstracts of the 60th Annual Meeting of the Society for American Archaeology, Minneapolis, pp. 74-75. Society for American Archaeology, Washington, D.C. Figgins, J. D. 1927 The Antiquity of Man in America. Natural History 27(3):229-239.

367

Fiorillo, A. R. 1989 An Experimental Study of Trampling: Implications for the Fossil Record. In Bone Modification, edited by R. Bonnichsen and M. Sorg, pp.61-72. Publication of the Center for the Study of the First Americans, Orono. Fisher, J. W. and H. C. Strickland 1991 Dwellings and Fireplaces: Keys to Efe Pygmy campsite Structure. In Ethnoarchaeo logical Approaches to Mobile Campsites, Hunter-Gatherer and Pastoralist Case Studies, edited by C.S. Gamble and W. A. Boismier, pp. 2115-236. International Monographs in Prehistory, Ethnoarchaeological Series 1, Ann Arbor. Fisk, M. C., S. K. Schmidt, and T. R. Seastedt 1998 Topographic Patterns of Above- and Belowground Production and Nitrogen Cycling in Alpine Tundra. Ecology 79(7):2253-2266. Flenniken, J. J. 1978 Reevaluation of the Lindenrneier Folsom: A Replication Experiment in Lithic Technology. American Antiquity 43:473-480. Fletcher, A. C. and F. La Flesche 1911 The Omaha Tribe. Extract from the Twenty-Seventh Annual Report of the Bureau of American Ethnology. Government Printing Office, Washington D.C. Forbis, R. G. and J. D. Sperry 1 953An Early Man Site in Montana. American Antiquity 18:127-133. Frank, D. A. and P. M. Groffman 1998 Ungulate VS Landscape Control of Soil C and N Processes in Grasslands of Yellowstone National Park. Ecology 79(7):2229-2241. Friedman, I., P. Carrara, and J. Gleason 1988 Isotopic Evidence of Holocene Climatic Change in the San Juan Mountains, Colorado. Quaternary Research 30:350-353. Frison, G. C. 1967 Archaeological Evidence of the Crow Indians in Northern Wyoming: A Study of a Late Prehistoric Period Buffalo Economy. Unpublished Ph.D. Dissertation, University of Michigan, Ann Arbor. University Microfilms, Ann Arbor. 1974 The Casper Site: A Hell Gap Bison Kill on the High Plains. Academic Press, New York. 1977 Paleoindian Sites and Economic Orientations in the Big Horn Basin. 116.

The Museum Journal 17:97-

1978 Prehistoric Hunters of the High Plains. Academic Press, New York. 1979 Observations on the Use of Tools: Dulling of Working Edges on Some Chipped Stone Tools in Bison Butchering. In Lithic Use-Wear Analysis, edited by B. Hayden, pp. 259-268. Academic Press: New York. 1982a Folsom Components. In The Agate Basin Site: A Record of the Paleoindian Occupation of the Northwestern High Plains, edited by G. C. Frison and D. J. Stanford, pp. 37-76. Academic Press, New York.

368

1982b Raw Stone Flaking Material Sources. In The Agate Basin Site A Record of the Paleoindian Occupation of the Northwestern High Plains, edited by G. C. Frison and D. J. Stanford, pp. 173178. Academic Press, New York. 1 982c Bison Dentition Studies. In The Agate Basin Site: A Record of the Paleoindian Occupation of the Northwestern High Plains, edited by G. C. Frison and D. J. Stanford, pp. 240-260. Academic Press, New York. 1982d Paleo-Indian Winter Subsistence Strategies on the High Plains. In Plains Indian Studies: A Collection of Essays in Honor of John C. Ewers and Waldo Wedel, edited by D. H. Ubelaker and H. J. Viola, pp. 193-201. Smithsonian Contributions to Anthropology Number 30, Wadhington DC. 1984 The Carter/Kerr-McGee Paleoindian Site: Cultural Resource Management and Archaeological Research. American Antiquity 49:288-314. 1987 Prehistoric, Plains-Mountain, Large-Mammal Communal Hunting Strategies. In The Evolution of Human Hunting, edited by M. H. Nitecki and D. V. Nitecki, pp. 177-223. Plenum Press, New York. 1991

Prehistoric Hunters of the High Plains. Second Edition. Academic Press, San Diego.

1992 The Foothills-Mountains and the Open Plains: The Dichotomy in Paleoindian Subsistence Strategies Between Two Ecosystems. In Ice Age Hunters of the Rockies, edited by D. J. Stanford and J. Day, pp. 323-342.. University Press of Colorado, Niwot. 1993 North American High Plains Paleo-Indian Hunting Strategies and Weaponry Assemblages. In From Kostenki to Clovis Upper Paleolithic-Paleo-Indian Adaptations, edited by 0. Soffer and N.D. Praslov, pp.237-249. Plenum Press, New York. Frison, G. C. (editor) 1996 The Mill Iron Site. University of New Mexico Press, Albuquerque. Frison, G. C. and B. A. Bradley 1980 Folsom Tools and Technology at the Hanson Site, Wyoming. Albuquerque. 1999

The Fenn Cache, Clovis Weapons and Tools. New Mexico.

University of New Mexico Press,

One Horse Land and Cattle Company, Santa Fe,

Frison, G. C. and C. Craig 1982 Bone, Antler, and Ivory Artifacts and Manufacture Technology. In The Agate Basin Site: A Record of the Paleoindian Occupation of the Northwestern High Plains, edited by G. C. Frison and D. J. Stanford, pp. 157-173. Academic Press, New York. Frison, G. C. and C. A. Reher 1970 Appendix 1: Age Determination of Buffalo by Tooth Eruption and Wear. In The Glenrock Buffalo Jump, 48CO304: Late Prehistoric Period Buffalo Procurement and Butchering in the Northwest Plains, by G. C. Frison. Plains Anthropologist Memoir No. 7, 46-50. Frison, G. C. and D. J. Stanford (editors) 1982 The Agate Basin Site: A Record of the Paleoindian Occupation of the Northwestern High Plains. Academic Press, New York.

369

Frison, G. C. and L. C. Todd (editors) 1987 The Homer Site The Type Site of the Cody Cultural Complex.

Academic Press, Orlando.

Fuller, W. A. 1966 The Biology and Management of the Bison of Wood Buffalo National Park. Bulletin Series 1(16). Canadian Wildlife Service. Gallagher, J. P. 1977 Contemporary Stone Tools in Ethiopia: Implications for Archaeology. Archaeology 4:407-414.

Wildlife Management

Journal of Field

Galloway, E. and G. Agogino 1961 The Johnson Site: A Folsom Campsite. Plains Anthropologist 6:205-208. Gamble, C. S. 1986 The Palaeolithic Settlement of Europe.

Cambridge University Press, Cambridge.

1991 An Introduction to the Living Spaces of Mobile Peoples. In Ethnoarchaeological Approaches to Mobile Campsites, Hunter-Gatherer and Pastoralist Case Studies, edited by C. S. Gamble and W. A. Boismier, pp. 1-23. International Monographs in Prehistory, Ann Arbor. Gaulin, S. J. and H. A. Hoffman 1988 Evolution and Development of Sex Differences in Spatial Ability. In Human Reproductive Behavior: A Darwinian Perspective, edited by L. Betzig, M. Borgerhoff Mulder and P. Turke, pp. 129-152. Cambridge University Press, London. Gero, J. M. 1989 Assessing Social Information in Material Objects: How Well do Lithics Measure Up? In Time, Energy and Stone Tools, edited by R. Torrence, pp.92-105. Cambridge University Press, Cambridge. 1990 Genderlithics: Women’s Roles in Stone Tool Production. In Engendering Archaeology: Women and Prehistory, edited by J. M. Gero and W. M. Conkey, pp. 163-193. Basil Blackwell, Oxford. Gifford-Gonzalez, D. 1993 Gaps in the Zooarchaeological Analyses of Butchery: Is Gender an Issue. In From Bones to Behavior, Ethnoarchaeological and Experimental Contributions to the Interpretation of Faunal Remains, edited by J. Hudson, pp. 181-199. Center for Archaeological Investigations, Southern Illinois University at Carbondale, Occasional Paper No. 21. Gifford, D. P. and A. K. Behrensrneyer 1977 Observed Formation and Burial of a Recent Human Occupation site in Kenya. Research 8:245-266.

Quaternamy

Gifford-Gonzalez, D. P., D. B. Damrosch, D. R. Damrosch, J. Pryor, and R. L. Thunen 1985 The Third Dimension in Site Structure: An experiment in Trampling and Vertical Dispersal. American Antiquity 50:803-818. Goode, G. 1999 Chipped Stone Resources of Central Texas. Paper presented at the Folsom Conference II, Austin, Texas March 2-6, 1999.

370

Goodyear, A. C. 1974 The Brand Site: A Techno-Functional Study of a Dalton Site in Northeast Arkansas.

Arkansas

Archeological Survey, Publications on Archeology, Research Series No. 7. 1979 A Hypothesis for the Use of C’ryptocrystalline Raw Materials Among Paleoindian Groups of North America. University of South Carolina, Institute of Archaeology and Anthropology, Research

Manuscript Series No. 156. Gould, R. A. 1980 Living Archaeology. Cambridge University Press, Cambridge. Gould, R. A., D. A. Koster, and A. H. L. Sontz 1971 The Lithic Assemblage of the Western Desert Aborigines of Australia. American Antiquity 36(2):149-169. Gould, R. A. and S. Saggers 1985 Lithic Procurement in Central Australia: A Closer Look at Binfords Idea of Embeddedness in Archaeology. American Antiquity 50:117-136. Graham, R. W. 1978 Identification of canid bones at the Linger Site. Vertebrate Paleontologist, Denver Museum of Natural History, Denver, Colorado. Graham, R. W. and E. L. Lundelius, Jr. 1984 Coevolutionary Disequilibrium and Pleistocene Extinctions. In In Quaternary Extinctions A Prehistoric Revolution, edited by P. S. Martin and R. G Klein, pp. 223-249. University of Arizona Press, Tucson. Greaves, R. D. 1997 Etlmoarchaeological Investigation of Subsistence Mobility, Resource Targeting, and Technological Organization Among Pume Foragers of Venezuela. Unpublished Ph.D. dissertation, University of New Mexico, Albuquerque. University Microfilms, Ann Arbor. Gregg, S. A., K. W. Kintigh, and R. Whallon 1991 Linking Ethnoarchaeological Interpretation and Archaeological Data: The Sensitivity of Spatial Analytical Methods to Postdepositional Disturbance. In The Interpretation ofArchaeological Spatial Patterning, edited by E. K. Kroll and T. D. Price, pp. 149-196. Plenum Press, New York. Greiser, S. T. 1983 A Preliminary Statement About Quarrying Activity at Flattop Mesa. 14.

