In the Eastern Fluted Point Tradition [1 ed.] 9781607812333, 9781607811701

Eastern North America has one of the largest inventories of Paleoindian sites anywhere in the Americas. Despite this ric

150 12 7MB

English Pages 439 Year 2013

Report DMCA / Copyright

DOWNLOAD FILE

Polecaj historie

In the Eastern Fluted Point Tradition [1 ed.]
 9781607812333, 9781607811701

Citation preview

i n  t h e

Eastern Fluted Point tradition edited by Joseph A.  M. Gingerich

in th e

Eastern Fluted Point tradition

in the

Eastern Fluted Point tradition

edited by Joseph A. M. Gingerich

The University of Utah Press Salt Lake City

Copyright © 2013 by The University of Utah Press. All rights reserved. The Defiance House Man colophon is a registered trademark of the University of Utah Press. It is based on a 4-ft-tall Ancient Puebloan pictograph (late PIII) near Glen Canyon, Utah. 17 16 15 14 13     1 2 3 4 5 Library of Congress Cataloging-in-Publication Data In the Eastern fluted point tradition / edited by Joseph A. M. Gingerich.    p.  cm.   Includes bibliographical references and index.   isbn 978-1-60781-170-1 (cloth : alk. paper)  isbn 978-1-60781-233-3 (ebook) 1. Paleo-Indians — East (U.S.) — Implements.  2. Projectile points — East (U.S.) 3. Stone implements — East (U.S.)  4. East (U.S.) — Antiquities.  I. Gingerich, Joseph A. M.   E78.E2E265 2013  974'.01 — dc23 2012037887 Printed and bound by Sheridan Books, Inc., Ann Arbor, Michigan.

This volume is dedicated to William M. Gardner, who gave many of us opportunities and continues to inspire Paleoindian research.

Contents

List of Figures   ix List of Tables   xiii Preface  xv Acknowledgments  xvii Introduction  1 Joseph A. M. Gingerich Part I. Paleoindian Chronology and Paleoenvironmental Considerations   1. Paleoindian Chronology and the Eastern Fluted Point Tradition   9 Appendix: Select Radiocarbon Dates from Eastern North America   22 D. Shane Miller and Joseph A. M. Gingerich

2. Paleoindian Environment and Subsistence Paradigm Case from New England to Virginia and Ohio   38 Lucinda J. McWeeney



3. Reconstructing the Pleistocene Environment of the Greater Southeast   58 Jessi J. Halligan

Part II. Reinvestigations of Classic Sites  

4. A Report on the 2008 Field Investigations at the Shoop Site (36DA20)   75 Kurt W. Carr, J. M. Adovasio, and Frank J. Vento



5. Spatial Organization at Bull Brook   104 Brian S. Robinson and Jennifer C. Ort



6. Fifty Years of Discovery at Plenge: Rethinking the Importance of New Jersey’s Largest Paleoindian Site   121 Joseph A. M. Gingerich



7. The Wells Creek Site: A Reinterpretation of Site Occupation   148 Jesse W. Tune



8. The Flint Run Complex: A Quarry-Related Paleoindian Complex in the Great Valley of Northern Virginia  156 Kurt W. Carr, R. Michael Stewart, Dennis Stanford, and Michael Frank

9. Revisiting Shawnee-Minisink  218 Joseph A. M. Gingerich

vii

Contents

Part III. New Sites and Perspectives  

10. Paleoindian Toolstone Provisioning and Settlement Organization at the Higgins Site (18AN489)  259 John C. Blong



11. Topper Site, South Carolina: An Overview of the Clovis Lithic Assemblage from the Topper Hillside   280 Ashley M. Smallwood, D. Shane Miller, and Douglas Sain



12. Tennessee’s Paleoindian Record: The Cumberland and Lower Tennessee River Watersheds  299 John B. Broster, Mark R. Norton, D. Shane Miller, Jesse W. Tune, and Jon D. Baker



13. Endscrapers, Use-Wear, and Early Paleoindians in Eastern North America   315 Thomas J. Loebel

Part IV. Observations on the Early Paleoindian Settlement of Eastern North America  

14. Is That All There Is? The Weak Case for Pre-Clovis Occupation of Eastern North America  333 Stuart J. Fiedel



15. The Weight and Meaning of Eastern Paleoindian Research: A View from West of the Rockies   355 Gary Haynes



16. Paleoindian Archaeology in Eastern North America: Current Approaches and Future Directions  371 David G. Anderson

Contributors  405 Index  407

viii

Figures

0.1. Map of selected sites and quarries discussed in the volume.   2 1.1. Summed probability distribution by region.   17 1.2. Summed probabilities corrected for taphonomic bias.   18 2.1. Map of sites discussed in the chapter.   39 3.1. Southeast area during the terminal Pleistocene.   59 3.2. Terminal Pleistocene and Holocene temperature proxy records.   60 3.3. Reconstructed biomes for the Southeast.   64 3.4. Modern biomes for the southeastern United States.   65 4.1. Map of the Middle Atlantic region.   76 4.2. Physiographic zones of Pennsylvania.   78 4.3a. Close-up of the area shown in Figure 4.3b.    79 4.3b. The Shoop site situated in the Ridge and Valley physiographic zone.   80 4.4. Topographic map.  81 4.5. Idealized stratigraphic profile of Steigman’s Bog.   81 4.6. Location of the 2008 excavation.   83 4.7. Schematic of an idealized Shoop site lithic core and the evolution of an endscraper.   87 4.8. Shoop endscrapers from the State Museum collection.   89 4.9. Two jasper adzes from the Shertzer collection.   92 4.10. Shoop fluted points from the State Museum collection.   93 4.11. Map of Fogelman concentrations superimposed over topographic map.   95 4.12. Revised artifact concentrations.   96 4.13. Bar graph comparing artifacts from the Shertzer South locus with total artifacts from the site.   97 5.1. Location map for the Bull Brook site.   107 5.2. Nick Vaccaro’s tool group from Locus 2.   109 5.3. Joe Vaccaro pointing to what William Fowler called the “meat pick.”   111 5.4. Bill Eldridge digging in Locus 11.   112 5.5. Mapping locations superimposed on aerial photograph.   113 5.6. Bull Brook activity distributions.   114 5.7. Selected artifact frequencies that most distinguished interior from exterior loci.   116 6.1. Location of Plenge with nearby rivers and Paleoindian sites.   123 6.2. Map of Plenge with artifact concentrations.   124 6.3. Examples of the different Paleoindian point styles recovered from the Plenge site.   126 6.4. Breakdown of Paleoindian points by phase.   127 6.5. Specimens resembling Vail-Debert-style points.   127 6.6a. Middle Paleoindian points.   128 6.6b. Middle Paleoindian preform.   129 6.6c. Complete Crowfield point.   129 6.7. Small or “stubby” Middle Paleoindian points.   130 6.8. Plano-like Late Paleoindian forms.   130 6.9. Pentagonal form assumed to be Late Paleoindian.   131 ix

Figures

6.10. Examples of fluted preforms from Plenge.   130 6.11. Channel flakes from Plenge.   132 6.12. End- and sidescrapers and limace-like specimens.   135 6.13. Probable Paleoindian tools made from a variety of raw materials.   140 6.14. Points from Plenge by raw material and time period.   141 6.15. Area around Plenge with isolated fluted point finds.   143 7.1. Physiographic provinces of Tennessee.   149 7.2. Clovis points from Wells Creek.   151 8.1. Map of the Flint Run area locating sites discussed in the text.   158 8.2. Topographic cross section of the study area.   158 8.3. Map of excavation areas and geomorphological sections at the Thunderbird site.   172 8.4. Generalized soil profile of the Thunderbird site.   172 8.5. Schematic drawing of activity areas at Thunderbird and Fifty.   179 8.6. Post mold pattern of possible house at the Thunderbird site.   180 8.7. Photograph of Feature 18.   181 8.8. Type I chipping cluster, Feature 68.   181 8.9. Chipping clusters with bipartite pattern, Features 62 and 75.   182 8.10. Chipping area, Feature 110.   182 8.11. Schematic of the Fifty site.   187 8.12. Cross section of the alluvial fan.   187 8.13. Soil profile of Area 3 at the Fifty site facing west.   189 8.14. Cross section of the interfan area at the Fifty site facing west.   190 8.15. Projectile point sequence from the interfan area at the Fifty site.   191 8.16. Correlation of arbitrary .25-ft levels to master levels at the Fifty site.   191 8.17. Drawing and picture of Feature 9.   194 8.18. Map of the reduction of a large primary flake.   195 8.19. Early Paleoindian fluted preforms from the Thunderbird site.   199 8.20. Early and Middle Paleoindian preforms from the Thunderbird site.   200 8.21. Late Paleoindian bifaces from the Thunderbird site.   201 8.22. Projectile points recovered in stratified contexts at the Fifty site.   203 9.1. Map of the first excavations at Shawnee-Minisink by Donald Kline.   219 9.2. Map of the American University excavations.   220 9.3. Location of Shawnee-Minisink at Brodhead Creek and the Delaware River.   223 9.4. Clovis points from Shawnee-Minisink.   225 9.5. Representative profile for the new excavation area.   225 9.6. Orientation of artifacts from Unit 4.   229 9.7. Inclination of artifacts’ long axis.   229 9.8. “Kline Point” from the “Early Early Archaic” levels.   230 9.9. Site map showing new excavations.   231 9.10. Sample of tools from Unit 1.   233 9.11. Excavation area with tools, cores, bifaces, and projectile points mapped in situ.   234 9.12. Excavation area with all artifacts mapped in situ.   234 9.13. Sample of tools from Unit 2.   235 9.14. Sample of tools from Unit 3.   236 9.15. Top of an artifact cluster (clean-up pile).   237 9.16. Photograph of a hearth in profile facing grid south.   238 9.17. Profile of Hearth 2.   238 9.18. Back plot of Unit 4E with artifacts shown above the top of Hearth 2.   242 10.1. Map of Higgins showing the location of sites discussed in the text.   260 10.2. Chart showing the distribution of chert artifacts.   262 x

Figures

10.3. Chart comparing the distributions of chert and quartz debitage types.   266 10.4. Chart comparing the distributions of cortex on chert and quartz debitage.   267 10.5. Chart comparing the distributions of platform preparation on chert and quartz flakes. 267 10.6. Chart comparing the distribution of chert and quartz debitage size.   268 10.7. Artifacts in the Paleoindian assemblage at the Higgins site.   269 10.8. Chart showing the distribution of flake tools by chert color.   270 11.1. Map of the Topper site, with U.S. map showing the locations of Clovis sites.   281 11.2. Map of the Topper site hillside excavation area.   282 11.3. Clovis point fragment found in the buried Clovis component.   283 11.4. Example of large original nodule size from the Topper outcrop.   283 11.5. Example of small original nodule size from the Topper outcrop.   283 11.6. Overshot flakes found in the buried Clovis component.   285 11.7. Examples of preform size variation at Topper.   285 11.8. Example of spall variation used for biface production — ​ big.  286 11.9. Example of spall variation used for biface production — ​ small.  286 11.10. Example of cores from the Topper hillside excavations.   287 11.11. Examples of bifacial tools from the Topper hillside excavations.   287 11.12. Examples of interior blades from the Topper hillside excavations.   288 11.13. Examples of blade cores from the Topper hillside excavations.   289 11.14. Examples of flake tools from the Topper hillside excavations.   290 11.15. Discoidal-shaped endscraper made from welded vitric tuff.   291 11.16. Example of macroscraper/planer from the hillside Block A.   291 11.17. Examples of flake and core tools.   292 11.18. Distribution of blade and irregular cores by excavation block.   293 11.19. Distribution of bifaces and blades by excavation block.   294 11.20. Distribution of flake tools by excavation block.   295 12.1. Map of the Tennessee and Cumberland watersheds.   300 12.2. Selected Paleoindian points from the Tennessee and Cumberland river drainages.   300 12.3. Distribution of Paleoindian bifaces.   303 12.4. The Johnson site.   304 12.5. Excavations at the Coates-Hines site.   306 12.6. Paleoindian points by type recorded in the Tennessee Fluted Point Survey.   307 12.7. John Broster pointing out a cluster of exposed artifacts at the Carson-Conn-Short site.  308 13.1. Map of sites and endscraper assemblages involved in the study.   316 13.2. Gainey endscraper 4026.   320 13.3. Gainey endscraper 3509.   320 13.4. Hawk’s Nest endscraper 71.   321 13.5. Hawk’s Nest endscraper 67/169.   322 13.6. Hawk’s Nest endscraper 27.   322 13.7. Nobles Pond endscraper 27476.   323 13.8. Nobles Pond endscraper 7020.   323 13.9. Nobles Pond endscraper 5570.   324 13.10. Shawnee-Minisink endscraper 181.   324 13.11. Shawnee-Minisink endscraper 1776.   325 13.12. Shawnee-Minisink endscraper 870.   325 15.1. Map of North America south of the ice sheets at about 11,200 Bp.  357 16.1. Baja/Colorado River Interior North American Colonization Model.   392 16.2. Close-up of the northern end of the Gulf of California showing possible movement pathways into the interior.   393 xi

Tables

1.1. Generalized Paleoindian Chronology for the Eastern United States.   10 1.2. Sample Description.  19 4.1. Tool Categories from the Shoop Site by Collection.   82 4.2. Tool Categories and Frequencies from the Shoop Site.   90 4.3. Metrics of a Sample of 72 Fluted Bifaces from the Shoop Site.   91 4.4. Cataloged Artifacts from the Shoop Site by Weight.   92 4.5. Tool Categories and Frequencies from the Shoop Site Excluding the Shertzer Collection.  97 4.6. Tool Categories and Frequencies from the Shertzer Collection.   97 5.1. Bull Brook Artifact Frequencies Separated by Interior and Exterior Loci.   115 5.2. Relative Artifact Frequencies.   115 6.1. Paleoindian Artifacts Collected from the Plenge Site.   125 6.2. Metrics and Descriptions of Channel Flakes.   133 6.3. Plenge Artifacts Examined by Pollock.   136 7.1. Metric Attributes Recorded on Clovis Points in Figure 7.2.   151 8.1. Climatic Episodes for the Northern Shenandoah Valley.   162 8.2. Stages of Manufacture for Clovis-Like Bifaces.   166 8.3. Cataloged Artifacts from the Lockhart Site.   170 8.4. Pedologic Description of Lower Terrace Area of the Thunderbird Site.   173 8.5. Preliminary Artifact Catalog List by Excavation Unit.   174 8.6. Preliminary Artifact Inventory by Excavation Unit and Level in Areas 1B and 4.   175 8.7. Preliminary Artifact Inventory for Features by Excavation Unit and Level in Areas 1B and 4.   178 8.8. Preliminary Counts on Artifacts from the Paleoindian — ​Early Archaic Levels at Thunderbird.  179 8.9. Artifacts Made from Possible Onondaga Chert at Thunderbird.   185 8.10. Typical Pollen Spectrum from the Bog at the Fifty Site.   188 8.11. Pedologic Description of Area 3, Fifty Site.   189 8.12. Tool Variation by Level at the Fifty Site.   193 8.13 Revised Chronology Based on Recent Radiocarbon Dates.   202 9.1. Radiocarbon Dates from the Original Excavations by American University.   219 9.2. Description of Possible Hearth Features from the Original Excavations.   221 9.3. Clovis Point Metrics.   224 9.4. Representative Soil Profile Description for Shawnee-Minisink.   226 9.5. Particle Size Analysis Data for the Excavation Area.   228 9.6. Tool Counts by Test Unit.   232 9.7. Summary Metrics for Complete Endscrapers.   232 9.8. Radiocarbon Ages from Shawnee-Minisink.   239 9.9. Chi-Square Test Comparing the Number of Occurrences of Local and Nonlocal Debitage by Excavation Quad.   240 9.10. Local and Nonlocal Distribution Among the Test Units.   243 xiii