Southwestern Lore 49(4):6-

Green, T., B. Cochran, T. Fenton, J. Woods, G. Titmus, L. Tieszen, M. Davis, and S. Miller 1998 The Buhl Burial: A Paleoindian Woman From Southern Idaho. American Antiquity 63:437-456. Grigg, L. D. and C. Whitlock 1998 Late-Glacial Vegetation and Climate Change in Western Oregon. 298. Grinnell, G. B. 1923 The Cheyenne Indian, Volume I.

Quaternary Research 49:287-

Yale University Press, New Haven.

Grimm, E. C. 1988 Data Analysis and Display. In Vegetation History, edited by B. Huntley and T. Webb, pp. 43-76. Kluwer Academic, Dordrecht, Netherlands.

371

Grove, B. 1997-99 Personal Communications, Ignacio, Colorado. Videotape in Ignacio, October 1998 by Pegi Jodry and Ted Timrek. Department of Anthropology, Pa!eoindian/Paleoecology Program, Smithsonian Institution, Washington, D.C. Guthrie, R. D. 1984 Mosaics, Allelochemics and Nutrients, An Ecological Theory of Late Pleistocene Megafaunal Extinctions. In Quaternary Extinctions A Prehistoric Revolution, edited by P. S. Martin and R. 0 Klein, pp. 259-298. University of Arizona Press, Tucson. Hagar, Ivol 1977 Site Inventory Record, Site No. 5CF84. Office of Archaeology and Historic Preservation, Colorado Historical Society, Denver. Harrison, B. R. and K. L. Killen 1978 Lake Theo: A Strat?/Ied, Early Man Bison Butchering and Camp Site, Briscoe County, Texas. Special Archeological Report No. 1. Panhandle-Plains Historical Museum, Canyon, Texas. Harrison, B. R. and H. C. Smith 1975 A Test Excavation of the Lake Then Site, Briscoe County, Texas. Panhandle-Plains Historical Review 48:70-106. Haury, C. 1979 Characterization of the Chert Resources of the El Dorado Project Area. In Finding, Managing, and Studying Prehistoric Cultural Resources at El Dorado Lake, Kansas (Phase I), edited by G. R. Leaf, pp. 209-227. University of Kansas Museum of Anthropology Research Series, Number 2, Lawrence. Hawley, A. W. L., D. G. Peden, and W. R. Stricklin 1981 Bison and Hereford Steer Digestion of Sedge Hay. Canadian Journal ofAnimal Science 61:165174. Hayden, B. 1979 Snap, Shatter, and Superfractures: Use-Wear of Stone Skin Scrapers. In Lithic Use-Wear Analysis, edited by B. Hayden, pp. 207-229. Academic Press, New York. 1981 Subsistence and Ecological Adaptations of Modern Hunter/Gatherers. Omnivorous Primates, Gathering and Hunting in Human Evolution, edited by R. S. 0. Harding and G. Teleki, pp. 344421. 1982 Interaction Parameters and the Demise of Paleo-Indian Craftsmanship. 27:109-123. 1986a Resource Models of Inter-Assemblage Variability. 1986b Use and Misuse: The Analysis of Endscrapers.

Plains Anthropologist

Lithic Technology 15(3):82-89.

Lithic Technology 15:65-70.

Hayden, B. (editor) 1979 Lithic Use- Wear Analysis. Academic Press, New York. Hayden, B. and R. Gargett 1988 Specialization in the Paleolithic. Lithic Technology 17:12-18.

372

Haynes, C.V., Jr. 1971 Time, Environment, and Early Man. Arctic Anthropology 8:3-14. 1975 Pleistocene and Recent Stratigraphy. In Late Pleistocene Environments of the Southern High Plains, edited by F. Wendorf and J. J. Hester, pp. 57-96. Fort Burgwin Research Center Publication No. 9. Southern Methodist University, Dallas. 1980 Geochronology and Paleoenvironments of the Murray Springs Clovis Site, Arizona. Geographic Society Research Reports 13:243-251.

National

1984 Stratigraphy and Late Pleistocene Extinction in the United States. In Quaternaiy Extinctions A Prehistoric Revolution, edited by P. S. Martin and R. G Klein, pp. 345-353. University of Arizona Press, Tucson. 1991 Geoarchaeological and Paleohydrological Evidence for a Clovis-Age Drought in North America and its Bearing on Extinction. Quaternary Research 35:438-450. 1993 Clovis-Folsom Geochronology and Climatic Change. In From Kostenki to Clovis Upper Paleolithic-Paleo-Indian Adaptations, edited by 0. Soffer and N.D. Praslov, pp.219-236. Plenum Press, New York. 1995 Geochronology of Paleoenvironmental Change, Clovis Type Site, Blackwater Draw, New Mexico. Geoarchaeology 10:317-388. Haynes, C.V., Jr. and G. A. Agogino 1986 Geochronology of Sandia Cave. Smithsonian Contributions to Anthropology No. 32. Washington, D. C. Haynes, C. V., Jr., D. Stanford, M. Jodry, J. Dickenson, J. Montgomery, P. Shelley, and G. A. Agogino 1999 A Clovis Well at the Type Site, 11,500 B.C.: The Oldest Prehistoric Well in America. Geoarcheology 14:455-470. Haynes, G. 1981 Bone Modifications and Skeletal Disturbances by Natural Agencies: Studies in North America. Unpublished Ph.D. dissertation, The Catholic University of America, Washington, D.C. University Microfilms, Ann Arbor. Heinrich, P. V. 1984 Appenix C: Petrographic Analysis of Jasper From 5CF84 Chaffee County, Colorado. In A Cultural Resource Evaluation of Site 5CF84, Salida Ranger District, Pike and San Isabel National Forests, Colorado, by C. Chambellan, M. Kadziel, T. J. Lennon, and E. K. Wade, pp. 96-103. Prepared Under Contract No. R2-12-83-72 for the U. S. Forest Service, Salida Ranger District Pike and San Isabel National Forests, Colorado by Western Cultural Resource Management, Inc, Boulder. Henderson, J. and G. T. Goode 1991 Pavo Real: An Early Paleoindian Site in South-Central Texas. Pleistocene 8:26-28.

Current Research in the

Hester, J. J. 1962 A Folsom Lithic Complex from the Elida Site, Roosevelt County, New Mexico. 69:92-113.

El Palacio

373

1972 Blackwater Locality No. 1 A StratfIed, Early Man Site in Eastern New Mexico. Fort Burgwin Research Center, No. 8, Southern Methodist University, Dallas.

Publication of

Hill, M. G. 1994 Subsistence Strategies by Folsom Hunters at Agate Basin Wyoming: A Taphonomic Analysis of the Bison and Pronghorn Assemblages. Unpublished Master’s thesis, University of Wyoming, Laramie. Hill, M. E. and J. L. Hofrnan 1997 The Waugh Site: A Folsom-Age Bison Bonebed in Northwestern Oklahoma. Anthropologist 42:63-83.

Plains

Hoard, R. J., S. R. Holen, M. D. Glascock, H. Neff, and J. M. Elam 1992 Neutron Activation Analysis of Stone from the Chadron Formation and a Clovis Site on the Great Plains. Journal ofArchaeological Science 19:655-665. Hofinan, J. L. 1980 Scapula Skin-Dressing and Fiber-Processing Tools.

Plains Anthropologist 25:135-141.

1986a An Initial Survey of the Folsom Complex in Oklahoma. Paper Presented at the Conference, Denver.

44th Annual Plains

1986b Eva Projectile Point Breakage at Cave Spring: Pattern Recognition and Interpretive Possibilities. Midcontinental Journal ofArchaeology 11:85. 1988 Folsom Research in Western Oklahoma: Rethinking the Folsom Occupation of the Southern Plains. Transactions of the Twenty-third Regional Archeological Symposium for Southeastern New Mexico and Western Texas, pp. 6-30. Southwestern Federation of Archeological Societies, Amarillo. 1990a Cedar Creek: A Folsom Locality in Southwestern Oklahoma. 7:19-23.

Current Research in the Pleistocene

1990b Salt Creek, Recent Evidence From the Eastern Folsom Margin in Central Oklahoma. Anthropologist 35:367-374.

Plains

1991 Folsom Land Use: Projectile Point Variability as a Key to Mobility. In Raw Material Economies Among Prehistoric Hunter-Gatherers, edited by A. Montet-White and S. Holen, pp. 335-355. University of Kansas Publications in Anthropology 19, Lawrence. 1992 Recognition and Interpretation of Folsom Technological Variability on the Southern Plains. In Ice Age Hunters of the Rockies, edited by D. J. Stanford and J. Day, pp. 193-224. University Press of Colorado, Niwot. 1993 Kansas Folsom Evidence. The Kansas Anthropologist 15(2):31-43. 1994a Paleoindian Aggregations on the Great Plains. Journal of Anthropological Archaeology 13:34 1370. 1994b Personal Communication. Department of Anthropology. University of Kansas, Lawrence. 1995 Dating Folsom Occupations on the Southern Plains: The Lipscomb and Waugh Sites. Field Archaeology 22:421-437.

Journal of

374

1996 Early Hunter-Gatherers of the Central Great Plains: Paleoindian and Mesoindian (Archaic) Cultures. In, Archeology and Paleoecology of the Central Great Plains, edited by J. L. Hofman, pp. 41-100. Arkansas Archeological Survey Research Series No. 48, Fayetteville. 1999 Folsom Fragments, Site Types and Prehistoric Behavior. In Folsom Lithic Technology: Explorations in Structure and Variation, edited by D. S. Amick, pp. 122-143. International Monographs in Prehistory, Archaeological Series 12, Ann Arbor. Hofman, J. L., D. S. Amick, and R. 0. Rose 1990 Shifting Sands: A Folsom-Midland Assemblage From a Campsite in Western Texas. Anthropologist 35:221-253. Hofman, J. L. and E. E. Ingbar 1988 A Folsom Hunting Overlook in Eastern Wyoming.