Tables

9.11. Edible Plant Totals from Old and New Excavations.   249 10.1. Debitage Types in the Fluted Point Assemblage.   264 10.2. Core Data.  270 10.3. Flake Tool Data.   271 11.1. Technological Variables Recorded in the Biface Analysis.   284 11.2. Cores by Excavation Block.   293 11.3. Bifaces and Blades by Excavation Block.   294 11.4. Counts of Other Artifact Classes by Excavation Block.   295 12.1. Paleoindian Period Chronology.   299 12.2. Preliminary Counts of Carson-Conn-Short Artifacts from Controlled Surface Collections and Test Excavations.   301 12.3. Radiocarbon Dates for the Johnson Site.   304 13.1. Summary Microwear Results.   319 13.2. Ratio of Hide-Processing Tasks Occurring at Sites in the Sample.   326 14.1. Dates from Meadowcroft, Stratum IIa and the I/IIa Interface.   334 14.2. Stone Tools from Stratum II.   337 14.3. Reported Botanical Remains from Meadowcroft Rockshelter, Stratum IIa.   338 14.4. Pre-Clovis Radiocarbon Dates from Cactus Hill.   343 14.5. Debitage Counts by Level, Units 1/9 and 2/9, Cactus Hill.   344 16.1. Raw Material Types and Incidence on Typed Clovis and Presumed Clovis Fluted Points.  380

xiv

Preface

The title of this book, The Eastern Fluted Point Tradition, is borrowed from Errett Callahan’s (1979) book in which he outlines the projectile point manufacturing techniques used by Early Paleoindians in the Middle Atlantic region of eastern North America. While his title may suggest a divergence in eastern fluted point styles from their western counterparts — ​where fluted points were first defined — ​Callahan’s intention was to examine the manufacture techniques used by the earliest occupants of eastern North America and provide a better understanding of fluted point variability and technology. While this book does not specifically focus on the variation of fluted point styles or the detailed work of Callahan, I could not think of a better title for a book that covers a broad range of topics and sites related to the people who once lived in eastern North America and are recognized for their distinctive fluted points. In coming up with the idea for this book, the contributors and I realized that there were few published summaries of eastern Paleoindian research. The first two that came to mind were Eastern Paleoindian Lithic Resource Use by Christopher Ellis and Jonathan C. Lothrop (1989) and The Paleoindian and Early Archaic Southeast by David G. Anderson and Kenneth E. Sassaman (1996). While these and others are useful research references, a single volume had yet to be compiled that incorporates individual site reports. It has also been over 15 years since the last book on eastern

Paleoindian research was published, and in the interim, a multitude of new and interesting research projects have been initiated. In an attempt to communicate some of these latest findings to a broad audience, I compiled a number of papers that provide reports on reinvestigations of old sites, preliminary reports on new sites, and updates on some of the current Paleoindian research in eastern North America. In preparation for this volume, I organized a symposium at the 75th Anniversary Meeting of the Society for American Archaeology in St. Louis, Missouri, in 2010. Although the s­ ession was held in the evening, over 200 people attended. Most of the papers presented at the symposium are included in this volume. I thank Jim Adovasio, Dick Bosivert, Jim Dunbar, Al Goodyear, Andrew C. Hemmings, and Stephen Loring for their participation in the symposium and regret that they could not be a part of this volume. David Anderson served as discussant in the symposium and agreed to expand his comments to form the summary chapter of this book. The chapters presented herein provide updated information on many of the best-known Paleoindian sites in eastern North America. My only regret is that this volume does not represent more regions in the east. The data provided in this volume, however, will undoubtedly encourage further consideration of the sites included and garner comparisons with other sites throughout eastern and greater North America.

xv

Acknowledgments

I would like to thank the anonymous readers who reviewed this volume. Their insights and edits were very helpful in finishing the final manuscript. I also appreciate comments from Dave Anderson, Gary Haynes, Bob Kelly, and Todd Surovell, who read and made suggestions on vari­ ous chapters. Mary Lou Larson and Reba Rauch (University of Utah Press Acquisitions Editor)

were very helpful through every step of the publication process. Jennifer Horvath, Willa Mullen, and Heather Rockwell all contributed to the volume in various ways. Their assistance helped to ensure the timely completion of the manuscript. Finally, this volume helped me fulfill the Surovell Program’s “Plan A” publication quota for graduation — t​ hanks, Todd!

xvii

Introduction Joseph A. M. Gingerich

Eastern North America has one of the largest inventories of Paleoindian sites anywhere in the Americas. Despite this rich record of early human settlement, there are few widely published reports or summaries of Paleoindian research in this region. Many of the best-known Paleoindian sites in eastern North America were discovered over 40 years ago. Some these sites have received better coverage, such as Shawnee-Minisink (Dent 2002; McNett 1985), Debert (MacDonald 1968), and Vail (Gramly 1982, 1984, 2009), but most have received limited exposure at regional conferences, in regional journals and special publications, or in hard-to-find dissertations and master’s t­ heses. The goal of this book is to provide a single edited volume that reports on over four decades of Early Paleoindian research in eastern North America, including previously unpublished site reports and updates on recent research (Figure I.1). The book is designed to be data-rich, providing counts and metrics on artifacts and basic site descriptions that will allow for comparisons to be made between sites in the region. Once I received the final chapter for this book, it became apparent how many sites and how much new research in eastern North America are not in this volume and in fact, could not be reported in a single volume. Nevertheless, the chapters in this book provide a step toward creating a more cohesive picture of the early human occupation of eastern North America. My hope is that additional volumes will follow that will continue to garner interest and encourage the pursuit of new Paleoindian research in eastern North America.

The book is divided into four sections: (1) Paleo­ indian Chronology and Paleoenvironmental Considerations, (2) Reinvestigations of Classic Sites, (3) New Sites and Perspectives, and (4) Observations on the Early Paleoindian Settlement of Eastern North America. A preview of each chapter follows. Paleoindian Chronology and Paleoenvironmental Considerations

Interpretations of the climate and environment that existed when people first arrived in eastern North America have strongly influenced perceptions of Paleoindian settlement patterns and adaptations (e.g., Custer and Stewart 1990; Gardner 1974; Meltzer 1988). It is therefore appropriate to start this volume with an overview of Paleo­ indian chronology in eastern North America and data relating to the environment at the time. Miller and I begin by summarizing our current knowledge of Paleoindian chronology in eastern North America. Using a large number of radio­ carbon dates from the east, we examine differences in the frequency of radiocarbon dates by region. These data underscore differences in the radiocarbon record that may be linked to settlement and environmental change. Miller and I evaluate gaps in the radiocarbon record and further address issues relating to our understanding of point typology and morphology during the Paleoindian period. The chapters by Lucinda McWeeney and Jessi Halligan summarize the ­environmental conditions during the Paleoindian period. ­McWeeney’s 1

Introduction

Figure 0.1. Map of selected sites and quarries discussed in the volume.

Reinvestigations of Classic Sites

update of a chapter previously published in 2007 offers an examination of ­pollen cores and macro­ botanical remains from or near archaeological sites in the Northeast. The increasing presence and identification of more thermophilic plants during the late Pleistocene suggests a more diverse environment than previously was thought to have exsited in areas of the Northeast. Environmental data from the Southeast are s­ ummarized by Halligan, who highlights the complexity of late Pleistocene environments in this region. Halligan’s chapter emphasizes the numerous changes that likely occurred during the Paleoindian period. How these changes would have affected people is unknown, but changes along coastal areas and the presence of complex faunal communities were at least two factors that likely had a major effect on the activities, survival, or lifestyles of early popu­lations.

Reports containing the most detailed summaries for many important Paleoindian sites make up the heart of this book. They are arranged by the date on which they were discovered. The Shoop site was one of the first to be widely published in eastern North America (Witthoft 1952). Not only is it considered to be one of the larger sites in eastern North America, but it is unique due to the occurrence of several discrete loci, along with the recovery of thousands of artifacts made from a single lithic raw material from over 300 km distant. Carr et al. provide an indepth summary of the work at Shoop and include some of the first detailed accounts of new excavations at the site. Recent analyses of the assemblage complement Cox’s (1986) earlier analysis of the site and help provide a better understanding of the site and its artifacts. 2

Introduction

Robinson and Ort’s Bull Brook chapter is important as it provides a detailed history and description of the excavation and analysis of this large and important Paleoindian site. This chapter complements their recent article on Bull Brook that documents their reanalysis and revised interpretation of the site (Robinson et al. 2009). As the collectors who discovered the site, Eldridge and the Vaccaros, did over 50 years ago, Robinson and Ort provide a record that will allow for future analyses, critiques, and expansion of data from this important site. The Plenge site is one of the best-known Paleo­ indian sites in the Middle Atlantic region. It has long been recognized as an extensively collected surface site, and new research highlights the potential for future research at the site. The recovery of nearly every known fluted point type in eastern North America at this site suggests that certain locations on the landscape remained important throughout the Paleoindian period. New counts, a delineation of activity loci, and descriptions of the Paleoindian assemblage in this chapter provide a useful database that will contribute to a better understanding of Paleoindian land use and technology in eastern North America. The Wells Creek site in Tennessee has long been a part of the Paleoindian literature in eastern North America. It is widely known as one of the earliest sites in Tennessee, and the location and occupation of this site have contributed to many of the Paleoindian settlement pattern models in the region. Tune’s reanalysis of the Wells Creek assemblage suggests that the Early Paleoindian occupation may have been more transient and less of a focal point on the landscape. This information and the complete recharacterization of the assemblage by Tune will require a reassessment of this site’s role in current models of Paleoindian settlement patterns. The report by Carr et al. on Thunderbird and the Flint Run complex summarizes over 20 years of work on this Early Paleoindian site in the Shenandoah Valley of Virginia. Most of the work summarized in this chapter was conducted between 1971 and 1992. Although some information has been disseminated in the form of lab reports, dissertations, and master’s theses, this is the first time these data have been compiled and published in a widely accessible volume. Kurt Carr

and Michael Stewart, two former students of Bill Gardner, do an exceptional job of not only summarizing this material but explaining how Gardner’s work at Flint Run influenced Early Paleo­ indian studies in the region “then” and today. The need for more work with the Thunderbird assemblage has long been apparent, as well as the need to further refine the models generated from the work at Flint Run. The information presented in the Thunderbird chapter demonstrates that sections of the site, especially the lowest levels, contain impeccable chipping clusters where further examination of technology and spatial patterns will be achievable. Dennis Stanford, Michael Frank, Pegi Jodry, and others at the Smithsonian have now made the Thunderbird material accessible for study. It will be exciting to learn more about this important site in the future. The new investigations at Shawnee-Minisink, which are summarized in this volume, provide additional data on the Paleoindian occupation. With four new radiocarbon dates, the excavation of two Paleoindian hearth features, and the recovery of over 30,000 new Paleoindian artifacts, the new excavations highlight the presence of several dense activity loci across the presumed extent of the site. Analysis of features and preliminary observations of the spatial arrangements of artifacts suggest a more substantial Paleoindian occupation than previously reported. New Sites and Perspectives

Although not a new site, the Higgins site is included in this section because it is not well known outside of the Middle Atlantic region. Carol Ebright, who was originally in charge of excavating and analyzing the assemblage from the site, provided one large and comprehensive report in 1992. The data in the report, however, are not easily accessible due to its limited distribution. Blong’s reanalysis of the Paleoindian assemblage accomplishes three goals: (1) It summarizes Ebright’s work on the site, (2) it enhances our understanding of the site’s stratigraphy and evidence of Paleoindian technology, and (3) it builds off Ebright’s work to further tie the site to regional patterns of lithic use. Focusing on this latter accomplishment, the Higgins site further illustrates differences in Paleoindian assemblages in coastal regions and the exploitation 3

Introduction

of ­secondary and lower-quality lithic sources. Studies like these create a more holistic picture of Paleoindian settle­ment that may be attributed to seasonal and environmental differences (see also Custer 1989; Dent 1995; Lowery 2002). Finally, the artifacts from Higgins, possibly manufactured from Normanskill chert, may further enhance our knowledge of lithic movement and patterns of curation or demonstrate the variety of chert available from cobble sources near the coast (i.e., Lowery 2002). Over the last decade, the Topper site has gained a lot of attention for its proposed pre-­ Clovis­occupation. However, it is equally important to document the well-established Clovis living floors at Topper. Smallwood et al. provide the first comprehensive report on the Clovis occupation at Topper. Although this chapter only highlights 128 m2 of the 622-m2 area of the Clovis component that has been excavated, it provides the first total artifact counts for this area and details the horizontal and vertical distribution of Clovis artifacts, which allows us to better understand the site and its early occupation(s) (see also Miller 2010). Smallwood et al. also document aspects of the lithic technology at Topper that comfortably fits with Clovis reduction strategies and tool kits. As with the other detailed discussions of assemblages in this volume, documenting reduction techniques and variation will lead to increased insight into and identification of different Paleo­ indian technologies. The limited spatial analyses presented in this report suggest that there may be several discrete activity loci represented at Topper, which may point toward the segregation of activities at the site. As concluded by the authors of this chapter, the activities at Topper appear to be more than just quarrying and tool production but, rather, a number of activities. In light of these observations and the number of artifacts that have been recovered from Topper, these characteristics are similar to those reported for the Thunderbird site (Carr et al., Chapter 8, this volume). In 2005, Goodyear characterized Topper as part of a complex of sites that are centered around the Allendale chert quarries. While there is much work to be completed at Topper and surrounding sites, it will be interesting if spatial patterns, site location, and site function closely resemble the patterns proposed by Gardner for Flint Run and

Paleoindians in the Shenandoah Valley. Perhaps Topper and nearby sites will provide another opportunity to refine and test aspects of the Flint Run Model. The summary of Tennessee Paleoindian archaeology by Broster et al. is a nice addition to this volume. In addition to better-known sites in Tennessee, such as Carson-Conn-Short and Dust Cave, this chapter highlights the rich late Pleistocene occupation of Tennessee and its river valleys. Most importantly, this chapter reports on recent Paleoindian finds, which include many sites that have not been the subject of publications. The sites discussed in this chapter underscore the need for additional work on Tennessee Paleo­indian sites. The discovery of sites such as Carson-Conn-Short and Johnson, which may represent some of the earliest sites in the region, provides a new opportunity to better understand Paleoindian technology, especially when compared with other sites in the volume that exhibit entire lithic reduction sequences. Loebel presents one of the first large comparisons of use-wear studies on Early Paleoindian tools in eastern North America. His work illustrates the value of use-wear to determine the presence of contact items that are not preserved in the archaeological record (i.e., organic perishables) and the application of use-wear to decipher the use and life history of stone tools. The identification of specific contact materials combined with evidence of a range of hide-working activities at these sites helps to provide a better picture of site function and season of occupation. A case in point is Shawnee-Minisink, where L ­ oebel’s work provides insight into site activities while lending support to the hypothesized fall occupation and the collection of berries as supplements to other dietary items (Dent 2007; Gingerich 2011, and Chapter 9, this volume). Loebel’s chapter also highlights the need for and utility of more usewear studies, which have steadily grown (Cox 1986; Evans 1978; Marshall 1981, 1985; Wilmsen 1970) and continue to expand in eastern Paleo­ indian studies (e.g., Pope 2010; Rockwell 2010; Shoberg 2005; Spiess and Wilson 1987). With Loebel’s work and that of others, we are building a database of use-wear studies that may assist in comparing broad patterns of Paleoindian settlement and lifeways. 4

Introduction

Observations on the Early Paleoindian Settlement of Eastern North America

with many of the early sites in this region and offers a unique perspective on the progress of eastern Paleo­indian research over the past 30 years. Linking this research to western Paleoindian sites, Haynes offers a continental perspective on eastern Paleoindian studies, emphasizing the importance of eastern North America in understanding the timing, direction, and speed of Paleoindian colonization. As a conclusion to the book, Anderson provides a thoughtful review of the chapters, highlighting their importance to our understanding of the early settlement of eastern North America. His discussion lays the groundwork for future studies, which may draw heavily from the data presented in this book.

Crucial to our understanding of the spread of populations and Paleoindian material culture is the recognition and understanding of sites that may represent the earliest occupations in eastern North America. Given the small number of very early or “pre-Clovis” sites reported across the Americas, a disproportionate number have been described for eastern North America. Fiedel, a staunch opponent of many reported pre-Clovis sites, gives an evenhanded review of many of these purported pre-Clovis sites. His detailed summary raises questions about the validity of designating these sites as pre-Clovis, although some do appear to represent occupations that predate the age range for most Clovis sites. ­Fiedel’s skepticism, however, is generally due to the lack of information presented in published reports. In such cases, many of Fiedel’s questions or challenges can be easily answered by additional (and limited) analyses of these assemblages. It is important to emphasize that many arguments for pre-Clovis sites have become more convincing over the years. It is now necessary to critically examine what these sites represent. The next chapter, by Haynes, offers a western perspective on the research presented in this volume. Haynes, trained in the east, is familiar

The research presented in this volume will be a necessary source for scholars, students, and avo­ cational archaeologists interested in the early settlement of the Americas. Although it is limited to a review of sites in eastern North America, the information presented will be of interest to anyone who studies Paleoindians and the settlement of the Americas. I hope that this volume will provide a better foundation for future late Pleistocene studies in the east and throughout North America.

References Cited ware Valley: Revisiting Shawnee Minisink and Nearby Sites. In Ice Age Peoples of Pennsylvania, edited by K. W. Carr and J. M. Adovasio, pp. 51–78. Pennsylvania Historical and Museum Commission, Harrisburg. 2007 Seed Collecting and Fishing at the Shawnee Minisink Site: Everyday Life in the Pleistocene. In Foragers of the Terminal Pleistocene in North America, edited by R. B. Walker and B. N. Driskell, pp. 116–131. University of Nebraska Press, Lincoln. Ebright, Carol A. 1992 Early Native American Prehistory on the Maryland Western Shore: Archaeological Investigations at the Higgins Site Vols. 1–2. Maryland State Highway Administration ­Archaeological Report No. 1. Prepared for the Maryland State Railroad Administration, Annapolis.

Cox, Steven L. 1986 A Re-Analysis of the Shoop Site. Archaeology of Eastern North America 14:101–173. Custer, J. F. 1989 Prehistoric Cultures of the Delmarva Peninsula: An Archaeological Study. University of Delaware Press, Newark. Custer, J. F., and M. Stewart 1990 Environment, Analogy and Early Paleo­ indian Economies in North Eastern North America. In Early Paleoindian Economies of Eastern North America, edited by K. B. Tankersley and B. L. Isaac, pp. 303–322. Research in Economic Anthropology 5. JAI Press, Greenwich, Connecticut. Dent, Richard J. 1995 Chesapeake Prehistory: Old Traditions, New Directions. Plenum Press, New York. 2002 Paleoindian Occupation of the Upper Dela­ 5

Introduction Stratified Paleoindian–Archaic Site in the Upper Delaware Valley of Pennsylvania, edited by C. W. McNett, Jr., pp. 165–209. Academic Press, Orlando. McNett, Charles W., Jr. (editor) 1985 Shawnee-Minisink: A Stratified Paleoindian– Archaic Site in the Upper Delaware Valley of Pennsylvania. Academic Press, Orlando. Meltzer, David J. 1988 Late Pleistocene Human Adaptations in Eastern North America. Journal of World Prehistory 2:1–52. Miller, D. Shane 2010 Clovis Excavations at Topper 2005–2007: ­Examining Site Formation Processes at an Upland Paleoindian Site Along the Middle Savannah River. Occasional Papers 1 of the Southeastern Paleoamerican Survey. South Carolina Institute of Archaeology and Anthropology, Columbia. Pope, Melody 2010 A Microwear Study on a Small Sample of End Scrapers from the Shoop Site. Manuscript on file at the State Museum of Pennsylvania, Harrisburg. Robinson, Brian S., Jennifer C. Ort, William E. Eldridge, Adrian L. Burke, and Bertrand G. Pelletier 2009 Paleoindian Aggregation and Social Context at Bull Brook. American Antiquity 74(3):423– 447. Rockwell, Heather M. 2010 Use Wear Analysis of the Potter Site: A Paleo­ indian Site in New Hampshire. Unpublished Master’s thesis, University of Tulsa Graduate School. Shoberg, Marilyn B. 2005 Unpublished report of use-wear analysis of artifacts from the Potter site. Randolph, New Hampshire. Spiess, Arthur E., and Deborah B. Wilson 1987 Michaud: A Paleoindian Site in the New ­England–​Maritimes Region. Occasional ­Publications in Maine Archaeology, No. 6. Augusta. Wilmsen, Edwin N. 1970 Lithic Analysis and Cultural Inference: A ­Paleo-Indian Case. University of Arizona Press, Tucson. Witthoft, J. 1952 A Paleo-Indian Site in Eastern Pennsylvania: An Early Hunting Culture. Proceedings of the American Philosophical Society 96(4). Philadelphia.

Evans, June 1978 Paleo-Indian to Early Archaic Transition at the Shawnee-Minisink Site. Ph.D. dissertation, American University. University Microfilms, Ann Arbor. Gardner, William M. 1974 The Flint Run Complex: Pattern and Process During the Paleo-Indian to Early Archaic. The Flint Run Paleo-Indian Complex: A Preliminary Report 1971–1973 Seasons, edited by William M. Gardner, pp. 5–47. Occasional Publication No. 1. Archeology Laboratory, Department of Anthropology, Catholic University of America, Washington, D.C. Gingerich, Joseph A. M. 2011 Down to Seeds and Stones: A New Look at the Subsistence Remains from ­Shawnee-​ Minisink. American Antiquity 76(1):127–144. Goodyear, Albert C., III 2005 Evidence for Pre-Clovis Sites in the Eastern United States. In Paleoamerican Origins: Beyond Clovis, edited by R. Bonnichsen, B. Bradley, D. Stanford, and M. Waters, pp. 103– 112. Center for the Study of the First Americans, Texas A&M University, College Station. Gramly, Richard Michael 1982 The Vail Site: A Palaeo-Indian Encampment in Maine. Bulletin of the Buffalo Society of Natural Sciences, Vol. 30. Buffalo. 1984 Kill Sites, Killing Ground and Fluted Points at the Vail Site. Archaeology of Eastern North America 12:110–114. 2009 Palaeo Americans and Palaeo Environment at the Vail Site, Maine. Persimmon Press, North Andover. Lowery, D. L. 2002 A Time of Dust: Archaeological and Geomorphological Investigations at the Paw Paw Cove Paleo-Indian Site Complex in Talbot County, Maryland. Maryland Historical Trust, Crownsville. MacDonald, G. F. 1968 Debert: A Paleo-Indian Site in Central Nova Scotia. Anthropology Papers No. 16. National Museum of Canada, Ottawa. Marshall, Sydne B. 1981 Artifact Form and Function: Implications of Morphological Classification and Wear Pattern Analysis on Cultural Interpretation. Ph.D. dissertation, Columbia University. University Microfilms, Ann Arbor. 1985 Paleoindian Artifact Form and Function at Shawnee Minisink. In Shawnee-Minisink: A

6

I

Paleoindian Chronology and Paleoenvironmental Considerations

1

Paleoindian Chronology and the Eastern Fluted Point Tradition D. Shane Miller and Joseph A. M. Gingerich

The study of the Paleoindian period in the eastern United States is an exercise in contrasts. On one hand, this region has some of the highest densities of artifacts in North America. This pattern became evident with the first survey of fluted points in Virginia in the late 1940s and subsequent state compilations by Mason (1962) and Brennan (1982). Beginning in the late 1980s, Anderson (1990, 1996) began integrating the various state fluted surveys as well as other published sources into a single database. Like previous researchers (e.g., Brennan 1982; Mason 1962), he found that the occurrences of fluted points in the eastern United States dramatically outnumbered those found in the western United States. While several studies have attempted to show that this discrepancy is due to biases such as land cover and modern population, certain areas such as the Tennessee, Ohio, Savannah, Aucilla, and Cumberland river drainages are significantly overrepresented (Buchanan 2003; Prasciunas 2008, 2011). In addition to the overall numbers of fluted points in the eastern United States, there is also considerable variation in the morphology of these points. For example, in the western United States, the Clovis and Folsom types are considered to be the primary expressions of fluted point technology (Holliday 2000; Meltzer 2009). In the eastern United States, there are multiple other fluted, post-Clovis types including Cumber­ land, Barnes, Crowfield, Debert-Vail, Gainey, Michaud-­Neponset­, Redstone, Dalton, and others

(­Anderson 2001; Anderson et al. 1996; Bradley et al. 2008; Curran 1996; Ellis and Deller 1997; Goodyear 1999; Table 1.1). Despite the abundance and variety of fluted bifaces, there is one glaring obstacle that hampers the study of the Paleoindian period in the eastern United States — ​the absence of stratified sites with datable material (Anderson et al. 2005; Curran 1996; Dunnell 1990; Levine 1990; Smith 1986; Steponaitis 1986). Most fluted bifaces come from shallow sites, usually from plowed contexts. Furthermore, most artifacts are not recovered by professional archaeologists but, instead, reside in private collections (e.g., Fogelman and Lantz 2006; Goodyear 1999). Meltzer (1988) interpreted this pattern as evidence of extreme mobility, whereby small groups moved rapidly across the landscape leaving little aside from small sites and scattered, isolated bifaces. Others have argued that this pattern of shallowly buried sites is most likely due to a broadscale geomorphological bias. Dunnell (1990) observed that compared with other areas in North America, the southeastern United States is situated on a much older landscape, with many upland areas receiving little to no sedimentation since the Tertiary period. Additionally, the warm, mesic climate of this region promotes the decay of materials that can be radio­carbon dated. This pattern is reflected in recent continental-scale databases of radio­ carbon dates, which show this region to have relatively fewer numbers of reported radiocarbon 9

Miller and Gingerich Table 1.1. Generalized Paleoindian Chronology for the Eastern United States.

Period

cal BP

rcybp

Major Cultural Components

Late Paleoindian

12,550–11,400

10,500–10,000

Middle Paleoindian Early Paleoindian Pre-Clovis

12,800–12,550 10,800–10,500 13,500–12,800 11,500–10,800 >13,500 >11,500

Quad, Beaver Lake, Barnes/Michuad-Neponsent, Crowfield, Dalton, Plano Cumberland?, Debert-Vail, Redstone, Suwannee Clovis, Gainey ?

Source: Adapted from Anderson 1996; Anderson et al. 2005.

dates (­Buchanan et al. 2008; Waters and Stafford 2007). Goodyear (1999) relates the lack of stratified sites to late Pleistocene erosive conditions, which has been supported by recent research by Leigh (2004, 2006), who has shown that the major rivers of the South Atlantic Coastal Plain transitioned from braided rivers during the Last Glacial Maximum to large meandering rivers during the terminal Pleistocene. As a result, many of the terraces that could have contained archaeological material may have been eroded and destroyed as a result of this transition. Likewise, landscapes in the Northeast and Mid-Atlantic have been subject to similar geologic processes that limit the number of stratified sites and datable material. Many northeastern Paleoindian sites, for example, are located on sandy nondepositional landforms, which results in site burial through pedoturbation (Curran 1996:4). These, along with the reworking of alluvial terraces during the late Pleistocene and early Holocene (Gingerich 2007; Stewart et al. 1991; Wagner 1994), are the main reasons why intact stratified deposits are rare in the region. In addition, the mixing of shallow deposits and the incorporation of noncultural or recent materials have produced erroneous dates for many sites in the Northeast (e.g., Bonnichsen and Will 1999; Byers 1959; Curran 1984, 1999). Consequently, eastern North America has perhaps the richest material record for the Paleo­ indian period but is the poorest in terms of context and preservation. In the western United Sates, stratified sites such as Blackwater Draw (Boldurian and Cotter 1999), Wilson-Leonard (Collins 1998), Jimmy Pitts (Sellet et al. 2009), Carter/Kerr McGee (Frison 1984), Hell Gap (Larson et al. 2009), and Lubbock Lake (Johnson and Holliday 2004) have helped to clarify the stratigraphic relationship among the Clovis, Folsom, and other point types (Meltzer 2004, 2009). In the

eastern United States, no such site exists (except possibly, to some degree, Thunderbird, discussed in this volume). Instead, there has been a long history of extrapolating culture-­historical sequences based on similarities in the biface morphology of more well-dated types from other areas. For example, Soday (1954) equated the fluted bifaces found at the Quad locality in northwestern Ala­ bama with those from sites in the northeastern United States, most notably Shoop and Vail. ­Dragoo (1965, 1973), in his description of artifacts recovered from the Well’s Creek Crater site in Tennessee, drew comparisons with Clovis sites in the western United States. This tradition has continued with more recent discussions over whether eastern and western Clovis types are equivalent (e.g., Beck and Jones 2010; Haynes 1983; Haynes 2002; Morrow and Morrow 1999; Waters and Stafford 2007) and if fully fluted forms such as Barnes and Cumberland are contemporaneous with Folsom (Anderson 2001; Ellerbusch and Yerka 2005; Ellis 2004). Finally, forms with deeply indented bases, most notably the Redstone type, are often equated with similar types from dated sites in the northeastern United States, such as Vail and Debert (Daniel and Goodyear 2006; Goodyear 2006). While this has provided an alternative strategy for sorting out some of the ambiguity in the relative sequence of eastern fluted types, there have been relatively few attempts to approach this issue in a systematic and quantitative manner. In this chapter, we will discuss the s­ tratigraphic and radiocarbon record of the early Holocene and terminal Pleistocene in eastern North America to illustrate how dated components occurring during the Younger Dryas are particularly rare. Additionally, we will present the results of a ­pilot analysis of metric attributes from bifaces proposed to have been produced during this time span and those from dated contexts in other re10

Paleoindian Chronology and the Eastern Fluted Point Tradition

gions. Finally, we will discuss future directions for clarifying the culture history of the early Holocene and Paleoindian periods of the eastern United States.