Plains

Plains Anthropologist 33:337-350.

Hofman, J. L., L. C. Todd, and M. B. Collins 1991 Identification of Central Texas Edwards Chert at the Folsom and Lindenmeier Sites. Anthropologist 36:297-308.

Plains

Hofman, J. L., L. C. Todd, C. B. Schultz, and W. J. Hendy 1989 The Lipscomb Bison Quarry: Continuing Investigations at a Folsom Kill-Butchery Site on the Southern Plains. Bulletin of the Texas Archeological Society 60:149-189. Hofinan, J. L. and D. G. Wyckoff 1987 Folsom Components at the Winters and Beckner Sites. Current Research in the Pleistocene 4:1011. Holland, E. A. and J. K. Detling 1990 Plant Responses to Herbivoiy and Belowground Nitrogen Cycling.

Ecology 71:1040-1049.

Holland, E. A., W. J. Patron, J. K. Detling, and D. L. Copprock 1992 Physiological Responses of Plant Production to Herbivory and Their Consequences for Ecosystem Nutrient Flow. American Naturalist 140:685-706. Holliday, V. T. and Meltzer, D. J. 1996 Geoarchaeology of the Midland (Paleoindian) Site, Texas.

American Antiquity 61:755-771.

Holsworth, W. N. 1960 Interactions between Elk and Buffalo in Elk Island National Park, Alberta. Unpublished Master’s thesis. University of British Columbia, Vancouver. Hopper, R. M., A. D. Geis, J. R. Grieb, and L. Nelson, Jr. 1975 Experimental Duck Hunting Seasons, San Luis Valley, Colorado, 1963-1970. Monographs 46. Howard, E. B. 1935 Evidence of Early Man in North America. Peimslyvania Museum, Philadelphia.

Wildlife

Museum Journal 24, Nos. 2-3. University of

Huckell, B. B. 1979 Of Chipped Stone Tools, Elephants, and the Clovis Hunters: An Experiment. Anthropologist 24:77-178.

Plains

375

1982 The Distribution of Fluted Points in Arizona: A Review and an Update. Cultural Resource Management Division, Arizona State Museum, University of Arizona, Archaeological Series No. 145. Hudson, J. (editor) 1993 From Bones to Behavior, Ethnoarchaeological Contributions to the Interpretation of Fauna! Remains. Center for Archaeological Investigations, Southern Illinois University Occasional Paper Number 21, Carbondale. Humphrey, J. D. and C. R. FelTing 1994 Stable Isotopic Evidence for Latest Pleistocene and Holocene Climatic Change in North-Central Texas. Quaternary Research 41:200-213. Hurst, C.T. 1941 A Folsom Location in the San Luis Valley, Colorado, A preliminary report. 7(2):31-34. 1943

A Folsom Site in a Mountain Valley of Colorado.

Southwestern Lore

American Antiquity 8(3):250-253.

Ingbar, E. E. 1992 The Hanson Site and Folsom on the Northwestern Plains. In Ice Age Hunters of the Rockies, edited by D. J. Stanford and J. S. Day, pp. 169-192. University Press of Colorado, Niwot. 1994

Lithic Material Selection and Technological Organization. In The Organization ofAinerican Prehistoric Chipped Stone Tool Technologies, edited by P. J. Carr, pp. 45-56. International Monographs in Prehistory, Archaeological Series 7, Ann Arbor.

Irwin, H. 1968 The Itama: Early Late-Pleistocene Inhabitants of the Plains of the United States and Canada and the American Southwest. Unpublished Ph.D. dissertation, Harvard University, Cambridge. Irwin-Williams, C. and C. V. Haynes 1970 Climatic Change and Early Population Dynamics in the Southwestern United States. Research 1:59-71.

Quaternary

Irwin-Williams, C., H. J. Irwin, G. A. Agogino, and C. V. Haynes, Jr. 1973 Hell Gap: Paleo-Indian Occupation on the High Plains. Plains Anthropologist 18:40-53. Issac, B. L. 1995 Introduction: Hunting and Gathering: Gender as an Economic Factor. In Research in Economic Anthropology, Volume 16, edited by B. L. Issac, pp. 1-12. JAI Press, Greenwich, Connecticut. Janes, Robert, R. 1983 Archaeological Ethnography Among MacKenzie Dene, Canada. America, Technical Paper Number 28, University of Calgary. Jochim, M. A. 1976 Hunter-Gatherer Subsistence and Settlement, A Predictive Model.

The Arctic Institute of North

Academic Press, New York.

1981 Strategies for Survival. Academic Press, New York. 1998 A Hunter-Gatherer Landscape, Southwest Germany in the Late Paleolithic and Mesolithic. Plenum Press, New York.

376

Jodry, M. A. 1987 Stewart’s Cattle Guard Site: A Folsom Site in Southern Colorado, A Report of the 1981 and 1983 Field Seasons. Unpublished Master’s Thesis, 247 p., Department of Anthropology, University of Texas, Austin. 1992 Fitting together Folsom: Refitted lithics and Site Formation Processes at Stewart’s Cattle Guard Site. In Piecing Together the Past: Applications of Refitting Studies in Archaeology, edited by J. L. Hofman and J. G. Enloe, pp. 179-209. British Archaeological Reports, International Series 578, Oxford. 1995 Archaeology of Stewart’s Cattle Guard: Report of the 1995 Field Season. Report submitted to Dr. Anna Behrensmeyer, Department of Paleobiology and Dr. David Pawson Acting Associate Director for Science, National Museum of Natural History. On file Paleoindian/Paleoecology Program, Department of Anthropology, Smithsonian Institution, Washington, D. C. 1998 The Possible Design of Folsom Ultrathin Bifaces as Fillet Knives for Jerky Production. Research in the Pleistocene 15:75-77.

Current

1999 Paleoindian Stage Paleoecological Records. In Colorado Prehistory: A Context for the Rio Grande Basin, edited by M.A. Martorano, T. Hoefer, M. A. Jodry, V. Spero, and M. L. Taylor. Colorado Council of Professional Archaeologists, In press. Jodry, M. A., D. S. Shafer, D. J. Stanford, and 0. K. Davis 1989 Late Quaternary Environments and Human Adaptation in the San Luis Valley, South Central Colorado. In Water in the Valley, A 1989 Perspective on Water Supplies, Issues, & Solutions in the San Luis Valley, Colorado, edited by E.J. Harmon, pp. 189-208. Colorado Ground-Water Association, Denver. Jodty, M. A. and D. J. Stanford 1992 Stewart’s Cattle Guard Site: An Analysis of Bison Remains in a Folsom Kill-Butchery Campsite. In Ice Age Hunters of the Rockies, edited by D. J. Stanford and J. S. Day, pp. 10 1-168. University Press of Colorado, Niwot. 1996 Changing Hydrologic Regimes and Prehistoric Landscape Use in the Northern San Luis Valley, Colorado. In Geologic Excursions to the Rocky Mountains and Beyond, Field Trip Guidebook for the 1996 Annual Meeting, Geological Society ofAinerica, Denver, Colorado October 28-31, Special Publication 44, CD-ROM, edited by R. A. Thompson, M. R. Hudson, and Charles L. Pillmore. Colorado Geological Survey, Denver. Jodry, M. A., M. D. Turner, V. Spero, J. C. Turner, and D. J. Stanford 1996 Folsom in the Colorado High Country. Current Research in the Pleistocene 13:25-27. Johnson, E.

1977 Vertebrate Remains. In Lubbock Lake - Late Quaternary studies on the Southern High Plains, edited by E. Johnson, pp. 49-89. Texas A & M Press, College Station. Johnson, E. (editor) 1987 Lubbock Lake, Late Quaternary Studies on the Southern High Plains. Press, College Station.

Texas A&M University

Johnston, A. 1969 Man’s Utilization of the Flora of the Northwest Plains. In Post-Pleistocene Man and His Environment on the Northern Plains, edited by R. G. Forbis, L. B. Davis, 0. A. Christensen, and

377

G. Fedirchuk, pp. 108-177. Chacmool Conference, Archaeological Association, Department of Archaeology, The University of Calgary. Judge, W. J. 1973 Paleoindian Occupation of the Central Rio Grande Valley in New Mexico. Mexico Press, Albuquerque.

University of New

Judge, W. J. and J. Dawson 1972 Paleolndian Settlement Technology in New Mexico. Science 176:1210-1216. Kay, M. 1996a Preliminary Descriptions of Use Wear Traces on Tools From Stewart’s Cattle Guard Site. Correspondence and Notes on File Regarding High-Magnification Analysis, Department of Anthropology, PaleoindianlPaleoecology Program, Smithsonian Institution. Washington, DC. 1996b Microwear Analysis of Some Clovis and Experimental Stone Tools. In Stone Tools Theoretical Insights into Human Prehistory, edited by G. H. Odell, pp. 315-344, Plenum Press, New York. 1998 A Use-Wear Catalog for Stewart’s Cattle Guard. Preliminary Technical Report Submitted to Margaret A. Jodry, Department of Anthropology, Paleoindian/Paleoecology Program, Smithsonian Institution, Washington D.C. Keckler, D. 1997 Surfer for Windows, Version 6, Contouring and 3D surface Mapping. Golden,Colorado.

Golden Software, Inc.,

Kehoe, T. F. 1973 The Gull Lake Site.’ A Prehistoric Bison Drive Site in Southwestern Saskatchewan. Public Museum Publications in Anthropology and History Number 1, Milwaukee.

Milwaukee

Kelley, K. B. and H. Francis 1994 Navajo Sacred Places. Indiana University Press, Bloomington. Kelly, R. L. 1983 Hunter-Gatherer Mobility Strategies. Journal ofAnthropologiccil Research 39:277-306. 1985 Hunter-Gatherer Mobility and Sedentism: A Great Basin Study. Unpublished Ph.D. dissertation, University of Michigan, Ann Arbor. University Microfilms, Am’i Arbor. 1988 The Three Sides ofaBiface. American Antiquity 53:717-734. 1994 Some Thoughts on Future Directions in the Study of Stone Tool Technological Organization. In The Organization ofAmerican Prehistoric Chipped Stone Tool Technologies, edited by P. J. Can, pp. 132-136. International Monographs in Prehistory, Archaeological Series 7, Ann Arbor. 1995 The Foraging Spectrum: Diversity in Hunter-Gatherer Lfeways. Smithsonian Institution Press, Washington. Kelly, R. L. and L. C. Todd 1988 Coming Into the Country: Early Paleoindian Hunting and Mobility. American Antiquity 53:231244. Kennedy, J., V. MacMillan. And J. Durr 1998 The Barger Gulch Troublesome Procurement Site, 5GA 195. In Early Prehistory of Middle Park:

378

The 1997 Project and Summary of Paleoindian Archaeology, edited by M. Kornfeld, pp. 34-48. Technical Report No. 15a. Department of Anthropology, University of Wyoming, Laramie. Prepared for the Colorado State Historic Preservation Fund. Kennedy M. S. (editor) 1961 The Assiniboines, From the Accounts of the Old Ones Told to First Boy (James Larpenteur Long). University of Oklahoma Press, Norman. Kessler, R. 1998 Old Spanish Trail North Branch and Its Travelers. Sunstone Press, Santa Fe. Kidder, A. V. 1932 The Artifacts of Pecos. Papers of the Southwestern Expedition 6. Yale University Press, New Haven.

Phillips Academy, Andover.

Kind, C-J. 1985 Die Verteilung von Steinartefaken in Grabungsjlachen: ein Modell zur Organisation A ltund Mittelsteinzeitlicher Siedlungsplaze. Tubingen: Archaeologica Venatoria 7. Kintigh, K. W. 1993 Tools for Quantitative Archaeology, Programs for Quantitative Analysis in Archaeology. Available from Keith Kintigh, Department of Anthropology. Arizona State University, Tempe, Arizona. Kintigh, K. W. and A. Ammerman 1982 Heuristic Approaches to Spatial Analysis in Archaeology.

American Antiquity 47:31-63.

Kluth, C. F. and C. H. Schaftenaar 1994 Depth and geometry of the northern Rio Grande Rift in the San Luis Basin, south-central Colorado. In Basins of the Rio Grande Rift. , Structure, Stratigraphy, and Tectonic Setting, edited by G. R. Keller and S. M. Cather, pp. 27-37. Geological Society of America Special Paper 291, Boulder. Knapp, A. K., J. M. Blair, J. M. Briggs, S. L. Collins, D. C. Harnett, L. C. Johnson, and B. G. Towne 1999 The Keystone Role of Bison in North American Tallgrass Prairie. BioScience 49: 3 9-50 Knudson, R. 1971 Site Inventory Record, Site No. 5EP1. Office of Archaeology and Historic Preservation, Colorado Historical Society, Denver. Kornfeld, M. (editor) 1992 Early Prehistory of Middle Park: The 1997 Project and Summary of Paleoindian Archaeology. Technical Report No. 15a. Department of Anthropology, University of Wyoming, Laramie. Prepared for the Colorado State Historic Preservation Fund. Kraft, K. C. 1997 The Distribution of Alibates Silicified Dolomite Clasts Along the Canadian River. Research in the Pleistocene 14:106-107.

Current

Kraft, K. C. and W. K. Lail 1998 Paleoindian Tool-stone Utilization in Eastern Oklahoma: An Argument for Limited Mobility. Current Research in the Pleistocene 15:27-32.

379

Kreutzer, L. A. 1992 Bison and Deer Bone Mineral Densities: Comparisons and Implications for the Interpretation of Archaeological Faunas. Journal of Archaeological Science 1992:271-294. Kroll, E. M. and T. D. Price (editors) 1991 The Interpretation of Archaeological Spatial Patterning. Plenum Press, New York. Kutzbach, J. E. 1987 Model Simulations of the Climatic Patterns during the Deglaciation of North America. In The Geology of North America, edited by W. F. Ruddiman and H. E. Wright, Jr., pp. 425-446. North America and Adjacent Oceans During the Last Deglaciation, vol. K-3, A. R. Palmer, general editor, The Geological Society of America, Boulder. La Barre, Weston (Supervisor) 1935 Papers of Weston LaBarre, Typescript of Students Notes. National Anthropological Archives, Smithsonian Institution. (Contains daily notes of all members of the 1935 field-training party of the Santa Fe Laboratory of Anthropology led by A. Lesser of Columbia University). Largent, F. B. Jr., M. R. Waters, and D. L. Carlson 1991 The Spatiotemporal Distribution and Characteristics of Folsom Projectile Points in Texas. Anthropologist 36:323-341.

Plains

Larson, M. L. 1994 Toward a Holistic Analysis of Chipped Stone Assemblages. In The Organization ofAmericon Prehistoric Chipped Stone Tool Technologies, edited by P. J. Can, pp. 5 7-69. International Monographs in Prehistory, Archaeological Series 7, Aim Arbor. Lee, R. B. 1979 The !Kung San: Men, Women, and Work in a Foraging Society. Cambridge University Press, Cambridge. 1978

The !Kung San. Cambridge University Press, Cambridge.

Leechman, D. 1951 Bone Grease. American Antiquity 16:355-356. LeTourneau, P. D. 1998 Folsom Use of Eastern New Mexico Look-alike Cherts on the Llano Estacado. in the Pleistocene 77-79.

Current Research

LeTourneau, P. D. and M. S. Shackley, and T. Baker 1998 Folsom Period Obsidian Use in New Mexico. Abstracts of the 63" Annual Meeting of the Society for American Archaeology, Seattle, Washington, March 25-29, 1998. LeTourneau, P. D., M. S. Shackley, J. M. Warnica, and J. Cummings 1996 Analysis of Obsidian Folsom Artifacts from New Mexcio. Current Research in the Pleistocene 13:59-61. Lehman, S. J. and L. D. Keigwin 1992 Sudden Changes in North Atlantic Circulation during the Last Deglaciation. Lightfoot, R. R. 1994 The Duckfoot Site, Volume II: Archaeology of the House and Household. 4, Crow Canyon Archaeological Center, Cortex, Colorado.

Nature 356:757-762.

Occasional Paper No.

WN

Lintz, C. 1991 Texas Panhandle-Pueblo Interactions from the Thirteenth through the Sixteenth Century. in, Farmers, Hunters, and Colonists, edited by K. A. Spielrnann, pp. 89-106. The University of Arizona Press, Tucson. Loring, S. 1997 On the Trail to the Caribou House: Some Reflections on Innu Caribou Hunters in Northern Ntessinan (Labrador). In Caribou and Reindeer Hunters of the Northern Hemispere, edited by L. Jackson and P. T. Thacker. pp. 185-220. Worldwide Archaeology Series 6. Avebury, Aldershot. Lowie, R. H. 1922 The Material Culture of the Crow Indians. Anthropological Papers of the American Museum of Natural History Volume XXI, Part III. New York. Lyman, R. L. 1984 Bone Density and Differential Survivorship of Fossil Classes. Archaeology 3:259-299. 1994

Journal ofAnthropological

Vertebrate Taphonomy. Cambridge University Press, Cambridge.

MacDonald, D. H. 1998a Subsistence, Tool-Use, and Reproductive Strategies of Northern Plains Folsom Hunter-Gatherers: A View From the Bobtail Wolf Site, North Dakota. Unpublished Ph.D. dissertation, Washington State University, Pullman. University Microfilms, Ann Arbor. 1998b Subsistence, Sex, and Cultural Transmission in Folsom Culture. Archaeology 17:217-239.

Journal ofAnthropological

1999 Modeling Folsom Mobility, Mating Strategies, and Technological Organization in the Northern Plains. Plains Anthropologist 44:141-161. MacDonald, D. H. and B. S. Hewlitt 1999 Reproductive Interests and Forager Mobility.

Current Anthropology 40: in press.

MacDonald, G. F. 1969 Debert, A Paieo-Indian Site in Central Nova Scotia. National Museums of Canada, Anthropology Papers Number 16. The Queen’s Printer, Ottawa Madole, R. F. 1986 Lake Devlin and Pinedale Glacial History, Front Range, Colorado. 54.

Quaternary Research 25:43-

Mallouf, R. J. 1989 A Clovis Quarry Workshop. Plains Anthropologist 34:81-103. Manabe, S. and R. J. Stouffer 1995 Simulation of Abrupt Climate Change Induced by Freshwater Input to the North Atlantic Ocean. Nature 378:165-67. Markgraf, V. and L. Scott 1981 Lower Timberline in Central Colorado during the Past 15,000 Years. Geology 9:231-234). Marriot, Alice 1935-36 Manufacture. Typed manuscript excerpted from field notes collected 1935-36. Contains

381

transcript of two older Kiowa women discussing manufacture. Western History Collection, University of Oklahoma. Marshall, L. 1960 !Kung Bushman Bands. Africa 29:335-365. Mason, 0. T. 1891 Aboriginal Skin-Dressing, A Study Based on Material in the U. S. National Museum. Report of the National Museum For the Year 1889. Smithsonian Institution. Government Printing Office, Washington D.C. Mautner, K. 1978 The Role of Koyukon Athabascan Women in Subsistence. In Tracks in the Wildiand, A Portrayal of Koyukon and Nunainiut Subsistence, edited by R. K. Nelson, K. H. Mautner, and G. R. Bane, pp. 135-207. Anthropology and Historic Preservation, Cooperative Park Studies Unit, Occasional Paper Number 9, University of Alaska, Fairbanks. Mayewski, P., L. Meeker, S. Whitlow, M. Twickler, M. Morrison, R. Alley, P. Bloomfield, and K. Taylor 1993 The Atmosphere During the Younger Dryas. Science 261:195-197. Mayle, F. E., A. J. Levesque, and L. C. Cwynar 1993 Accelerator-mass Spectrometer Ages for the Younger Diyas Event in Atlantic Canada. Quaternary Research 39:355-360. McCalpin, S. P. 1983 Quaternary Geology and Neotectonics of the West Flank of the Northern Sangre de Cristo Mountains, South-central Colorado. Colorado School of Mines Quarterly 77(3):1-89. 1996 General Geology of the Northern San Luis Valley, Colorado. In Geologic Excursions to the Rocky Mountains and Beyond, Field Trip Guidebook for the 1996 Annual Meeting, Geological Society of America, Denver, Colorado October 28-31, Special Publication 44, CD-ROM, edited by R. A. Thompson, M. R. Hudson, and Charles L. Pillmore. Colorado Geological Survey, Denver. McCartney, P. H. 1990 Alternative Hunting Strategies in Plains Adaptations. In Hunters of the Recent Past, edited by L. B. Davis and B. 0. Reeves, pp. 111-121. Unwin Hyman, London. McGowan, I. R. and D. Plazak 1996 Water Level Changes in the Unconfined Aquifer of the San Luis Valley, 1980-1995. In Geologic Excursions to the Rocky Mountains and Beyond, Field Trip Guidebookfor the 1996 Annual Meeting, Geological Society ofAmerica, Denver, Colorado October 28-31, Special Publication 44, CD-ROM, edited by R. A. Thompson, M. R. Hudson, and Charles L. Pillmore. Colorado Geological Survey, Denver. McNaughton, S. J 1976 Serengeti Migratory Wildebeest: Facilitation of Energy Flow by Grazing. Science 191:92-94. 1979 Grassland Herbivore Dynamics. In Serengeti: Dynamics of an Ecosystem, edited by A. R. E. Sinclair and M. Norton-Griffiths, pp. 46-81. University of Chicago Press, Chicago. McNaughton, S. J., M. B. Coughenour, and L. L. Wallace 1982 Interactive Processes in Grassland Ecosystems. In Grasses and Grasslands, edited by S. R. Estes, R. J. Tyrl, and S. N. Brunken, pp. 167-193. University of Oklahoma Press, Norman.