These hearths produced two dates: 10,530 ± ​650 14C  bp (ISGS-48 [Coleman 1972:154]) and 10,200 ± ​ 330 14C  bp (M-2333 [Crane and Griffin 1972:159]). Consequently, Goodyear (1982, 1999) argues that Dalton represents a Late Paleoindian horizon distinct from later Early Archaic components containing notched bifaces that dates to between 10,500 and 9900 14C  bp. This time frame has been subsequently upheld with recent research at sites such as Dust Cave in northern Alabama (Sherwood et  al. 2004) and Big Eddy in Missouri (Hajic et al. 2000; Lopinot et al. 2000). At Dust Cave, Sherwood et al. (2004) identified a Quad and Beaver Lake component below the Dalton component, although for technological analyses they combined all of these into a general “Late Paleoindian” assemblage. Beaver Lake points are slightly waisted lanceolates with faint ears, slightly concave bases, and moderate basal thinning (Cambron and Hulse 1975; Justice 1995:35– 36). Quad points have distinct ears, a concave base, and pronounced basal thinning (Cambron and Hulse 1975; Justice 1995:​35–36). A sample of eight radiocarbon dates from zone U, which is associated with the Quad and Beaver Lake component, averages 10,388 ± ​32 14C  bp.1 However, despite the abundance of fluted bifaces recovered in this region, the radiocarbon record is relatively sparse (Anderson 1995, 1996; Anderson and Faught 1998; Anderson et al. 2005; Prasciunas 2008). Moreover, this area has some of the greatest variation in fluted forms, or types, in North America. Unfortunately, as previously discussed, most of these types are found out of context by private collectors, and there are only a limited number of buried, preserved components or single-component surface sites. For example, famous sites such as Nuckolls (Lewis and Kneberg 1959; Norton and Broster 1992), Wells Creek Crater (Dragoo 1965, 1973), and Quad (Cambron and Hulse 1960; Soday 1954) were surface sites that were heavily collected. On the other hand, sites such as Carson-Conn-Short in Tennessee (Broster and Norton 1996), Topper in South Carolina (Goodyear and Steffy 2003; Miller 2010; Smallwood et al., this volume), and Harney Flats in Florida (Daniel and Wisenbaker 1987) have stratigraphically discrete components with Paleo­ indian bifaces. Organic preservation, however, is poor at these sites, and material for radio­carbon

The Paleoindian Period in the Southeastern United States

In the southeastern United States, the Late Paleo­ indian period spans the boundary between the early Holocene and the beginning of the Younger Dryas (ca. 10,800–10,000 14C  bp or ca. 12,900– 11,500 cal Bp), a return to glacial-like conditions at the end of the Pleistocene epoch (Anderson et al. 2005; Goodyear 1999). In the southeastern and Mid-Atlantic/Midwestern United States, the Late Paleoindian period is signaled by the appearance of the Dalton point type (Gardner 1974, 1983; Goodyear 1982:390). Dalton bifaces typically begin their use-life as lanceolates with concave bases and serrated edges. Often the basal margins are parallel to slightly incurvate, while the blade portion is initially excurvate (Justice 1987:40–42). Several studies have shown that the blade margins transition from excurvate to incurvate through repeated resharpening (Goodyear 1982; Shott and Ballenger 2007). Dalton points were originally thought to have been an Early Archaic–type biface dated to between 10,000 and 8,000 radiocarbon years ago as a result of a series of dates from cave and rockshelter sites such as Graham Cave (Crane and Griffin 1956) and Arnold Research Cave (Crane and Griffin 1968; Shippee 1966) in Missouri and Stanfield-Worley Rockshelter in Alabama (DeJarnette et al. 1962). Goodyear (1982), however, argued that in several instances, it is likely that several of these dates may actually come from Dalton assemblages mixed with subsequent side-notched and corner-notched occupations. Moreover, due to research at open-air sites such as Brand (Goodyear 1974) and Sloan (Morse 1997) in Arkansas, the case can be made that Dalton is distinct from the notched varieties and is more similar to fluted bifaces and their associated technologies. In addition to Dalton being technologically distinct from notched forms, Goodyear also cites research from Rogers Shelter in Missouri (Wood and ­McMillan 1976), which contained a pure Dalton component with well-preserved hearths that was capped with culturally sterile alluvium. 11

Miller and Gingerich

dating has not been found. In fact, the only reliable Clovis-age date from the Southeast is a proboscidean ivory rod from Sloth Hole, a submerged sinkhole site in Florida (11,050 ± 50, SL285 [Hemmings 2004; Waters and Stafford 2007]). Despite stratigraphic ambiguity, several different Paleoindian point types have been defined in this region (Anderson et al. 2005; Goodyear 1999; Meltzer 2009). These include Clovis-, Redstone-, and Cumberland-type bifaces. Clovis bifaces are large, parallel-sided lanceolate bifaces with slightly concave bases and single or multiple flutes that rarely extend more than a third of the body (Bradley et al. 2010; Howard 1988, 1990; Justice 1995:​17–21; Sellards 1952). The Redstone type is defined by an overall triangular form that is widest at the base. These bifaces often have indented bases and more extensive fluting than Clovis bifaces (Goodyear 2006; Justice 1995:22). Cumberland bifaces are identified by their narrow, deeply fluted and slightly waisted appearance. The bases are also slightly concave and often have faint ears (Justice 1995:​25–27; Lewis 1954). Redstone and Cumberland bifaces are believed to immediately postdate Clovis, although there is no single site that clearly underwrites this proposed sequence (Anderson et al. 1996:9–11; Goodyear 1999:435). In the South Atlantic Coastal Plain and Florida, the Suwannee and Simpson types are more prevalent (Anderson 1996; Dunbar 2006a; Goodyear 1999). Suwannee points have incurvate basal margins and slightly projecting ears, whereas Simpson points have strong incurvate basal margins and more pronounced ears. These bifaces are for the most part unfluted but have shallow basal thinning (Bullen 1975; Goodyear 1999:439). As with other Paleoindian bifaces, there appears to be no site that clearly defines the chronological relationship among Suwannee, Simpson, and Clovis points (Dunbar 2006b). In the southeastern United States, there are several sites that are reported to have potential pre-Clovis components; the most well known are the Topper (Goodyear 2005) and Page-­Ladson sites (Dunbar 2006a, 2006b). The Topper site is located on an alluvial terrace along a chute channel of the Savannah River and was originally documented as having Early Archaic through ­historic-​ period components (Goodyear and Charles 1984). Through the course of ensuing excavations, an ad-

ditional Clovis component was discovered both on the alluvial terrace (Goodyear and Steffy 2003) and on the adjacent hillside (Goodyear et al. 2007; Miller 2010). Excavations below the Clovis component revealed an additional component below culturally sterile braided stream deposits that were dated to 15,200 ± 1,500 using optically stimulated luminescence (Goodyear 2005:​106; Waters et al. 2009a, 2009b). Even older components have been subsequently reported from the site (Goodyear 2005). The pre-​Clovis assemblage at Topper appears to be dominated by smashed cores, flakes, and crudely worked unifaces. Excavations at this site are ongoing, and there is no comprehensive report on the pre-Clovis investigations, which makes it nearly impossible to evaluate this assemblage. Page-Ladson is another potential pre-Clovis site that is found in a submerged sinkhole in the Aucilla River of Florida (Webb 2006). As a result of an extraordinary amount of multidisciplinary research, combined with exceptional organic preservation, Dunbar (2006a:414) argues that there is evidence for multiple Paleoindian components at the site. The earliest includes a sample of 11 flakes with seven associated radiocarbon dates that produce an averaged date of 12,425 ± 32 rcybp. However, despite the presence of several bifaces recovered out of context, none were recovered from the excavated deposits. Some other lesser-known pre-Clovis sites include LaGrange Shelter in Alabama and CoatsHines and Johnson in Tennessee. At LaGrange Shelter, debitage and a graver were recovered in a white sand zone below the Dalton component that was dated to 11,290 ± 635 (GX-2774 [DeJarnette and Knight 1976; Futato 1977]). The Johnson site is a multicomponent site that was found eroding into the Cumberland River near Nashville, Tennessee (Barker and Broster 1996; Broster and Norton 1996). In addition to several Early Archaic components, a fluted component was dated from charcoal from a possible hearth (11,780 ± 980, TX-7000; 11,980 ± 110, TX-7454). Finally, the Coats-Hines site contains the remains of a disarticulated mastodon in Late Pleistocene pond deposits along with 34 chert artifacts including a prismatic blade, a bifacial knife, and several other tools (Breitburg et al. 1996; Broster and Norton 1996). This site has produced three radiocarbon 12

Paleoindian Chronology and the Eastern Fluted Point Tradition

dates, the first (6530 ± 70, Beta-75403) from soil and plant remains from the dental cusps and the second (27,050 ± ​200, Beta-80169) from the sedi­ ments below the deposit. Finally, a third date (12,030 ± ​40, Beta-125350) might represent a better approximation of the age of the Paleoindian component (John Broster, personal communication 2011).

sylvania (Gingerich 2007, 2011; Gingerich and ­Waters 2007; McNett 1985), Cactus Hill in Virginia (McAvoy and McAvoy 1997; Wagner and McAvoy 2004), and Sheriden Cave (Redmond and Tankersley 2005; Waters et al. 2009a) and ­Paleo Crossing in Ohio (Brose 1994). The Shawnee-­Minisink site in Pennsylvania is perhaps the best dated site in eastern North America. Five accelerator mass spectrometry dates from three The Paleoindian Period in the separate hearth features associated with debitage Midwest and Mid-Atlantic United States and fluted points provide age estimates between There is much overlap in the Paleoindian chrono- 11,020 and 10,900 14C  bp (Dent 1999; Gingerich logical sequences between the southeastern 2007, and Chapter 9, this volume; Gingerich and United States and the Mid-Atlantic and Midwest- Waters 2007). The Cactus Hill site in Virginia also ern United States. Specifically, in the midconti- dates to 10,920 ± ​250 14C  bp, with assays run on nent and Atlantic Coastal Plain, Dalton bifaces charcoal from a hearth in a level containing fluted are the primary bifaces associated with the Late points and other artifacts (­McAvoy and McAvoy Paleoindian period (Daniel 2001; Gardner 1974, 1997). In the Midwest, the Sheriden Cave site in 1983; Goodyear 1999; Morse 1997). However, the Ohio contains two bone points, a fluted point, and geographic extent of Dalton-type bifaces is un- other stone artifacts. While there are some out­ clear. For example, in a survey of fluted bifaces liers within the dated profile of the cave, the levels from Pennsylvania, Fogelman and Lantz (2006:​ containing early artifacts are bracketed between 30) note that Dalton bifaces are extremely rare. 10,960 and 10,550 14C  bp, with dates of 10,960 ± ​ Like the southeastern United States, the Midwest- 60 to 10,840 ± ​90 14C  bp providing the “acceptable ern and Mid-Atlantic regions are also known for maximum limiting age of the culture-bearing derelatively high densities of fluted bifaces with a posits” (Redmond and Tankersley 2005:514). An variety of types that include Clovis, Gainey, Red- independent assay on collagen from one of the stone, Cumberland, Barnes, Holcombe, and bone points has also provided an age estimate of Crowfield. In most cases, these were recorded by 10,915 ± ​30 14C  Bp (UCIAMS-38249 [Waters et al. private collectors (e.g., Fogelman and Lantz 2006; 2009a]). Finally the Paleo Crossing site in Ohio McCary 1983; Peck 2002). has produced three Clovis-age dates ranging from With the exception of the aforementioned ra- 11,060 ± ​120 to 10,800 ± ​185 14C  bp (Brose 1994). diocarbon dates associated with Dalton compoA few sites in the Mid-Atlantic region have nents, there are only three other dated compo- produced ages that predate the known ages for nents that occur during the span of the Younger Clovis: Cactus Hill, Meadowcroft Rockshelter, Dryas. These include dates from a deeply buried, and Saltville. Excavators of these sites all report and yet to be fully explored, component at the ages several millennia older than Clovis. While Manning site in Ohio (10,240 ± 110, Beta-1774 Saltville has received little acceptance, Cactus Hill [Lepper and Cummings 1993; Maslowski et al. and Meadowcroft remain as plausible c­ ontenders 1995:​​8]), carbonized wood fragments associated for pre-Clovis status (see Fiedel, this volume; with a “Plano-like” point from Smith Mountain in Meltzer 2009). Saltville is known primarily as a Virginia (10,150 ± 70, Beta-93017 [Childress and paleontological locality but has been promoted Blanton 1997:​12]), and debitage and stone tools as a possible pre-Clovis site. The assemblage inat the Barton site in Maryland dated between cludes several pieces of microdebitage and a pos9540 ± ​50 (Beta-203866) and 10,370 ± ​50 (Beta- sible proboscidean bone tool, which produced a 201185 [Wall 2008; R. Wall, personal communi- date of 14,510 ± ​80 from bone collagen (McDoncation 2010]). ald and Kay 1999). There are several sites in the region with CloThe Cactus Hill site in Virginia has produced vis components that have produced ­radiocarbon several radiocarbon ages in artifact-bearing levels dates. These include Shawnee-Minisink in Penn- below the dated Clovis occupation, 10,920 ± ​50 13

Miller and Gingerich

(Beta-81589 [McAvoy and McAvoy 1997; Wagner and McAvoy 2004]). Two charcoal samples from two separate surface hearths with associated artifacts produced radiocarbon assays of 15,070 ± ​70 (Beta-81590) and 16,940 ± ​50 14C  bp (Beta-128330 [Feathers et al. 2006:170; McAvoy and M ­ cAvoy 1997]). Artifacts from these levels consist of quartzite debitage, prismatic blades, and two triangular projectile points (McAvoy and McAvoy 1997). The Meadowcroft Rockshelter is located in Washington County in southwestern Pennsylvania. Archaeological deposits in the rock­shelter range from Paleoindian through historic. ­Fifty-​ two radiocarbon dates suggest good stratigraphic integrity throughout the rockshelter (Adovasio et al. 1990). Although artifacts were found in levels dating to the Clovis era, no fluted or diagnostic Paleoindian artifacts were recovered from the rockshelter. Stratum IIa of the rockshelter has been the focus of much controversy, as this artifact-bearing level has dates that range from 16,175 to 11,300 14C  bp. Unfluted lanceo­late points and other artifacts within Stratum IIa were found within a level bracketed between 12,800 ± ​870 and 11,300 ± ​700 14C  bp (Adovasio et al. 1988; Adovasio et al. 1990). Below this level, artifacts have also been found in levels associated with radiocarbon dates between 16,175 ± ​975 and 12,800 ± ​870 14C  bp (Adovasio et al. 1988). While the antiquity of this site still remains in question (see ­Fiedel, this volume), flaked stone recovered from the earliest levels is unquestionably man-made, and the earliest projectile point (Miller lanceolate) resembles the triangular points from the pre-Clovis levels at Cactus Hill. Despite questions raised by ­Fiedel and others, Cactus Hill and Meadowcroft offer plausible evidence for the earliest occupations in the region. The Paleoindian Period in the Northeastern United States