382

McNaughton, S. J., F. F. Banyikwa, and M. M. McNaughton 1997 Promotion of the Cycling of Diet-Enhancing Nutrients by African Grazers. 1800. 1998 Root Biomass and Productivity in a Grazing Ecosystem: The Serengeti. McNaughton, S. J., R. W. Ruess, and S. W. Seagle 1988 Large Mammals and Process Dynamics in African Ecosystems.

Science 278:1798-

Ecology 79:587-592.

BioScience 38:794-800.

McNaughton, S. J., L. L. Wallace, and M. B. Coughenour 1983 Plant Adaptation in an Ecosystem Context: Effects of Defoliation, Nitrogen, and Water on Growth of an African Sedge. Ecology 64:307-318. Mead, J. I. and D. J. Meltzer 1984 North American Late Quaternary Extinctions and the Radiocarbon Record. In Quaternary Extinctions: A Prehistoric Revolution, edited by P. S. Martin and R. G. Klein, pp. 440-450. University of Arizona Press, Tucson. Meagher, M. M.

1973 The Bison of Yellowstone Park.

National Park Service Monograph Series 1.

Melson, B. 1996 Personal Communication, Department of Minerology, National Museum of NaturaL History, Smithsonian Institution, Washington, DC. Meltzer, D. J. 1989 Was Stone Exchanged Among Eastern North American Paleoindians? In Eastern Paleoindian Lithic Resource Use, edited by C. J. Ellis and J. C Lothrop, pp. 11-39. Westview Press, Boulder. Menounos, B. and M. A. Reasoner 1997 Evidence for Cirque Glaciation in the Colorado Front Range during the Younger Dryas Chronozone. Quaternary Research 48:38-47. Merrill, W. L., M. K. Hansson, C. S. Greene, and F. J. Reuss 1997 A Guide to the Kiowa Collections at the Smithsonian Institution. Washington D.C.

Smithsonian Institution Press,

Metcalf, G. 1970 Some Wooden Scraper Handles From the Great Plains and the Southwest. Plains Anthropologist

15:46-53. Metcalf, M. D., R. J. Rood, P. K. O’Brien, B. R. Overturf 1991 Kreinrnling Chert Procurement in the Middle Park Area, Colorado: 5GA 1144 and 5GA 1172. Prepared by Metcalf Archaeological Consultants, Inc. Eagle, Colorado. Prepared for J. F. Sato and Associates, Golden, Colorado. Miller, S. J. and W. Dort, Jr. 1978 Early Man at Owl Cave: Current Investigations at the Wasden Site, Eastern Snake River Plain, Idaho. In Early Man in America From a Circum-Pacific Perspective, edited by A. L. Bryan, pp. 129-139. Occasional Papers No. 1, Department of Anthropology, University of Alberta. Mithen, S. J. 1990 Thoughtful Foragers, A Study of Prehistoric Decision Making. Cambridge.

Cambridge University Press,

383

Montet-White, A. and S. Helen (editors) 1991 Raw Material Economies Among Hunter Gatherers. University of Kansas, Lawrence.

Publications in Anthropology Volume 19.

Morrow, T. A. and J. E. Morrow 1999 On the Fringe: Folsom Points and Preforms in Iowa. In Folsom Lithic Technology: Explorations in Structure and Variation, edited by D. S. Amick, pp. 65-81. International Monographs in Prehistory, Archaeological Series 12, Ann Arbor. Muehiberger, W. R. 1967 Geology of the Chama Quadrangle, New Mexico and the Chama Valley Survey. New Mexico Anthropology Series Bulletin 2, No. 1, Albuquerque.

University of

Munson, P. S. 1990 Folsom Fluted Projectile Points East of the Great Plains and their Biogeographical Correlates. North American Archaeologist 11:255-272. Myers, T. P. 1995 Paleoindians of the Eastern Sand Hills. Plains Anthropologist 40:61-68. Naranjo, Aldan 1998 Personal Communication, Historian, Southern Ute Tribe, Ignacio, Colorado. Videotaped in Ignacio, October 1998 by Pegi Jodry and Ted Tinirek. Department of Anthropology, PaleoindianlPaleoecology Program, Smithsonian Institution, Washington, D.C. Naze, B. 1986 The Folsom Occupation of Middle Park, Colorado. Southwestern Lore 52(4): 1-32. 1994 The Crying Woman Site: A Record of Prehistoric Human Habitation in the Colorado Rockies. Unpublished Master’s thesis, Colorado State University, Fort Collins. Nelson, R. 1978 Cultural and Ecological Dimensions of Subsistence. In Tracks in the Wildland: A Portrayal of Koyukon and Nunamiut Subsistence, by R. K. Nelson, K. H. Mautner, and G. R. Bane, pp. 218353. Occasional Paper No. 9, Anthropology and Historic Preservation Cooperative Park Studies Unit, University of Alaska, Fairbanks. Nelson, M. C. 1991 The Study of Technological Organization. Archaeological Method and Theory 3:57-95. University of Arizona Press, Tucson. Newman, J. R. 1993 The Effects of Distance on Lithic Material Reduction Technology. 21:491-501.

Journal ofField Archaeology

O’Connell, J. F. and K. Hawkes 1981 Alyawara Plant Use and Optimal Foraging Theory. In Hunter-Gatherer Foraging Strategies, Odell, G. H. 1996 Stone Tools and Mobility in the Illinois River Valley. Archaeological Series 10, Ann Arbor.

International Monographs in Prehistory,

384

Odess, D. 1996 Interaction, Adaptation and Culture Change: Lithic Exchange in Frobisher Bay Dorset Society, Baffin Island, Arctic Canada. Unpublished Ph.D. Dissertation, Department of Anthropology, Brown University, Providence. Olterman, J. (editor) 1995 San Luis Valley Waterbird Plan. Report jointly prepared by the Colorado Division of Wildlife, United States Fish and Wildlife Service, and U. S. Bureau of Land Management, Alamosa, Colorado. Available from the Colorado Division of Wildlife, Monte Vista, Colorado. Opler, M. K. 1940 The Southern Ute of Colorado. In Acculturation in Seven American Indian Tribes, Linton, pp. 119-203 D. Appleton-Centruy Company, New York.

edited by R.

Pannell, J. P., J. M. Yenter, S. 0. Woodyard, and R. E. Mayhugh 1973 Soil Survey of Alamosa Area, Colorado. Soil Conservation Service, In Cooperation with the Colorado Agricultural Experiment Station. United States Department of Agriculture, Washington D. C. Patten, B. 1999 Old Tools-New Eyes, A Primal Primer of Flintknapping.

Stone Dagger Publications, Denver.

Patterson, R. T., J. P. Guilbault, R. E. Thomson, and L. L. Luternauer 1995 Foraminiferal Evidence of Younger Diyas Age Cooling on the British Columbia shelf. Geographic physique et Quaternaire 49:409-427. Patrick, R. 1938 The Occurrence of Flints and Extinct Animals in Pluvial Deposits Near Clovis, New Mexico, Part V- Diatom Evidence From the Mammoth Pit. Proceedings of the Philadelphia Academy of Natural Sciences 90:15-25. Peden, D. G., G. M. Van Dyne, R. W. Rice, and R. M. Hansen 1974 The Trophic Ecology of Bison bison L. on Shortgrass Plains. Journal ofApplied Ecology 11:489498. People of Ksan, 1980 Gathering What the Great Nature Provided. University of Washington Press, Seattle. Peteet, D. M. 1995 Global Younger Dryas? Quaternary International 8:93-104. Peteet, D. M., and D. H. Mann 1994 Late-glacial Vegetational, tephra, and Climatic History of Southwest Kodiak Island, Alaska. Ecoscience 1:255-267. Peteet, R. K. 1978 Latitudinal Variation in Southern Rocky Mountain Forests. Journal of Biogeography 5:275-289. Peters, V. B. 1995 Women of the Earth Lodges, Tribal Life on the Plains. Archon Books, North Haven, Connecticut. Pettit, J. 1990 Utes, The Mountain People. Johnson Books, Boulder.

385

Quade, J., R. M. Forester, W. L. Pratt, and C. Carter 1998 Black Mats, Spring-fed Streams, and Late-Glacial-Age Recharge in the Southern Great Basin. Quaternary Research 49(2):129-148. Quigg, J. M. 1997 Bison Processing at the Rush Site, 41TG346, and Evidence for Pemmican Production in the Southern Plains. Plains Anthropologist 42:145-161. Rapson, D.J. 1990 Pattern and Process in Intra-Site Spatial Analysis, Site Structural and Faunal Research at the Bugas -Holding Site. Unpublished Ph.D. dissertation, Department of Anthropology, University of New Mexico, Albuquerque. Ray, C. N.

1947 Chemical Alteration of Silicate Artifacts. Bulletin of the Texas Archeological and Paleontological Society 18:28-39. Reasoner, M. A. and M. A. Jodry 1998a Response of the Alpine Timberline Ecotone to the Younger-Dryas Climate Oscillation in the Colorado Rocky Mountains: Evidence from Two Well-Dated, High Resolution Pollen Records. Poster presented at 15th Biennial Meeting of the American Quaternary Association, Puerto Vallarta. 1998b Postglacial Vegetation History of Black Mountain Lake. Preliminary report submitted to the Department of Anthropology, Smithsonian Institution and to the Colorado Natural History Small Grants Program. Copies available from the PaleoindianlPaleoecology Program, Smithsonian Institution. 1999 Rapid Response of Alpine Timberline Vegetation to the Younger Dryas Climate Oscillation in the Colorado Rocky mountains, USA. Geology, in press. Reasoner, M. A., G, Osborn, and N. W. Rutter 1993 Age of the Crowfoot Advance in the Canadian Rocky Mountains: A Glacial Event Coeval with the Younger Dryas Oscillation. Geology 22(5):439-442. Redder, A. J. 1985 Horn Shelter Number 2: The South End, a Preliminary Report. Central Texas Archeologist 10:37-65. Reher, C. A. 1974 Population Study of the Casper Site Bison. In The Casper Site, A Hell Gap Bison Kill on the High Plains, edited by G. C. Frison, pp. 113-124. Academic Press, New York. 1977 Adaptive Process on the Shortgrass Plains. In For Theory Building in Archaeology, edited by L. R. Binford, pp. 13-40. Academic Press, New York. 1978 Buffalo Population and Other Deterministic Factors in a Model of Adaptive Process on the Shortgrass Plains. Bison Procurement and Utilization: A Symposium, edited by L. B. Davis and M. Wilson, pp. 23-39. Memoir 14, Plains Anthropologist, Lincoln. 1991 Large Scale Lithic Quarries and Regional Transport Systems on the High Plains of Eastern

Wyoming, Spanish Diggings Revisited. In Raw Material Economies Among Prehistoric HunterGatherers, edited by A. Montet-White and S. Holen, pp. 251-284. University of Kansas Publications in Anthropology 19.

386

Reber, C. A. and G. C. Frison 1980 The Vore Site, 48CK302, A Stratified Buffalo Jump in the Wyoming Black Hills. Plains Anthropologist, Lincoln.

Memoir 16,

Reider, R. G. 1990 Late Pleistocene and Holocene Pedogenic and Environmental Trends at Archaeological Sites in Plains and Mountain Areas of Colorado and Wyoming. In Archaeological Geology of North America, edited by N. P, Lasca and J. Donohue, pp.335-359. Geological Society of America, Centennial Special Volume 4, Boulder. Reynolds, M. 1999 Personal Communication, Colorado State University, Co-operative Extension Service, La Jara, Colorado. Reynolds, H. W., R. M. Hansen, D. G. Peden 1978 Diets of the Slave River Lowland Bison Herd, Northwest Territories, Canada. Management 42:581-590.

Journal of Wildlife

Reynolds, M. 1999 Colorado State University, Co-operative Extension Service, Alamosa, Colorado. Richards, M. 1997 Deerskins into Buckskins How to Tan with Natural Materials, A Field Guide for Hunters and Gatherers. Backcountry Publishing, Cave Junction, Oregon. Richmond, P. J. 1990 Trail to Disaster. University Press of Colorado, Niwot. Riggs, J.

1980 Blue Mountain Buckskin A Working Manual, Dry-Scrape Brain-Tan.

Second edition. Available from Jim Riggs, 72501 Highway 82, Wallowa, Oregon 97885, Sixth Printing 1995.

Roberts, F. H. H. 1935 A Folsom Complex: Preliminary Report on Investigations at the Lindenmeier Site in Northern Colorado. Smithsonian Miscellaneous Collections 94(4). Roberts, M. F. 1970 Late Glacial and Postglacial Environments in Southeastern Wyoming. Palaeoclimatolo,gy, and Palaeoecology 8:5-17.

Palaeogeography,

Rogers, E. S. 1969 Band Organization Among the Indians of Eastern Subarctic Canada. In Contributions to Anthropology: Band Societies, edited by David Damas, pp. 21-52. National Museums of Canada Bulletin 228. Root, M. J. 1993 Analyses of Stone Tools and Flaking Debris. In Site 32DU955A: Folsom Occupation of the KnJ River Flint Primary Source Area, Phase III (Part 1) Archaeological Data Recovery at Lake lb National Wildlife Refuge, Dunn County, North Dakota: Interim Report For 1992-1993 Investigations at 32DU955A, edited by M. J. Root, pp. 179-227. Project Report Number 22, Center for Northwest Anthropology, Department of Anthropology, Washington State University, Pullman.

387

1994a Analysis of Stone Tools and Flaking Debris. In Archaeology of the Bobtail Wolf site (32DU955A): 1993-94 Progress Report, by M. J Root, and A. M. Emerson, pp. 147-190. Project Report Number 26, Center for Northwest Anthropology, Department of Anthropology, Washington State University, Pullman. 1994b Archaeological Field and Laboratory Techniques. In Archaeology of the Bobtail Wolf site (32DU955A): 1993-94 Progress Report, by M. J Root. and A. M. Emerson, pp. 23-48. Project Report Number 26, Center for Northwest Anthropology, Department of Anthropology, Washington State University, Pullman. 1994c Folsom Lithic Technology and Raw Material Availability at the Bobtail Wolf Site (32DU955A), A Lithic Workshop and Camp in the Knife River Flint Quarry Area. Paper Presented at the 52 nd Annual Meeting of the Plains Anthropological Society, Lubbock. Root, M. Sand A. M. Emerson 1994 Archaeology of the Bobtail Wolf site (32DU955A) 1993-94 Progress Report. Project Report Number 26, Center for Northwest Anthropology, Department of Anthropology, Washington State University, Pullman. Root, M. J., J. D. William, M. Kay, and L. K. Shiftmn 1999 Folsom Ultrathin Biface and Radial Break Tools in the Knife River Flint Quarry Area. In Folsom Lithic Technology, edited by D. S. Amick, pp. 144-168. International Monographs in Prehistory, Archaeological Series 12, Ann Arbor. Rowe, P. R. 1952 Early Horizons of Mills County, Iowa: Part I, Evidences of Early Man. Io,’a Archaeological Society Journal 1(3):6-13. Sampson, C. G. 1988 Stylistic Boundaries Among Mobile Hunter-Foragers. D.C.

Smithsonian Institution Press. Washington

Savishinsky, S. S. 1994 The Trail of the Hare, Environment and Stress in a Sub-Arctic Community, Gordon and Breach, USA. Schaeffer, J. B. 1958 The Alibates Flint Quarry, Texas.

second edition.

American Antiquity 24:189-191.

Schneider, F. E. 1982 A Preliminary Investigation of Paleo-Indian Cultures in North Dakota. Quarterly 6(4):16-43. Sellards, E. H. 1952 Early Man in America: A Study of Prehistory.

Manitoba Archaeological

The University of Texas Press, Austin.

Schneider, F. F. 1982 A Preliminary Investigation of Paleo-Indian Cultures in North Dakota. Quarterly 6(4):16-43). Schultz, C. B. 1932 Association of Artifacts and Extinct Mammals in Nebraska. Museum 1(33):271-282.

Manitoba Archaeological

Bulletin of the Nebraska State

388

1943 Some Artifacts of Early Man in the Great Plains and Adjacent Areas, 249. Schultz, J. M. 1992 The Use-Wear Generated by Processing Hides,

American Antiquity 8:242-

Plains Anthropologist 37:333-351.

Seeman, M. F. 1994 Interciuster Lithic Patterning at Nobles Pond: A Case For ’Disembedded Procurement Among Early Paleoindian Societies. American Antiquity 59:273-288. Shafer, D. S. 1989 The Timing of Late Quaternary Monsoon Precipitation Maxima in the Southwest United States. Unpublished Ph.D. dissertation, University of Arizona, Tucson. University Microfilms, Ann Arbor. Shafer, H. J. 1970 Notes on Uniface Retouch Technology. American Antiquity 35:480-487. 1983 The Tranchet Technique in Lowland Maya Lithic Tool Production.

Lithic Technology 12:57-56.

Shariff, A. R., M. E. Biondini, and C. E. Grygiel 1994 Grazing Intensity Effects on Litter Decomposition and Soil Nitrogen Mineralization. Range Management 47:444-449. Shifrmn, L. K. and J. D. William 1996 Investigations at Young-Man-Chief, a Folsom Site in Western North Dakota. the Pleistocene 13:42-44.

Journal of

Current Research in

Shott, M. J. 1986a Settlement Mobility and Technological Organization Among Great Lakes Paleo-Indian Foragers. Unpublished Ph.D. dissertation, University of Michigan, Ann Arbor. University Microfilms, Ann Arbor. 1986b Technological Organization and Settlement Mobility: An Ethnographic Examination. Anthropological Research 42:15-51.

Journal of

1989 Technological Organization in Great Lakes Paleoindian Assemblages. In Eastern Paleoindian Lithic Resource Use, edited by C. J. Ellis and J. C. Lothrop, pp. 221-237. Westview Press, Boulder. Siegel, P. E. 1977 Functional Variability Within an Assemblage of Endscrapers.

Lithic Technology 13:35-51.

Silberbauer, G. 1981 Hunter and Habitat in the Central Kalahari Desert. Cambridge University Press, Cambridge. Smith, C. S. and L. M. McNees 1990 Rattlesnake Pass Site: A Folsom Occupation in South-Central Wyoming, Plains Anthropologist 35:273-289. Smith, E. C. and R. Boyd 1990 Risk and Reciprocity: Hunter-Gatherer Socioecology and the Problem of Collective Action. In Risk and Uncertainty in Tribal and Peasant Economies, edited by E, Cashadan, pp. 167-191. Westview Press, Boulder.

389

Smith, J. E., M. J. Risk, H. P. Schwarcz, and T. A. McConnaughey 1997 Rapid Climate Change in the North Atlantic during the Younger Dryas Recorded by Deep-Sea Corals. Nature 386(6627):818-820. Snow, Chief J. 1977 These Mountains are our Sacred Places The Story of the Stoney Indians. Toronto.