In the Northeast there are a number of early sites, but few provide dates older than 10,700  14C  bp. Unlike in the Mid-Atlantic and Southeast, there have been no viable claims for pre-Clovis sites in the region. While much of the Northeast was glaciated during the Last Glacial Maximum, paleo­climate research shows that the Laurentide ice sheet had retracted into Canada by 14,000– 14

13,500 cal Bp (Ridge 2003). These data suggest that colonization could have occurred during or well before the Clovis era. But Clovis-age dates, associated with secure archaeological contexts, have yet to be recorded north of New York. Dates from the Hiscock site, 10,990 ± ​100 (TO-3194) and 10,810 ± ​50 14C  bp (CAMS-62560), which date bone tools made from mastodon, appear to be the earliest human-associated dates in the Northeast (Laub 2002). Although the Hiscock dates are early and are associated with fluted points, the approximate time of occupation will likely remain in question until other assays are run on materials or features that are not associated with the mastodon(s), which could have been scavenged from a natural death. Regardless, the fluted point deposit at Hiscock is likely early and provides remarkable evidence of human association with and use of mastodon in eastern North America (also see Anderson 2003:301; Laub and Spiess 2003). The Hiscock points have been interpreted as Gainey (Storck and Holland 2003) and represent what is believed to be one of the earliest fluted forms in the region (Deller and Ellis 1988). Points fitting the description of Clovis appear to be rare or not significantly documented in the Northeast (Bradley et al. 2008). The “Gainey type” also shows considerable variation, and refinement of this typological category is likely needed (Ellis et al. 2003). Using metrics alone, the fluted points at Hiscock most closely resemble those from the Shoop and Lamb sites and various “Gainey” isolates (Ellis et  al. 2003:231). Gainey forms have been argued to be coeval with later Clovis dates and show some overlap in technology that could be indicative of a true Clovis variant. Although researchers admit to the wide variation in western Clovis points (e.g., Deller and Ellis 1992:131), Gainey points and other proposed post-­Clovis forms are different as they often have longer flutes, slightly deeper basal concavities, and occasionally the presence of a small and wide basal thinning flake that is removed over the primary flute (Deller and Ellis 1992:132; for other differences with Clovis, see Bradley et al. 2008). Despite these differences, the overall similarity to Clovis and technological relationships to other dated fluted points in the region form the temporal placement of Gainey (e.g., Deller and Ellis

Paleoindian Chronology and the Eastern Fluted Point Tradition

1988:255; Ellis et al. 2003:229), though few if any good dates actually document the type(s). Dates from the Great Lakes region generally provide a wide range for Gainey occupations (12,360 ± ​1,240–10,370 ± ​108 14C  bp [Storck and Holland 2003:282–283]). Dates for Paleo Crossing in Ohio, however, which has been described as both Gainey and Clovis (Brose 1994; Eren et al. 2004; Storck and Holland 2003; Waters and Stafford 2007), have produced the tightest age range and may represent one of the most secure age estimates for Gainey or early fluted forms, averaging 10,980 ± ​75 14C  bp (Brose 1994). Outside of the Great Lakes region a more recent point typology has been proposed for the Northeast (Bradley et al. 2008). In this scheme researchers have defined Gainey as being similar to what they call Kings Road–Whipple (Bradley et al. 2008). Although this type is not established by radiocarbon dates and is again defined on technological attributes, Kings Road–Whipple is suggested to represent the first occupants in the region. In this typology, Bull Brook–style points, which were once considered similar to and possibly coeval with Gainey (Ellis and Deller 1997; Spiess et al. 1998), are now considered to occur slightly later in the Paleoindian temporal or technological sequence. Dates for Bull Brook have been problematic. Assays run by Byers (1959) and then Grimes (1979) — ​9300 ± ​400, 8720 ± ​400, 6940 ± ​800, 8940 ± ​400, 7590 ± ​255, 5440 ± ​160 — w ​ ere quickly dismissed as too young. Recent work by Robinson et al. (2009) held high hopes for producing credible ages for Bull Brook. Two dates were obtained from calcined caribou bone: 10,410 ± ​60 (Beta-240629) and 10,380 ± ​60 (Beta-240630). Although these dates are consistent with other well-dated fluted point occupations in the Northeast and may fit the Bradley et al. (2008) typology, they could still be too young. As discussed by Robinson et al. (2009:426; Robinson and Ort, this volume), further tests are needed to determine if calcined bone dates and charcoal dates are equivalent. Earlier work on bone dating using both collagen and apatite suggests that bone dates can be 200–300 years younger (e.g., Ambrose and Krigbaum 2003; Surovell 2000); if this study is correct, an appropriate date for Bull Brook may be closer to 10,700 14C  bp. Nevertheless, Gainey,

Kings Road–Whipple, and Bull Brook are considered to be the earliest point types in the region; however, the amount of technological and temporal overlap will eventually have to be tested through secure dating or stratigraphic relationships. The largest number of dates to help establish one hypothesized type in the Northeast comes from the Debert site in Nova Scotia. The Debert site has a total of 28 dates, providing a tight age estimate of 10,600 ± ​47 14C  bp for the occupation of the site (MacDonald 1968). Many of these dates are from feature contexts in direct association with Paleoindian artifacts (MacDonald 1968:52– 56). Although some have questioned whether the Debert dates were obtained from natural features, resulting from forest fires (Bonnichsen and Will 1999), MacDonald (1968:52–53) claims to have been able to distinguish between those features representing hearths and forest fires. M ­ acDonald reports that 14 samples date 10 hearths associated with the Paleoindian occupation. The projectile points from Debert, which are considered a temporal type (Vail-Debert), exhibit very deep basal concavities and are often subtriangular in shape (though probably due to resharpening [see ­Ellis 2004]). Almost exact replicas of the Debert points were found at Vail, which has recently been redated to 10,710 ± ​50 14C  bp (Beta-207579 [R. Gramly, personal communication 2007; Puryear 2009]). This date further supports MacDonald’s dates for Debert and the temporal placement of the Vail-Debert type. After the dates from Vail and Debert, the chronology of northeastern Paleoindian sites becomes even less straightforward, with a significant gap in dated forms. The next recognizable point type is Barnes or Michaud-Neponset, appropriately named after regional sites. ­Michaud-​ Neponset points were named at the Michaud site in Maine, where they are associated with a date of 10,200 ± ​620 14C  bp (Beta-15660 [Spiess and Wilson 1987:84]). Subsequent dates for this point type confirm its age range, with dates from Colebrook, 10,290 ± ​170 (Beta-107429 [Bunker and Potter 1999:70]) and 10,230 ± ​50 14C  bp (Beta258579 [Kitchel and Boisvert 2011]); Neponset, 10,190 ± ​60 14C  bp (Beta-75527 [Curran 1996]); and Templeton, 10,190 ± ​300 (W-3391 [Moeller 1980]) and 10,215 ± ​90 (AA-7160 [McWeeney 1994]). 15

Miller and Gingerich

Michaud-Neponset points are usually medium to large, exhibit fine collateral flaking, are fluted over half the length, and have a moderately deep basal concavity with distinct basal ears (Bradley et al. 2008:141). The distinct style of these points and the aforementioned radiocarbon dates make Michaud-Neponset points one of the best-dated types in the Northeast. They also appear to represent one of the last widespread fluted styles in the region. Crowfield and Cormier-Nicholas points (Bradley et al. 2008), which are assumed to be the latest fluted point forms, are rare in the region and are only documented at a handful of sites. Crowfield points are rather prevalent throughout Ontario and the Great Lakes region but are rare in the Northeast and have a nonpatterned distribution. They have been found at the ­Reagan site, Vermont, and the Trull Farm site, Massachusetts, and there have been reported isolated finds in Maine and New York (Bradley et al. 2008:146– 148). Other isolated finds are known as far south as Pennsylvania and New Jersey (e.g., Fogelman and Lantz 2006; Gingerich, Chapter 6, this volume). A recent Crowfield point base excavated at the Nesquehoning site in Pennsylvania was associated with a radiocarbon date of 9940 ± ​50 14C  bp and may offer an accurate age estimate for this point style in the region (Stewart et  al. 2012). Cormier-Nicholas points are only known from seven sites, with a few known isolated finds in southeastern Massachusetts (for description, see Bradley et al. 2008). Late Paleoindian forms in the region are unfluted and are usually referred to as “Plano” style. These points are well flaked, usually collaterally, and are best described as elongated lanceolate in shape. Under the Plano type, we include both the Agate Basin–related and St. Anne–Varney types, which have recently been described in detail by Bradley et al. (2008). Again, this point style is also considered rare in the region and has not been well dated. The only date associated with the ­Plano-​like forms is from the Varney-style points, which have produced assays ranging from about 9500 to 7000 14C  bp (Bradley et  al. 2008). Despite the uncertainty of the radiocarbon dates, it would appear that these Plano-like forms are younger than 10,000 radiocarbon years and likely date to between 9500 and 9000 14C  bp (­Bradley

et al. 2008:161). It should be noted, however, that the Smith Mountain site in Virginia has produced a Plano-like point that may be associated with a date of 10,200 14C  bp (Childress and Blanton 1997). The geographical extent of Plano-style points is not known, but they appear to be more common in northerly latitudes, becoming rare the farther south one moves, with finds throughout the Great Lakes and the Northeast and only a few documented finds south of Pennsylvania. Dalton points, a marker of the Late Paleoindian period in the Mid-Atlantic and Southeast, are rare in the Northeast. Other types, however, such as Hi-Lo have been argued to be temporal equivalents in the Great Lakes region (Koldehoff and Loebel 2009). The summary of point types provided above is the current working framework used to discuss technological and temporal patterns throughout the Northeast. Over time new point styles will likely be defined, and hopefully these new definitions and refined typologies will come with the discovery of new sites with well-dated components. Until such sites are found, articles like that of Bradley et al. (2008) will go a long way in providing distinct typological categories from which researchers can generate a more meaningful discussion of technology, chronology, and variation in the Paleoindian record. Although limited, sites in the Northeast have produced several radiocarbon assays associated with distinctive fluted point styles. These dates span into the Younger Dryas and show some continuity in the radiocarbon record when compared with the other two regions. Radiocarbon dates of 10,700 Bp from as far north as Maine and the Canadian Maritimes suggest that people moved into the region several ­centuries earlier. Dates from the ­Shawnee-​ Minisink site in northeastern Pennsylvania provide secure evidence of people along the southern boundaries of the Northeast by 11,000 14C  bp. The lack of early dates from the Northeast at this time is unfortunate, but we should not be surprised to eventually find Clovis-age dates associated with the earliest fluted point forms in the region. Discussion

For most of the eastern United States, the time period encompassed by the Younger Dryas marks a time of chronological ambiguity. In the south16

Paleoindian Chronology and the Eastern Fluted Point Tradition

Figure 1.1. Summed probability distribution by region.

eastern United States, the transition from Early Archaic side-notched to Late Paleoindian Dalton assemblages is based on a handful of sites and radiocarbon dates, most of which occur in Missouri and Illinois with the exception of sites such as Dust Cave in Alabama (Goodyear 1982, 1999; Sherwood et al. 2004). This pattern is also consistent with the Midwestern and Mid-­Atlantic regions, where there are several dated Clovis sites but almost none for the remainder of the Younger Dryas. On the other hand, in the northeastern United States, there are multiple fluted point sites that have radiocarbon dates that place them squarely within the Younger Dryas but almost none that predate it, the exception being the Hiscock site. To further illustrate this pattern, we compiled a list of radiocarbon dates from the eastern United States for components older than 8000 14C  bp (see Appendix).2 For sites or components with multiple radiocarbon dates, we averaged the dates using the “R_Combine” function in OxCal 4.1. We then calibrated the dates (again using OxCal 4.1) and placed them in order from youngest to oldest by region (Southeast, Mid-Atlantic/Midwest, and Northeast). For the Southeast, the central tendencies of only four components fall within the Younger Dryas. These include zones T and U from Dust Cave, the side-notched component from Page-Ladson, and a component from 8LE2105, a submerged sinkhole site in Florida. For the

Midwestern/Mid-Atlantic region, only a component from the Manning site in Ohio and the Smith Mountain site in Virginia fall within the Younger Dryas. While the corner-notched component from Thunderbird in Virginia and the side-notched component from St. Albans in West Virginia overlap significantly with the Younger Dryas, this may in large part be due to the large standard deviations of the original dates. Finally, 12 different components from the northeastern United States fall within the Younger Dryas, which is in stark contrast to the southeastern and Midwestern and Mid-Atlantic regions. To further illustrate the variation in the distribution of Paleoindian and Early Archaic radio­ carbon dates across eastern North America, we generated summed probability distributions using CalPal (Weninger et al. 2007; Figure 1.1). We also corrected our probability distribution for ­taphonomic bias, as it is assumed that older dates might be underrepresented due to ­natural destructive processes (Surovell and ­Brantingham 2007; Surovell et al. 2009; Figure 1.2). In both distributions, for the southeastern and Mid-­ Atlantic/​Midwestern samples, the same general pattern emerges, with low probabilities spanning the Younger Dryas. However, in the northeastern sample, the Younger Dryas is marked by a large peak in the probability distribution. Despite the chronological ambiguity for the time encompassing the Younger Dryas in the 17

Miller and Gingerich

Figure 1.2. Summed probabilities corrected for taphonomic bias.

southeastern and Midwestern/Mid-Atlantic regions, many have proposed temporal spans for several different fluted point types based on biface morphology (Anderson 2001; ­Anderson et  al. 1996; Carty and Spiess 1992; Curran 1996; Daniel and Goodyear 2006; Dincauze 1993; ­Ellis 2004; Goodyear 1999, 2006). These generally come in the form of cursory comparisons to some feature or attribute observed on bifaces from dated components from other regions, such as the northeastern or western United States. However, there have been relatively few attempts to quantify variation in the morphology of eastern fluted point types. For example, some approaches utilize cladistic analyses that create “a nested series of taxa based on homologous characters shared by two or more taxa and their immediate common ances-

tor” (O’Brien et al. 2001:1115; see also Buchanan and Collard 2007). Others have applied geometric morphometrics, which involves multivariate statistics that use a set of comparable landmarks to assess the similarities and differences in the overall shape of a sample population, as an alternative approach to cladistic analyses (Buchanan 2006; Ellis 2004; Meltzer and Cooper 2006; Morrow and Morrow 1999; Thulman 2008). One example of using morphology to “crossdate” biface types is the recent attempt to assign the chronological position of the Redstone type relative to other point types in the southeastern United States (Daniel and Goodyear 2006; Goodyear 2006). This type is often separated from the Clovis type because of morphological differences and because of its similarity to bifaces from 18

Paleoindian Chronology and the Eastern Fluted Point Tradition

well-dated, post-Clovis sites in the northeastern United States such as Vail and Debert. Moreover, similarities are also drawn between Redstones and Gainey-type bifaces, which are also undated. As a case study, we used a simplified geometric morphometrics approach to quantitatively assess whether the Clovis and Redstone bifaces are indeed morphologically different types and to what degree they are similar to bifaces from well-dated contexts from outside the southeastern United States. We chose to focus on this particular type due to the pivotal role it plays in recent discussions of possible demographic collapse during the Younger Dryas as a consequence of a proposed extraterrestrial impact (Anderson et  al. 2009; Firestone et al. 2007; Meltzer and Holliday 2010). The primary samples we used are from the North Carolina and South Carolina fluted point surveys. These samples were used for three reasons. First, Goodyear (2006) and Daniel and Goodyear (2006) used the same type designations to distinguish Clovis and Redstone-type bifaces in these two samples. Second, using both of these samples together increases the sample size for statistical analyses. Finally, both of these fluted point surveys are available online through the Paleoindian Database of the Americas project (PIDBA).3 Biface measurements were acquired from three additional sources. The first source is the Northeastern Fluted Point Survey, which is also available from the PIDBA. The second source is the published literature specific to particular sites. The final source data were acquired by measuring casts from the C. Vance Haynes Paleoindian Cast Collection.4 Measurements on a total of 728 bifaces were used in this analysis (Table 1.2). In particular, we chose to analyze the ratio of basal concavity depth to basal width because this measurement is one of the key criteria that Goodyear (2006) and Daniel and Goodyear (2006) used to distinguish Clovis and Redstone bifaces. Moreover, other researchers have also pointed to relative basal concavity depth as varying across time and space in a consistent pattern (­Anderson et  al. 2009; Carty and Spiess 1992; Curran 1996; Daniel and Goodyear 2006; Ellis and Deller 1997; Goodyear 2006; McAvoy 1992; Morrow and Morrow 1999). Since many of the samples in this analysis either suffer from small sample sizes or are not normally distributed, the

Table 1.2. Sample Description.