Samuel Stevens,

Snyder, L. 1999 Archaeozoologist and Research Collaborator, Analysis of Stewart’s Cattle Guard Faunal Assemblage, National Museum of Natural History, Smithsonian Institution, Washington, DC. Soffer, 0. 1985 The Upper Paleolithic of the Central Russian Plain. Academic Press, Orlando. Soilberger, J. B. 1971 A Technological Study of Beveled Knives. Plains Anthropologist 16:209-218. 1977 On Fluting Folsom: Notes on Recent Experiments. 48:47-52. Speth, J. D. 1983 Bison Kills and Bone Counts.

Bulletin of the Texas Archeological Society

University of Chicago Press, Chicago.

Spero, V. 1999a Sources of Lithic Material In and Near the Rio Grande Basin. In Rio Grande Basin, Prehistoric Regional Context Colorado, edited by Marilyn Martorano and Ted Hoefer. The Prehistory of Colorado: A Publication Series. Colorado Council of Professional Archaeologists and the Colorado Historical Society, Denver. In press. 1999b Rio Grande Basin Native American Perspectives. In Rio Grande Basin, Prehistoric Regional Context Colorado, edited by Marilyn Martorano and Ted Hoefer. The Prehistory of Colorado: A Publication Series. Colorado Council of Professional Archaeologists and the Colorado Historical Society, Denver. In press. Spielmann, K. A. 1983 Late Prehistoric Exchange Between the Southwest and Southern Plains. 28:257-272. 1991

Plains Anthropologist

Farmers, Hunters, and Colonists. The University of Arizona Press, Tucson.

Stafford, M. D. 1990 The Powars II site 48PL330: A Paleoindian red ochre mine in eastern Wyoming. Unpublished Master’s Thesis, Department of Anthropology, University of Wyoming, Laramie. Stanford, D. J. 1978 The Jones-Miller Site: An Example of Hell Gap Bison Procurement Strategy. In Bison Procurement and Utilization: A Symposium, edited by L. B. Davis and M. Wilson, pp. 90-97. Society for American Archaeology, Memoir 1 14. 1990 A History of Archaeological Research in the San Luis Valley, Colorado. Historian. Vol. XXII (3):33-39.

The San Luis Valley

Wif

1991 Folsom Mobility as Reflected by the Distribution of Lithic Raw Materials in Sites From Montana to Texas. Paper presented at the 56th Annual Meeting of the Society for American Archaeology, New Orleans. 1999a Paleoindian Archeology and Late Pleistocene Environments in the Plains and Southwestern United States. In Ice Age Peoples of North America, edited by R. Bonnichsen. 110 pp. Center for the Study of the First Americans, Corvallis. 1999b Personal Communication, Department of Anthropology, National Museum of Natural History, Smithsonian Institution, Washington, DC. Stanford, D. J. and F. Broilo 1981 Franks Folsom Campsite. In Archaeological Essays in Honor of Mark Wimberley, S. Foster. The Artifact 19(3):1-13, El Paso.

edited by M.

Steward, J. H. 1938 Basin-Plateau Aboriginal Sociopolitical Groups. Bureau of American Ethnology Bulletin 20, Washington D.C. 1969 Postscript to Bands: On Taxonomy, Processes, and Causes. . In Contributions to Anthropology.’ Band Societies, edited by David Damas, pp. 288-295. National Museums of Canada Bulletin 228. Stiner, M. C. 1994 Honor Among Thieves, A Zooarchaeological Study of Neanderthal Ecology. Press, Princeton. Stoffle, R. W., D. B. Halmo, J. E. Olmsted, M. J. Evans 1990 Native American Cultural Resource Studies at Yucca Mountain, Nevada. Research, University of Michigan, Ann Arbor.

Princeton University

Institute for Social

Stromberg, B. 1994 Younger Dryas Deglaciation at Mt. Billingen, and Clay Varve Dating of the Younger Dryas/Preboreal Transition. Boreas 23(2):177-193. Stockton, E. D. 1973 Shaw’s Creek Shelter: Human Displacement of Artefacts and its Significance. Mankind 9:112117. Storck, P. L. 1997 The Fisher Site, Archaeological, Geological, and Paleobotanical Studies at an Early Pa/co-Indian Site in Southern Ontario, Canada. Memoirs Museum of Anthropology, University of Michigan Number 30, Ann Arbor. Storck, P. L. and J. Tomenchuk 1990 An Early Paleoindian Cache of Informal Tools at the Udora Site, Ontario. In Early Pa/eoindian Economies of Eastern North America, edited by K. B. Tankersley and B. L. Isaac, pp. 45-93. Research in Economic Anthropology, Supplement 5. Stuart, D. E. and R. P. Gauthier 1984 Prehistoric New Mexico. New Mexico Archaeological Council, Albuquerque. Stuiver, M. and P. J. Reimer 1993 Extended C Data Base and Revised CALIB 3.0 35(l):215-230. ’4

’ 4

C Age Calibration Program. Radiocarbon

WI’

Tanner, A. 1979 Bringing Home Animals, Religious Ideology and Mode of Production of the Mistassini Cree Hunters. St. Martins Press, New York. Taylor, K., P. Mayewski, R. Alley, E. Brook, A. Gow, P. Grootes, D. Meese, E. Saltzman, J. Severinghaus, M. Twickler, J. White, S. Whitlow, and G. Zielinski 1997 The Holocene-Younger Dryas Transition Recorded at Summit, Greenland. Science 278(5339):825-827. Tieszen, L. L., L. Stretch, and J. V. Kooi 1997 Staple Isotopic Determination of Seasonal Dietary Patterns in Bison at Four Preserves Across the Great Plains. In International Symposium on Bison Ecology and Management in North America, edited by L. R. Irby and J. E. Knight, pp. 130-140. Montana State University, Bozeman. Thomas, D. H. 1983 The Archaeology of Monitor Valley, 1: Epistemology, Anthropological Papers No. 58, Pt. 1. American Museum of Natural History, New York. 1986 Contemporary Hunter-Gatherer Archaeology in America. In American Archaeology Past and Future, A Celebration of the Society for American Archaeology 1935-1985, edited by D. J. Meltzer, D. D. Fowler, and J. A. Sabloff, pp. 237-276. Smithsonian Institution Press, Washington D. C. 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 yr BP In Global Changes Since the Last Glacial Maximum, edited by H. E. Wright, J. E. Kutzbach, W. F. Ruddiman, F. A. StreetPerrott, T. Webb III, and P. J. Bartlein, pp. 468-513. University of Minnesota Press, Minneapolis. Titmus, G. L. and J. C. Woods 1991 A Closer Look at Margin Grinding’ on Folsom and Clovis Points. Journal of California and Great Basin Anthropology 13:194-203. Todd, L. C. 1987 Taphonomy of the Homer II Bonebed. In The Homer Site: The Type Site of the Cody Cultural Complex, edited by G. C. Frison and L. C. Todd, pp. 107-198. Academic Press, Orlando. 1991 Seasonality Studies and Paleoindian Subsistence Strategies. In Human Predators and Prey Mortality, edited by M. C. Stiner, pp. 217-276. Westview Press, Boulder. 1995 AP460/660 Field Class in Archaeology. 1995 Excavations at the Hudson-Meng Bison Bonebed 25SX1 15. Department of Anthropology, Colorado State University, Fort Collins. Todd, L. C. and J. L. Hofman 1987 Bison Mandibles from the Homer and Finley Sites. In The Homer Site, The Type Site of the Cody Cultural Complex, edited by G. C. Frison and L. C. Todd, pp. 493-539. Academic Press, Orlando. Todd, L. C., J. L. Hofman and C. B. Schultz 1990 Seasonality of the Scottsbluff and Lipscomb Bison Bone Beds: Implications for Modeling Paleoindian Subsistence. American Antiquity 55:813-827. 1992 Faunal Analysis and Paleoindian Studies: A Reexamination of the Lipscomb Bison Bonebed. Plains Anthropologist 37:137-165.

392

Todd, L. C., D. J. Rapson, and J. L. Hofman 1996 Dentition Studies of the Mill Iron and Other Early Paleoindian Bison Bonebed Sites. In The Mill Iron Site, edited by G. C. Frison, pp. 145-175. University of New Mexico Press, Albuquerque. Toll, 0. W. 1962 Arapaho Names and Trails. A Report of a 1914 Pack Trip.

Published by the author. 43 pp.

Tomenchuk, J. 1995 Status Report on a Use-Wear Feasibility Study of Two Samples of Chipped Stone Tools From Clusters 3 and 5 at the Folsom-Age Stewart’s Cattle Guard Site. Preliminary Technical Report Submitted to Margaret A. Jodry, Department of Anthropology, Paleoindian/Paleoecology Program, Smithsonian Institution, Washington D.C. 1997 A Parametric Use-Wear Study of Artifacts from Areas C and C-east. In the Fisher Site: Archaeological, Geological and Paleobotanical Studies at an Early Paleo-Indian Site in Southern Ontario, Canada, edited by P. Storck, pp. 95-161. Memoirs of the Museum of Anthropology, Number 30. University of Michigan, Ann Arbor. 1999 Preliminary Functional Interpretations of Chipped Stone Tools From Clusters 3 and 5 at the Folsom-Age Stewart’s Cattle Guard Site. Notes on File, Department of Anthropology, PaleoindianfPaleoecology Program, Smithsonian Institution, Washington D.C. Tomenchuk, J. and P. L. Storck 1997 Two Newly Recognized Paleoindian Tool Types: Single- and Double-Scribe Compass Gravers and Coring Gravers. American Antiquity 63:508-522. Torrence, R. 1982 Time Budgeting and Hunter-Gatherer Technology. In Hunter-Gatherer Economy in Prehistoiy: A European Perspective, edited by G. Bailey, pp. 11-22. Cambridge University Press, Cambridge. 1989a Tools as Optimal Solutions. In Time, Energy and Stone Tools, edited by R. Torrence, pp. 1-6. Cambridge University Press, Cambridge 1989b Retooling: Towards a Behavioral Theory of Stone Tools. In Time, Energy and Stone Tools, edited by R. Torrence, pp. 57-66. Cambridge University Press, Cambridge Torrence, R. (editor) 1989c Time, Energy and Stone Tools.

Cambridge University Press, Cambridge.

Tunnel], C. 1976 Fluted Projectile Point Production as Revealed by Lithic Specimens from the Adair-Steadman Site in Northwest Texas. The Museum Journal 17:140-168. 1978 The Gibson Lithic Cache from West Texas. Office of the State Archeologist Report 30. Historical Commission.