Sample

State/Province

Northeastern Clovis Paleo Crossing Ohio Shawnee-​ Pennsylvania  Minisink

Reference/ Count Source

26 2

Eren et al. 2004 Gingerich 2007:115

Northeastern Fluted Bull Brook Massachusetts Debert Nova Scotia Gainey Ontario Lamb New York Vail Maine

32 17 14 8 42

Ellis 2004 Ellis 2004 Ellis 2004 Ellis 2004 Ellis 2004

Southeastern Fluted Clovis South Carolina Clovis North Carolina Redstone South Carolina

236 203 49

Redstone

32

Goodyear 2006 Goodyear 2006 Daniel and Goodyear 2006 Daniel and Goodyear 2006

North Carolina

Western Clovis Blackwater New Mexico Draw Dent Colorado Domebo Oklahoma Escapule Arizona Kimmswick Missouri Lehner Arizona Leikem Arizona Murray Arizona  Springs Naco Arizona Total

15

CVH

3 7 2 3 13 1

CVH CVH CVH CVH CVH CVH

15

CVH

8

CVH

728

Note: CVH = C. Vance Haynes Paleoindian Cast Collection.

nonparametric Wilcoxin rank-sum test was used to distinguish whether or not the samples are from the same population. Using this measurement there was a significant difference between the Clovis and Redstone samples from the North Carolina and South Carolina fluted point surveys, albeit with a considerable amount of overlap between the two distri­ butions (z = 9.03; p < .001). Furthermore, there appears to be both a spatial and a ­temporal trend in the distribution of this attribute. First, the sample of Clovis bifaces from the western United States displayed rather minimal expressions of basal indention. This measurement is 19

Miller and Gingerich

also ­statistically consistent with a combined sample of Clovis bifaces from Paleo Crossing in Ohio and ­Shawnee-­​Minisink in Pennsylvania (z = –.63; p = .53), which also overlap temporally with their western counterparts (see ­Waters and Stafford 2007). However, with increasing latitude the samples display an increase in relative basal concavity (also see Curran 1999). This pattern is temporal in nature because those sites from dated contexts in the northeastern United States and southeastern Canada are post-­Clovis in age. This pattern also has implications for the sample of bifaces in the southeastern United States. Specifically, the samples of Clovis bifaces from North and South Carolina are statistically similar to those from dated contexts from the western United States (z = 1.22; p = .22) as well as those from Paleo Crossing and Shawnee-Minisink (z = –.30; p = .76). Meanwhile, the sample of Redstone bifaces is statistically different from western Clovis bifaces (z = –6.11; p < .001), the combined sample of bifaces from Paleo Crossing and Shawnee-­Minisink (z = –3.89; p < .001), and the bifaces typed as Clovis from the North and South Carolina fluted point surveys (z = 8.88; p < .001). Additionally, the Redstone sample is statistically different from the bifaces from Vail (z = 8.88; p < .001), Debert (z = 4.36; p < .001), and Shoop (z = –2.61; p = .009). Instead, this sample is most similar to a sample of bifaces from the Bull Brook site in Massachusetts (z = –.58; p = .56) and a sample of Gainey-type bifaces from Ontario (z = –.19; p = .85). While samples from the Lamb (n = 4) and Hiscock (n = 3) sites suffered from small sample sizes, the mean and median values vary considerably from those in the Redstone sample. Consequently, relative basal concavity depth may provide a useful “first approximation” for cross-dating early fluted point types in the eastern United States.5 For example, bifaces that come from dated Clovis contexts from across North America have relatively minimal basal concavity depth. This observation, coupled with the relatively narrow time span of Clovis dates from across North America (Buchanan et  al. 2008; Faught 2008; Waters and Stafford 2007), appears to support the assertion that Clovis bifaces in the eastern United States are most likely contemporary with those found elsewhere in the Americas.

Relative basal concavity depth also differentiates Clovis from Redstone bifaces, which provides a quantitative basis for arguing that these are indeed separate types. Instead, the Redstone sample falls between the dated sample of ­western and eastern Clovis bifaces (ca. 11,050–10,800 14C  bp) and the samples from Vail (10,710 ± ​43 14C  bp; n = 2) and Debert (10,591 ± ​33 14C  bp; n = 31). If these serve as bounding dates for the Redstone type, the temporal span may have only been on the order of roughly a radiocarbon century. However, the Redstone type is also statistically similar to the Gainey bifaces from the Northeastern Fluted Point Survey and the bifaces from the Bull Brook site. If recent dates on calcined bone from Bull Brook (10,395 ± ​43 14C  bp; n = 2) are accurate, then the temporal span for bifaces with basal concavity depths in the Redstone/Gainey/Bull Brook range may actually be much broader and run counter to the trend of increasing basal concavity depth through time and with increasing latitude. On the other hand, as discussed previously, the dates derived from calcined bone may not be equivalent to those from charcoal and consequently may be several centuries too young. As a result, the infrequent occurrences of Redstone and Gainey bifaces relative to Clovis across eastern North America may be due to the fact that they were produced and discarded over a much shorter period of time. However, this time span, which occurs during the initial part of the Younger Dryas, coincides with a “cliff ” in the calibration curve, and adds further ambiguity to identifying the temporal span of this point type. Consequently, even with more radiocarbon dates, attempts to “pin down” point types produced during this time span may be beyond the current resolution provided by radiocarbon dating. Conclusion

There are many areas where our approach to understanding the chronology of the Paleoindian and Early Archaic periods in eastern North America can be improved. First and foremost would be the discovery of sites comparable to Blackwater Draw, Agate Basin, and Lubbock Lake east of the Mississippi River. While several sites such as Hardaway, Ice House Bottom, Koster, and Dust Cave have components that encompass the Early Archaic period and the latter 20

Paleoindian Chronology and the Eastern Fluted Point Tradition

half of the Younger Dryas, there is still an appreciable gap in the chronology for the initial half of the Younger Dryas, with the exception being the northeastern United States. However, given the amount of research conducted in the eastern United States, opportunities to find archaeological sites with components dating to the Younger Dryas are extremely rare. Instead of waiting for the “holy grail” to appear, an alternative would be to extract as much information from preexisting sites and collections as possible. This could come in the form of reexamining older or minimally published collections from sites, which is the focus of several chapters in this volume. Moreover, redating many sites that were radiocarbon dated prior to the invention of the accelerator mass spectrometer method has been a recent goal of several researchers (Dent 1999; Gingerich 2007, 2011; Haynes et al. 1984; Waters and Stafford 2007) and would continue to be a beneficial avenue of research for the foreseeable future, as well as continued experimentation with alternative forms of deriving absolute dates. However, in many instances, the ability to derive absolute dates from particular sites or regions will be difficult, and archaeologists will instead have to rely on assumptions based on biface morphology. While the preliminary study in this analysis, as well as prior studies by other researchers (e.g., Curran 1999; Ellis and Deller 1997; Morrow and Morrow 1999; Spiess et al. 1998), indicates that relative basal concavity depth may be one useful metric that varies consistently across time and space, it will not be applicable to some Paleo­indian point types in eastern North Amer-

ica, such as Suwannee, Simpson, Cumberland, and Barnes (Michaud-­Neponset), that are undated or show great variation in form. However, the development of multivariate approaches and appropriate reference samples should allow for a more comprehensive assessment of the variation in eastern fluted point types. A second avenue for improving these types of analyses is to standardize the way in which metric attribute data are collected. The measurements used in the analysis were recorded by many people using calipers, which increases the likelihood that interobserver error could influence statistical tests. There are two potential solutions for overcoming this issue. The first would be to create a comparative library of images to use as a basis for statistical comparisons. Morrow and Morrow (1999) measured photographs to increase their sample in regions where they did not have direct access to the bifaces or measurement data. Thulman (2008) used a similar method whereby he photographed bifaces in private collections and used desktop publishing software to quickly and consistently acquire measurements. As a step in creating an online library of artifact images, the Paleoindian Database of the Americas project has now started including artifact images on over 8,000 artifacts that can be related to other contextual information (Anderson et al. 2010). Consequently, expanding data sets such as this, as well as maximizing the research potential from preexisting collections, may provide some of the only tools available for addressing culture-­historical gaps until eastern equivalents of Blackwater Draw are located, excavated, and published.

Notes 4. The C. Vance Haynes Paleoindian Cast Collection is located online at http://www.argonaut.arizona​ .edu/projects/castcollection.htm. 5. The use of basal concavity may be appropriate for early fluted point forms (i.e., Gainey, Kings Road– Whipple, Debert-Vail, Redstone, Bull Brook), but fluted forms such as Michaud-Neponset, ­Cormier-​Nicolas, and Crowfield appear to break the pattern of increased basal concavity.

1. All radiocarbon dates in this chapter are averaged using methods outlined by Long and Rippeteau (1974). 2. This not considered to be an exhaustive list but merely a table of dates from important sites and dated samples that are from contexts that exhibit clear association with artifacts or features. 3. The Paleoindian Database of the Americas is ­located online at http://pidba.utk.edu.

21

Appendix: Select Radiocarbon Dates from Eastern North America Site

State/ Province

Debert

Nova Scotia P.743

Debert

Nova Scotia P.739

Debert

Nova Scotia P.741

Debert

Nova Scotia P.967

Debert

Nova Scotia P.966

Debert

Nova Scotia P.970

Debert

Nova Scotia P.970A

Debert

Nova Scotia P.971

Debert

Nova Scotia P.972

Debert

Nova Scotia P.973

Debert

Nova Scotia P.974

Debert

Nova Scotia P.975

Debert

Nova Scotia P.977

Varney Farm Varney Farm Varney Farm Varney Farm Varney Farm Varney Farm

ME ME ME ME ME ME

Lab #

Beta-81251 Beta-88674 Beta-81250 Beta-93001 Beta-88673 Beta-79658

14C Date

Reference(s)

10,519 ± 143 10,413 ± 167 10,466 ± 167 10,651 ± 179 10,661 ± 173 10,656 ± 134 10,574 ± 180 10,516 ± 177 10,545 ± 126 10,429 ± 320 10,644 ± 308 10,851 ± 341 10,564 ± 163 10,580 ± 163 10,572 ± 121 10,544 ± 159 10,481 ± 161 10,518 ± 120 10,469 ± 179 10,465 ± 144 10,467 ± 118 10,561 ± 367 10,960 ± 377 10,802 ± 375 10,769 ± 374 10,773 ± 226 10,485 ± 152 10,536 ± 151 10,511 ± 120 10,569 ± 179 10,734 ± 185 10,652 ± 114 10,853 ± 168 10,820 ± 110 10,837 ± 114 11,106 ± 311 10,946 ± 307 11,026 ± 225 10,043 ± 371 10,213 ± 376 10,128 ± 275 8380 ± 100 8420 ± 60 8430 ± 100 8620 ± 60 8700 ± 60 9410 ± 190

MacDonald 1968 MacDonald 1968 MacDonald 1968 MacDonald 1968 MacDonald 1968 MacDonald 1968 MacDonald 1968 MacDonald 1968 MacDonald 1968 MacDonald 1968 MacDonald 1968 MacDonald 1968 MacDonald 1968 MacDonald 1968 MacDonald 1968 MacDonald 1968 MacDonald 1968 MacDonald 1968 MacDonald 1968 MacDonald 1968 MacDonald 1968 MacDonald 1968 MacDonald 1968 MacDonald 1968 MacDonald 1968 MacDonald 1968 MacDonald 1968 MacDonald 1968 MacDonald 1968 MacDonald 1968 MacDonald 1968 MacDonald 1968 MacDonald 1968 MacDonald 1968 MacDonald 1968 MacDonald 1968 MacDonald 1968 MacDonald 1968 MacDonald 1968 MacDonald 1968 MacDonald 1968 Inashima 2008; Petersen et al. 2000:127 Inashima 2008; Petersen et al. 2000:127 Inashima 2008; Petersen et al. 2000:127 Inashima 2008; Petersen et al. 2000:127 Inashima 2008; Petersen et al. 2000:127 Inashima 2008; Petersen et al. 2000:127

22

Site

State/ Province

Lab #

14C Date

Reference(s)

Varney Farm Esker

ME ME

Beta-103384

9615 ± 225 10,060 ± 60

Inashima 2008; Petersen et al. 2000:127 Spiess et al. 1998:238

Michaud Michaud Hedden Hedden Vail Vail Vail Vail

ME ME ME ME ME ME ME ME

Beta-15660 Beta-13833 Beta-68806 Beta-70668 AA-114 AA-115 AA-116 C-457

10,200 ± 620 9010 ± 210 10,510 ± 60 10,590 ± 60 10,610 ± 330 10,550 ± 800 10,040 ± 395 14,020 ± 1,120

Vail

ME

Beta-207579

10,710 ± 50

Vail Vail Vail Table Land Weirs Beach Weirs Beach Weirs Beach Whipple Whipple Whipple Whipple Whipple Colebrook Colebrook

ME ME ME NH NH NH NH NH NH NH NH NH NH NH

SI-4617 Beta-1833 AA-117 Beta-71607 GX-4571 GX-5445 GX-4569 average of AA-149 average of AA-150 AA-884 AA-150a AA-150c Beta-107429 Beta-258579

10,300 ± 90 11,120 ± 180 10,460 ± 325 8490 ± 60 8985 ± 210 9155 ± 395 9615 ± 225 9550 ± 320 11,050 ± 300 10,250 ± 260 10,300 ± 500 11,400 ± 360 10,290 ± 170 10,220 ± 50

Saugus Quarry Double P WMECO Charleston Meadows Niponset Bull Brook Bull Brook Templeton (6LF21) Templeton (6LF21) Bolton-Spring Bolton-Spring Johnsen No. 3 Johnsen No. 3 Johnsen No. 3 Johnsen No. 3 Johnsen No. 3 Johnsen No. 3 Johnsen No. 3 Johnsen No. 3 Gardepe Richmond Hill

MA MA MA MA MA MA MA CT CT CT CT NY NY NY NY NY NY NY NY NY NY

I-8256 GX-7508 GX-6986 GX-10925 Beta-75527 Beta-240630 Beta-240629 W-3931 AA-7160 Beta-13177 Beta-1880 GX-8206 GX-8225 GX-8207 GX-8223 GX-8205 GX-8224 GX-8204 GX-9331 Dic-261 I-4929

8095 ± 200 8555 ± 415 8685 ± 370 9120 ± 280 10,210 ± 60 10,380 ± 60 10,410 ± 60 10,190 ± 300 10,215 ± 90 7790 ± 100 10,700 ± 70 8385 ± 230 8585 ± 190 8735 ± 210 8830 ± 210 8880 ± 255 9000 ± 230 9140 ± 260 9665 ± 550 9380 ± 100 9360 ± 120

Spiess and Wilson 1987 Spiess and Wilson 1987 Asch Sidell 1999 Asch Sidell 1999 Haynes et al. 1984 Haynes et al. 1984 Haynes et al. 1984 R. M. Gramly, personal communication 2006 R. M. Gramly, personal communication 2006 Gramly 1982 Gramly 1982 Haynes et al. 1984 Robinson 1996 Bolian 1980:125 Bolian 1980:125 Bolian 1980:124 Haynes et al. 1984 Haynes et al. 1984 Curran 1996 Haynes et al. 1984 Haynes et al. 1984 Bunker et al. 1997 N. R. Kitchel, personal communication 2010 Hoffman 1988:25 Hoffman 1988:25; Thorbahn 1982 Hoffman 1988:25 Hoffman 1987; Hoffman 1988:25 Curran 1996 Robinson et al. 2009 Robinson et al. 2009 Moeller 1980 McWeeney 1994 Calogero and Philpotts 1995:14 Calogero and Philpotts 1995:14 Funk and Wellman 1984:91 Funk and Wellman 1984:91 Funk and Wellman 1984:91 Funk and Wellman 1984:91 Funk and Wellman 1984:91 Funk and Wellman 1984:91 Funk and Wellman 1984:91 Funk and Wellman 1984:91 Funk and Wellman 1984:91 Funk and Wellman 1984:91

23

Appendix: Select Radiocarbon Dates from Eastern North America (continued). Site

State/ Province

Lab #

14C Date

Reference(s)

Arc Arc Hiscock Hiscock Hiscock Hiscock Hiscock Hiscock Hiscock Hiscock Hiscock

NY NY NY NY NY NY NY NY NY NY NY

BGS-1794 BGS-1795 NZA-1106 NZA-1107 NZA-1108 AA-4943 AA-6970 AA-6971 AA-6968 CAMS-6340 CAMS-6341

Tankersley et al. 1997 Tankersley et al. 1997 Laub 2002:108 Laub 2002:108 Laub 2002:108 Laub 2002:108 Laub 2002:108 Laub 2002:108 Laub 2002:108 Laub 2002:108 Laub 2002:108

Turkey Swap West Creek Sheep Rockshelter Sheep Rockshelter Central Builders Sandts Eddy Sandts Eddy Meadowcroft Meadowcroft Meadowcroft Meadowcroft Meadowcroft Meadowcroft Meadowcroft Meadowcroft Meadowcroft Meadowcroft Meadowcroft Meadowcroft Shawnee-Minisink Shawnee-Minisink Shawnee-Minisink Shawnee-Minisink Shawnee-Minisink Shawnee-Minisink Shawnee-Minisink

NJ NJ PA PA PA PA PA PA PA PA PA PA PA PA PA PA PA PA PA PA PA PA PA PA PA PA

DIC-1059 Beta-71577 M-1908 M-1909 A-10053 Beta-53142 Beta-51501 SI-2064 SI-2061 SI-2491 SI-2489 SI-2065 SI-2488 SI-1872 SI-1686 SI-2354 SI-2062 SI-2060 DIC-2187 W-2994 W-3134 W-3388 W-3391 Beta-101935 Beta-127162 Beta-203865

10,360 ± 400 10,375 ± 110 10,850 ± 140 10,420 ± 120 10,220 ± 120 10,465 ± 110 10,945 ± 185 10,545 ± 160 10,705 ± 80 9475 ± 95 9260 ± 70/9150 ± 80 8739 ± 160 9850 ± 165 7050 ± 250 8870 ± 320 9165 ± 210 9300 ± 130 9420 ± 90 8010 ± 110 9075 ± 115 11,300 ± 700 12,800 ± 870 13,240 ± 1010 13,270 ± 340 14,925 ± 620 15,120 ± 165 16,175 ± 975 19,100 ± 810 19,600 ± 2,400 21,070 ± 475 10,590 ± 300 10,750 ± 600 9310 ± 1,000 11,050 ± 1,000 10,940 ± 90 10,900 ± 40 10,820 ± 50

Shawnee-Minisink Shawnee-Minisink Shawnee-Minisink Squaw Rockshelter Squaw Rockshelter Cranberry Prairie Manning Manning Manning Manning

PA PA PA OH OH OH OH OH OH OH

UCIAMS-24865 UCIAMS-24866 OxA-17371 DIC-321B DIC-586 DIC-2555 Beta-23047 Beta-27476 Beta-27477 Beta-1774

10,915 ± 25 11,020 ± 30 10,970 ± 50 7450 ± 85 9480 ± 160 9370 ± 70 9010 ± 120 9720 ± 290 9890 ± 200 10,240 ± 110

24

Cavallo 1981 Stanzeski 1996 Herbstritt 1988:12; Michels and Smith 1967 Herbstritt 1988:12; Michels and Smith 1967 K. Carr, personal communication 2010 Inashima 2008 Inashima 2008 Adovasio et al. 1988 Adovasio et al. 1988 Adovasio et al. 1988 Adovasio et al. 1988 Adovasio et al. 1988 Adovasio et al. 1988 Adovasio et al. 1988 Adovasio et al. 1988 Adovasio et al. 1988 Adovasio et al. 1988 Adovasio et al. 1988 Adovasio et al. 1988 Dent 2002:57; McNett et al. 1985 Dent 2002:57; McNett et al. 1985 Dent 2002:57; McNett et al. 1985 Dent 2002:57; McNett et al. 1985 Dent 1999 Dent 1999 Gingerich 2007; Gingerich and Waters 2007 Gingerich 2007;Gingerich and Waters 2007 Gingerich 2007;Gingerich and Waters 2007 Gingerich 2007, 2011 Maslowski et al. 1995:10 Maslowski et al. 1995:8 Maslowski et al. 1995 Maslowski et al. 1995:9 Maslowski et al. 1995:8 Maslowski et al. 1995:8 Maslowski et al. 1995:8

Site

State/ Province

Lab #

14C Date

Reference(s)

Sheriden Cave

OH

Beta-117604

10,550 ± 70

Sheriden Cave

OH

Beta-117605

10,570 ± 70

Sheriden Cave

OH

Beta-117603

10,600 ± 60

Sheriden Cave

OH

Beta-117606

10,620 ± 70

Sheriden Cave

OH

AA-21710

10,680 ± 80

Sheriden Cave

OH

Beta-127909

10,840 ± 80

Sheriden Cave Sheriden Cave

OH OH

UCIAMS-38249 Beta-127910

10,915 ± 30 10,960 ± 60

Paleo Crossing Paleo Crossing Paleo Crossing Barton Barton Barton Barton Barton West Water Street Amos Power Plant Crosscutter St. Albans St. Albans St. Albans St. Albans St. Albans St. Albans Smith Mountain Smith Mountain Fifty Fifty Thunderbird Cactus Hill

OH OH OH MD MD MD MD MD DE WV WV WV WV WV WV WV WV VA VA VA VA VA VA

AA-8250-E AA-8250-D AA-8250-C Beta-201182 Beta-201183 Beta-201184 Beta-203866 Beta-201185 UGA-516 Beta-94532 Y1540 Y-1539 M-1820 M-1821 Y-1538 M-1827 Beta-77373 Beta-93017 UGA-544 W-3006 W-2816 Beta-81589

10,980 ± 110 10,800 ± 185 11,060 ± 120 7790 ± 50 8390 ± 50 8410 ± 50 9540 ± 50 10,370 ± 50 9430 ± 310 7750 ± 450 7960 ± 60 8160 ± 100 8250 ± 100 8820 ± 500 8830 ± 700 8930 ± 160 9900 ± 500 8810 ± 130 10,150 ± 70 8895 ± 160 9250 ± 300 9900 ± 340 10,920 ± 250

Cactus Hill

VA

Beta-81590

15,070 ± 70

Cactus Hill Cactus Hill Barber Creek Barber Creek Barber Creek Barber Creek Barber Creek Harrison Branch Harrison Branch

VA VA NC NC NC NC NC TN TN

Beta-97708 Beta-128330 Beta-188955 Beta-166239 Beta-150188 Beta-166237 Beta-166238 GX-4120 GX-4119

16,670 ± 730 16,940 ± 50 8950 ± 40 8440 ± 50 8940 ± 70 9280 ± 60 9860 ± 60 7570 ± 250 8545 ± 245

Redmond and Tankersley 2005:503; Waters and Stafford 2007 Redmond and Tankersley 2005:503; Waters and Stafford 2007 Redmond and Tankersley 2005:503; Waters and Stafford 2007 Redmond and Tankersley 2005:503; Waters and Stafford 2007 Redmond and Tankersley 2005:503; Waters and Stafford 2007 Redmond and Tankersley 2005:503; Waters and Stafford 2007 Waters et al. 2009a:109 Redmond and Tankersley 2005:503; Waters and Stafford 2007 Brose 1994 Brose 1994 Brose 1994 Wall 2008 Wall 2008 Wall 2008 Wall 2008 Wall 2008 Custer et al. 1994 Maslowski et al. 1995:10 Maslowski et al. 1995:10 Broyles 1966:27, 40–41 Broyles 1966:27, 40–41 Broyles 1966:25, 40–41 Broyles 1966:25, 40–41 Broyles 1966:19, 40–41 Broyles 1966:18, 40–41 Childress and Blanton 1997:12 Childress and Blanton 1997:12 Carr 1986 Carr 1986 Gardner 1974:5 Feathers et al. 2006; McAvoy and McAvoy 1997 Feathers et al. 2006; McAvoy and McAvoy 1997 Feathers et al. 2006 Feathers et al. 2006 Daniel et al. 2008 Daniel et al. 2008 Daniel et al. 2008 Daniel et al. 2008 Daniel et al. 2008 Chapman 1973, 1976:3–4 Chapman 1973, 1976:3–4

25

Appendix: Select Radiocarbon Dates from Eastern North America (continued). Site

State/ Province

Lab #

14C Date

Reference(s)

Ice House Bottom Ice House Bottom Ice House Bottom Ice House Bottom Ice House Bottom Ice House Bottom Patrick Puckett Puckett Puckett Rose Island Rose Island Rose Island Rose Island Rose Island Rose Island Rose Island Rose Island Widemeier Johnson Johnson Johnson Johnson Johnson Johnson Johnson Johnson Johnson Johnson Johnson Johnson Whalen Whalen Big Bone Lick Cloudsplitter Shelter Cloudsplitter Shelter Cloudsplitter Shelter Deep Shelter Enoch Fork Shelter Enoch Fork Shelter Enoch Fork Shelter Longworth-Gick Longworth-Gick Longworth-Gick Longworth-Gick Main Main Main

TN TN TN TN TN TN TN TN TN TN TN TN TN TN TN TN TN TN TN TN TN TN TN TN TN TN TN TN TN TN TN KY KY KY KY KY KY KY KY KY KY KY KY KY KY KY KY KY

GX-4123 I-9137 I-9138 GX-4127 GX-4125 GX-4126 GX-4122 Tx-7413 Tx-7412 Beta-48045 GX-3563 GX-3169 GX-3598 GX-3168 GX-3167 GX-3597 GX-3565 GX-3564 Beta-234592 Tx-7694 Tx-7693 AA-8860 Tx-7543 Tx-7695 AA-9164 AA-9168 Beta 66202 AA-9165 Tx-7000 Tx-7454 Tx-6999 SFU-221 Beta-15080 W-1358 GX-5874 GX-5785 UCLA-23131 RL-68 Beta-27767 Beta-1524 Beta-27769 TX-2951 TX-3011 UGA-1336 TX-3012 Beta-56433 Beta-5638 Beta-59066

7790 ± 215 8525 ± 355 8715 ± 140 9175 ± 240 9350 ± 215 9435 ± 270 9410 ± 290 8490 ± 180 8820 ± 180 9790 ± 160 8020 ± 190 8060 ± 350 8660 ± 180 8700 ± 300 8800 ± 270 8920 ± 325 9110 ± 145 9330 ± 250 9390 ± 50 8810 ± 80 8830 ± 170 8925 ± 75 8940 ± 110 8980 ± 90 9050 ± 85 9090 ± 85 9510 ± 250 9555 ± 90 11,700 ± 980 11,980 ± 110 12,660 ± 970 8500 ± 460 9340 ± 100 10,600 ± 200 8200 ± 225 9215 ± 290 9228 ± 100 8520 ± 470 8740 ± 130 10,960 ± 240 13,480 ± 350 8420 ± 110 8440 ± 380 8440 ± 125 9490 ± 230 8030 ± 100 8190 ± 90 8190 ± 110

Chapman 1973, 1976:3–4 Chapman 1973, 1976:3–4 Chapman 1973, 1976:3–4 Chapman 1973, 1976:3–4 Chapman 1973, 1976:3–4 Chapman 1973, 1976:3–4 Chapman 1973, 1976:3–4 Norton and Broster 1993 Norton and Broster 1993 Norton and Broster 1993 Chapman 1976:3–4 Chapman 1976:3–4 Chapman 1976:3–4 Chapman 1976:3–4 Chapman 1976:3–4 Chapman 1976:3–4 Chapman 1976:3–4 Chapman 1976:3–4 Broster et al. 2008 Barker and Broster 1996:98 Barker and Broster 1996:98 Barker and Broster 1996:98 Barker and Broster 1996:98 Barker and Broster 1996:98 Barker and Broster 1996:98 Barker and Broster 1996:98 Barker and Broster 1996:98 Barker and Broster 1996:98 Barker and Broster 1996:98 Barker and Broster 1996:98 Barker and Broster 1996:98 Maslowski et al. 1995:9 Maslowski et al. 1995:8 Tankersley 1985:41, 1990:82 Maslowski et al. 1995:9 Maslowski et al. 1995:8 Maslowski et al. 1995:8 Maslowski et al. 1995:9 Maslowski et al. 1995:9 Maslowski et al. 1995:8 Maslowski et al. 1995:8 Maslowski et al. 1995:9 Maslowski et al. 1995:9 Maslowski et al. 1995:9 Maslowski et al. 1995:8 Maslowski et al. 1995:10 Maslowski et al. 1995:9 Maslowski et al. 1995:9