Texas

Turner, E. S. and P. Tanner 1994 The McFaddin Beach Site on the Upper Texas Coast. Bulletin of the Texas Archeological Society 65:3 19-336. Turney-High, H. H. 1937 The Flathead Indians of Montana. American Anthropological Association, Memoir 48.

393

United States Department of the Interior, Bureau of Land Management 1977 Final San Luis Research Area Grazing Management Environmental Statement. Bureau of Land Management, Department of the Interior, Washington D.C. United States Forest Service, Rio Grande National Forest 1996 Soil Resource and Ecological Inventory of Rio Grande National Forest - West Part, Colorado. Available from San Juan Rio Grande National Forest, Monte Vista, Colorado. Van Noten, F., D. Cahen, and L. Keeley 1980 A Paleolithic Campsite in Belgium. Scientific American 242(4):48-55. Vargas, D. 1694 Journal of the Vargas Expedition into Colorado, 1694. Archivo de la Nacin, Mexico City, Mexico. Copy available from Adams State College, Alamosa, Colorado. Varien, M. D. 1997 New Perspectives on Settlement Patterns: Sedentism and Mobility in a Social Landscape. Unpublished Ph.D. dissertation, Arizona State University. University Microfilms, Ann Arbor. Villa, P. 1982 Conjoinable pieces and Site Formation Processes. American Antiquity 47:276-290. Villa, P. and J. Courtin 1983 The Interpretation of Stratified Sites: A View From Underground. Journal ofArchaeological Science 10:267-281. Walde, D. and N. D. Willows, (editors) 1991 The Archaeology of Gender. Proceedings of the Twenty-Second Annual Conference of the Archaeological Association of the University of Calgary, Calgary. Walker, D. N. 1982 Early Holocene Vertebrate Fauna. In The Agate Basin Site: A Record ofPaleoindian Occupation on the Northwestern High Plains, edited by G. C. Frison and D. J. Stanford, pp. 274-308. Academic Press, New York. 1987 Late Pleistocene/Holocene Environmental Changes in Wyoming: The Mammalian Record. In Late Quaternaiy Mammalian Biogeography and Environments of the Great Plains and Prairies, edited by R. W. Graham, H. A. Semken, Jr., and M. A. Graham, pp. 334-392. Illinois State Museum Scientific Papers 22, Springfield. Warnica, J. 1961 The Elida Site: Evidence of a Folsom Occupation in Roosevelt County Eastern New Mexico. Bulletin of the Texas Archaeological Society 30:209-2 15. Warren, A. H. 1967 Petrographic Analysis of Pottery and Kithics. In An Archaeological Survey of the Chuska Volley and the Chaco Plateau New Mexico, Part 1, Natural Science Studies, by A. H. Harris, J. Schoenwetter, and A. H. Warren, pp. 104-134. Watanabe, H. 1968 Subsistence and Ecology of Northern Food Gatherers with Special Reference to the Ainu. In Man the Hunter, edited by R. Lee and I. DeVore, pp. 69-77. Aldine, Chicago.

394

Wedel. W. R. 1970 Antler Tine Scraper Handles in the Central Plains. 1986

Plains Anthropologist 15:36-45.

Central Plains Prehistory: Holocene Environments and Culture Change in the Republican River Basin. University of Nebraska Press, Lincoln.

Vv’iessner, P. W. 1977 Hxaro: A Regional System of Reciprocity for Reducing Risk Among the !Kung San Volume 1. Unpublished Ph.D. dissertation, University of Michigan, Am Arbor. University Microfilms, Ann Arbor. 1983 Style and Social Information in Kalahari San Projectile Points.

American Antiquity 48:253-276.

1986 Kung San Networks in a Generational Perspective. In The Past and Future of!Kung Ethnography: Critical Reflections and Symbolic Perspectives Essays in Honor of Lorna Marshall, edited by M. Biesele, R. Gordon, and R. Lee, pp.103-136. Helmut Buske Verlag, Hamburg. Wendorf, F., A. D. Krieger, C. C. Albritton, and T. D. Stewart 1955 The Midland Discovery, A Report on the Pleistocene Human Remains From Midland, Texas. University of Texas Press, Austin. Wendorf, F. and T. H. Thomas 1951 Early Man Sites Near Concho, Arizona. American Antiquity 17:107-14. Wenzel, G. W. 1995 Ningiqtuq: Resource Sharing and generalized reciprocity in Clyde River, Nunavut. Anthropology 32(2):43-60.

Arctic

Wessel, R. L., P. L. Eidenbach, C. S. Corner, B. Knight, and L. M. Meyer 1996 From Playas to Highlands: Paleoindian Adaptations to the Region of the Tularosa. Submitted to MEV ATEC, Corporation, White Sands Missle Range, New Mexico, Contract No. DAAD07-95C-0 125, Subcontract No. MCH-96-C-0011. Copies available from Humans Systems Research, Inc., Tularosa, New Mexico. Wheat, J. B. 1972 The Olsen-Chubbuck Site: A Paleo-Indian Bison Kill. Society for American Archaeology, Memoir 26. 1978 Olsen-Chubbuck and Jurgens Sites: Four Aspects of Paleo-Indian Economy. In Bison Procurement and Utilization: A Symposium, edited by L. B. Davis and M. Wilson, pp. 84-89. Memoir 14, Plains Anthropologist, Lincoln. 1979

The Jurgens Site. Memoir 15. Plains Anthropologist, Lincoln.

1991 The Chispa Creek Folsom Site, Culberson County, Texas. Abstracts of the 56" Annual Meeting, Society for American Archaeology, New Orleans, p. 173. Wheeler, R. 1999 Personal Communication, Bureau of Land Management, Alamosa, Colorado. Whitaker, J. 0., Jr. 1996 National Audubon Society Field Guide to North American Mammals. York.

Alfred A. Knopf, New

395

William, J. D. (editor) 1994 Site 32DU955C: A Folsom Complex Workshop in the Knife River Flint Primary Source Area. Quaternary Studies Program, Northern Arizona University, Flagstaff. Submitted to the U.S. Fish and Wildlife Service, Denver. William, J. D., M. J. Root, and L. K. Shifrin 1997 Lake Ilo Ultrathin Bifaces and Folsom Points: Separate Production Sequences. in the Pleistocene 14:111-112.

Current Research

William, J. D. and L. K. Shifrmn (editors) 1995 The Big Black Site (32DU955C): A Folsom Complex Workshop in the Knife River Flint Primary Source Area, A Progress Report for the 1994 Field Season, Quaternary Studies Program, Northern Arizona University, Flagstaff. Submitted to the U.S. Fish and Wildlife Service, Denver. Williams, M. W., P. D. Brooks, and T. Seastedt 1998 Nitrogen and Carbon Soil Dynamics in Response to Climate Change in a High-elevation Ecosystem in the Rocky Mountains, U.S.A. Arctic andAipine Research 30:26-30. Wilmsen, E. N. 1973 Interaction, Spacing Behavior, and the Organization of Hunting Bands. Journal of Anthropological Research 29:1-31. 1974 Lindenmeier: A Pleistocene Hunting Society. Harper and Row, New York. Wilmsen, E. N. and F. H. H. Roberts, Jr. 1978 Lindenmeier 1934-1974, Concluding Report on Investigations. Anthropology, Number 24. Washington D.C.

Smithsonian Contributions to

Wilson, G. L. 1987 Buffalo Bird Woman’s Garden, Agriculture of the Hidatsa Indians. Society Press, St. Paul.

Minnesota Historical

Winterhalder, B. and E. A. Smith (editors) 1981 Hunter-Gatherer Foraging Strategies. The University of Chicago Press, Chicago. Wobst, H. M. 1974 Boundary Conditions for Palaeolithic Social Systems: A Simulation Approach. American Antiquity 39:147-78. 1976 Locational Relationships in Paleolithic Society. Journal of Human Evolution 5:49-58. Wohlfarth, B. 1996 The Chronology of the Last Termination: A Review of Radiocarbon-dated, High-Resolution Terrestrial Stratigraphies. Quaternary Science Reviews 15(4):267-284. Wright, H. T. 1989 Late Glacial Foragers in Eastern North America. In Eastern Paleoindian Lithic Resource Use, edited by C. J. Ellis and J. C Lothrop, pp. 345-351. Westview Press, Boulder. Wyckoff, D. G. 1996 An Introductory Study of Alibates Gravel Occurrences Along Western Oklahoma’s Canadian River. In In the Light of Past Experience: Papers in Honor of Jack T Hughes, edited by B. C. Roper, pp. 405-452. Panhandle Archeological Society Publication 5. 1993 Gravel Sources of Knappable Alibates Silicified Dolomite.

Geoarchaeology 8:35-58.

396

1993 Grave! Sources of Knappable Alibates Silicified Dolomite. Geoarchaeology 8:35-58. Yellen, J. E. 1977a Cultural Patterning in Fauna! Remains. In Experimental Archaeology, edited by J. E. Yellen and W. MacDonald, pp. 271-331. Columbia University Press, New York. 1977b Archaeological Approaches to the Present: Models for Reconstructing the Past. New York.

Academic Press,

Yengoyan, A. A. 1976 Structure, Event, and Ecology. In Aboriginal Australia: A Comparative Viewpoint. In Tribes and Boundaries in Australia, edited by N. Peterson, pp. 121-132. Australian Institute of Aborigican

Studies, Social Anthropology Series, No. 10, Camberra. Young, R. K. 1997 The Ute Indians of Colorado in the Twentieth Century.

University of Okiahooma Press, Norman.

Zeimens, G. M. 1982 Analysis of Postcranial Remains. In The Agate Basin Site: A Record of the Paleoindian Occupation of the Northwestern High Plains, edited by G. C. Frison and D. J. Stanford, pp. 213-

240. Academic Press, New York. Zimov, S. A., V. I. Chuprynin, A. P. Oreshko, F. S. Chapin III, M. C. Chapin, and J. F. Reynolds 1995 Effects of Mammals on Ecosystem Change at the Pleistocene-Holocene Boundary. In Arctic and Alpine Biodiversity: Patterns, Causes and Ecosystem Consequences, edited by M. C. Chapin, III and C. KOrner, pp. 128-135. Ecological Studies, Vol. 113. Springer-Verlag, Berlin.

View more...

Comments

Copyright © 2017 DATENPDF Inc.