26

Site

State/ Province

Lab #

14C Date

Reference(s)

Main Main Morrissroe Big Eddy Big Eddy Big Eddy Big Eddy Big Eddy Big Eddy Big Eddy Big Eddy Big Eddy Big Eddy Big Eddy Big Eddy Big Eddy Big Eddy Big Eddy Big Eddy Big Eddy Big Eddy Big Eddy Big Eddy Big Eddy Big Eddy Gregg Shoals Gregg Shoals Gregg Shoals Phinizy Swamp Rae’s Creek Rae’s Creek Rae’s Creek Dust Cave Dust Cave Dust Cave Dust Cave Dust Cave Dust Cave Dust Cave Dust Cave Dust Cave Dust Cave Dust Cave Dust Cave Dust Cave Dust Cave Dust Cave Dust Cave

KY KY KY MO MO MO MO MO MO MO MO MO MO MO MO MO MO MO MO MO MO MO MO MO MO GA GA GA GA GA GA GA AL AL AL AL AL AL AL AL AL AL AL AL AL AL AL AL

Beta-59067 Beta-5634 SFU-271 Beta-112982 AA-35461 Tx-9329 AA-27479 AA-35365 AA-35462 AA-26653 AA-25778 Tx-9325 AA-27480 AA-27487 AA-29022 AA-27488 AA-29021 AA-26654 Tx-9327 AA-27481 AA-27482 AA-27485 AA-35489 Tx-9326 AA-27486 – – – – Beta-35187 Beta-35233 Beta-35235 Beta-65178 Beta-48756 Beta-65184 Beta-147136 Beta-81608 Beta-81611 Beta-133788 Beta-65177 Beta-147132 Beta-81610 Beta-133791 Beta-147135 Beta-133790 Beta-65181 Beta-41063 Beta-81609

8450 ± 120 8500 ± 70 8220 ± 100 9190 ± 90 9270 ± 240 9450 ± 61 9525 ± 65 9800 ± 1,100 9835 ± 70 10,185 ± 75 10,260 ± 85 10,336 ± 110 10,340 ± 100 10,400 ± 75 10,430 ± 70 10,470 ± 80 10,680 ± 60 10,710 ± 85 11,076 ± 86 11,160 ± 75 11,190 ± 75 11,280 ± 75 11,375 ± 80 11,384 ± 107 11,900 ± 80 10,000 ± 140 10,170 ± 140 10,370 ± 140 8953 ± 51 7570 ± 130 8370 ± 280 9060 ± 110 7040 ± 110 7680 ± 170 8330 ± 170 8830 ± 50 8470 ± 50 9890 ± 70 9950 ± 50 9990 ± 140 10,010 ± 40 10,070 ± 70 10,100 ± 50 10,140 ± 40 10,310 ± 60 10,310 ± 230 10,330 ± 120 10,340 ± 130

Maslowski et al. 1995:9 Maslowski et al. 1995:9 Maslowski et al. 1995:9 Hajic et al. 1998:91; Hajic et al. 2000:31 Hajic et al. 1998:91; Hajic et al. 2000:31 Hajic et al. 1998:91; Hajic et al. 2000:31 Hajic et al. 1998:91; Hajic et al. 2000:31 Hajic et al. 1998:91; Hajic et al. 2000:31 Hajic et al. 1998:91; Hajic et al. 2000:31 Hajic et al. 1998:91; Hajic et al. 2000:31 Hajic et al. 1998:91; Hajic et al. 2000:31 Hajic et al. 1998:91; Hajic et al. 2000:31 Hajic et al. 1998:91; Hajic et al. 2000:31 Hajic et al. 1998:91; Hajic et al. 2000:31 Hajic et al. 1998:91; Hajic et al. 2000:31 Hajic et al. 1998:91; Hajic et al. 2000:31 Hajic et al. 1998:91; Hajic et al. 2000:31 Hajic et al. 1998:91; Hajic et al. 2000:31 Hajic et al. 1998:91; Hajic et al. 2000:31 Hajic et al. 1998:91; Hajic et al. 2000:31 Hajic et al. 1998:91; Hajic et al. 2000:31 Hajic et al. 1998:91; Hajic et al. 2000:31 Hajic et al. 1998:91; Hajic et al. 2000:31 Hajic et al. 1998:91; Hajic et al. 2000:31 Hajic et al. 1998:91; Hajic et al. 2000:31 Goodyear 1999:453 Goodyear 1999:453 Goodyear 1999:453 Elliott et al. 1994 Crook 1990:126 Crook 1990:126 Crook 1990:124, 126 Sherwood et al. 2004:538 Sherwood et al. 2004:538 Sherwood et al. 2004:538 Sherwood et al. 2004:538 Sherwood et al. 2004:538 Sherwood et al. 2004:539 Sherwood et al. 2004:539 Sherwood et al. 2004:539 Sherwood et al. 2004:539 Sherwood et al. 2004:539 Sherwood et al. 2004:539 Sherwood et al. 2004:539 Sherwood et al. 2004:539 Sherwood et al. 2004:539 Sherwood et al. 2004:539 Sherwood et al. 2004:538

27

Appendix: Select Radiocarbon Dates from Eastern North America (continued). Site

State/ Province

Lab #

14C Date

Reference(s)

Dust Cave Dust Cave Dust Cave Dust Cave Dust Cave Dust Cave Russell Cave Russell Cave Russell Cave Stanfield-Worley Stanfield-Worley Stanfield-Worley Stanfield-Worley Stanfield-Worley Lagrange Shelter Windover Windover Windover Little Salt Spring Little Salt Spring Little Salt Spring Little Salt Spring Little Salt Spring Little Salt Spring Little Salt Spring Little Salt Spring Little Salt Spring Little Salt Spring Little Salt Spring Little Salt Spring Little Salt Spring 8LE2105 8LE2105 8LE2105 Page-Ladson

AL AL AL AL AL AL AL AL AL AL AL AL AL AL AL FL FL FL FL FL FL FL FL FL FL FL FL FL FL FL FL FL FL FL FL

Beta-40680 Beta-100506 Beta-65179 Beta-40681 Beta-81613 Beta-81599 I-828 I-822 I-2239 M-1153 M-1348 M-1347 M-1346 M-1152 GX-2774 TO-518 Beta-18295 TO-241 UM-1100 I-6458 I-6512 Tx-2594 UM-1101 Tx-2460 I-6460 Beta-36591 Tx-2461 Tx-2595 I-6459 Tx-2636 Tx-2335 Beta-81467 Beta-81468 Beta-81469 Beta-15089

10,345 ± 80 10,370 ± 180 10,390 ± 80 10,490 ± 360 10,490 ± 60 10,500 ± 60 8145 ± 275 8485 ± 275 8550 ± 320 8920 ± 400 9040 ± 400 9340 ± 400 9440 ± 400 9640 ± 450 11,290 ± 635 7830 ± 80 7930 ± 80 8120 ± 70 8145 8455 8955 9080 9100 9500 9645 9660 9920 10,190 10,980 12,030 13,450 9850 ± 50 9900 ± 60 10,090 ± 70 9450 ± 100

Page-Ladson

FL

Beta-11905

9730 ± 120

Page-Ladson

FL

Beta-58858

9930 ± 60

Page-Ladson

FL

Beta-103888

9950 ± 70

Page-Ladson

FL

Beta-21750

10,000 ± 120

Page-Ladson

FL

Beta-58857

10,000 ± 80

Page-Ladson

FL

Beta-21752

10,280 ± 110

Sherwood et al. 2004:538 Sherwood et al. 2004:538 Sherwood et al. 2004:538 Sherwood et al. 2004:538 Sherwood et al. 2004:539 Sherwood et al. 2004:539 Futato 1977:39; Griffin 1974 Futato 1977:39; Griffin 1974 Futato 1977:39; Griffin 1974 Futato 1977:39; Griffin 1974 DeJarnette et al. 1962:85–87 DeJarnette et al. 1962 DeJarnette et al. 1962 DeJarnette et al. 1962 DeJarnette and Knight 1976:38 Doran and Dickel 1988:367 Doran and Dickel 1988:367 Doran and Dickel 1988:367 Clausen et al. 1979:611 Jones and Tesar 2000 Clausen et al. 1979:611 Clausen et al. 1979:611 Clausen et al. 1979:611 Clausen et al. 1979:611 Clausen et al. 1979:611 Jones and Tesar 2000 Clausen et al. 1979:611 Clausen et al. 1979:611 Clausen et al. 1979:611 Clausen et al. 1979:611 Clausen et al. 1979:611 Faught et al. 2003:17 Faught et al. 2003:17 Faught et al. 2003:17 Dunbar et al. 1988:449; Dunbar et al. 1989:477–482 Dunbar et al. 1988:449; Dunbar et al. 1989:477–482 Dunbar et al. 1988:449; Dunbar et al. 1989:477–482 Dunbar et al. 1988:449; Dunbar et al. 1989:477–482 Dunbar et al. 1988:449; Dunbar et al. 1989:477–482 Dunbar et al. 1988:449; Dunbar et al. 1989:477–482 Dunbar et al. 1988:449; Dunbar et al. 1989:477–482

Note: Each sample at Debert was counted twice and then averaged for the date (the last date in the sequence in boldface equals the average).

28

Paleoindian Chronology and the Eastern Fluted Point Tradition

References Cited

Anderson, D. G., Albert C. Goodyear, Thomas W. Stafford, Jr., James Kennett, and Allen West 2009 Potential Human Population Decline/Reorganization During the Younger Dryas in North America. Poster presented at the American Geophysical Union Fall Meeting, December 16, San Francisco. Anderson, David G., D. Shane Miller, Stephen J. Yerka, and Michael K. Faught 2005 Paleoindian Database of the Americas: 2005 Status Report. Current Research in the Pleistocene 22:91–92. Anderson, David G., D. Shane Miller, Stephen J. Yerka, J. Christopher Gillam, Erik N. Johanson, Derek T. Anderson, Albert C. Goodyear, and Ashley M. Smallwood 2010 PIDBA (Paleoindian Database of the ­Americas) 2010: Current Status and Findings. Archaeology of Eastern North America 38:63–90. Anderson, David G., L. D. O’Steen, and K. E. Sassaman 1996 Environmental and Chronological Considerations. In The Paleoindian and Early Archaic Southeast, edited by David G. Anderson and Kenneth E. Sassaman, pp. 3–15. University of Alabama Press, Tuscaloosa. Asch Sidell, Nancy 1999 Prehistoric Plant Use in Maine: Paleo­indian to Contact Period. In Current Northeast Paleoethnobotany, edited by John P. Hart, pp. 191–223. Bulletin No. 494. New York State Museum, Albany. Barker, G., and J. B. Broster 1996 The Johnson Site (40Dv400): A Dated Paleo­ indian and Early Archaic Occupation in Tennessee’s Central Basin. Journal of Alabama Archaeology 42:97–153. Beck, C., and G. T. Jones 2010 Clovis and Western Stemmed: Population Migration and the Meeting of Two Technologies in the Intermountain West. American Antiquity 75(1):81–116. Boldurian, Anthony T., and John L. Cotter 1999 Clovis Revisited: New Perspectives on Paleo­ indian Adaptations from Blackwater Draw, New Mexico. University Museum Monograph 103. University Museum, University of Pennsylvania, Philadelphia. Bolian, Charles E. 1980 The Early and Middle Archaic of the Lakes Region, New Hampshire. In Early and Middle Archaic Cultures in the Northeast, edited by David R. Starbuck and Charles E. ­Bolian, pp. 115–134. Occasional Publications in

Adovasio, J. M., A. T. Bouldurin, and R. C. Carlisle 1988 Who Are Those Guys? Some Biased Thoughts on the Initial Peopling of the New World. In Americans Before Columbus: Ice Age ­Origins: Selected Papers from the First C ­ olumbian Quincentenary Symposium, edited by R. C. Carisle, pp. 45–62. Department of Anthropology, Ethnology Monographs No. 12. University of Pittsburg, ­Pittsburg. Adovasio, J. M., J. D. Gunn, J. Donahue, and R. Stuckenrath 1990 The Meadowcroft Rockshelter Radiocarbon Chronology 1975–1990. American Antiquity 55:348–354. Ambrose, Stanley, and John Krigbaum 2003 Bone Chemistry and Bioarchaeology. Journal of Anthropological Archaeology 22:193–199. Anderson, David G. 1990 Paleoindian Colonization of Eastern North America: A View from the Southeastern United States. In Early Paleoindian Economies of Eastern North America, edited by Kenneth Tankersley and Barry Isaac, pp. 163– 216. Research in Economic Anthropology Supplement 5. JAI Press Inc., Greenwich, Connecticut. 1995 Paleoindian Interaction Networks in the Eastern Woodlands. In Native American Interactions: Multiscalar Analysis and Interpretations in the Eastern Woodlands, edited by Michael S. Nassaney and Kenneth E. Sassaman, pp. 3–26. University of Tennessee Press, Knoxville. 1996 Models of Paleoindian and Early Archaic Settlement in the Lower Southeast. In The Paleo­ indian and Early Archaic Southeast, edited by David G. Anderson and Kenneth Sassaman, pp. 29–57. University of Alabama Press, Tuscaloosa. 2001 Climate and Culture Change in Prehistoric and Early Historic Eastern North America. Archaeology of Eastern North America 29:​ 143–186. 2003 The Hiscock Site Archaeological Record. In The Hiscock Site: Late Pleistocene and Holocene Paleoecology and Archaeology of Western New York State, edited by Richard S. Laub, pp. 300–​303. Bulletin of the Buffalo Society of Natural Sciences 37. Buffalo. Anderson, D. G., and M. K. Faught 1998 The Distribution of Fluted Paleoindian Projectile Points: Update. Archaeology of Eastern North America 26:163–187. 29

Miller and Gingerich (46KA27), Kanawha County. West Virginia Archaeologist 19:1–43. Buchanan, Briggs 2003 The Effects of Sample Bias on Paleoindian Fluted Point Recovery in the United States. North American Archaeologist 24(4):311–338. 2006 An Analysis of Folsom Projectile Point Resharpening Using Quantitative Comparisons of Form and Allometry. Journal of Archaeological Science 33:185–199. Buchanan, Briggs, and Mark Collard 2007 Investigating the Peopling of the Americas Through Cladistic Analysis of Early Paleo­ indian Projectile Points. Journal of Anthropological Archaeology 26:366–393. Buchanan, Briggs, Mark Collard, and Kevan Edinborough 2008 Paleoindian Demography and the Extraterrestrial Impact Hypothesis. Proceedings of the National Academy of Sciences 105(33):11651– 11654. Bullen, Ripley P. 1975 A Guide to the Identification of Florida Projectile Points. Kendall Books, Gainesville, Florida. Bunker, Victoria, Edna Feighner, and Jane Potter 1997 Technical Report Archeological Resources Phase I-B Preliminary Archeological Assessment and Phase II Intensive Survey. Portland Natural Gas Transmission System Northern New Hampshire Revision Route, M.P. 0.0– 68.6, FERC Docket No. CP 96-249-003. Report on file at the New Hampshire Division of Historical Resources, Concord. Bunker, Victoria, and Jane Potter 1999 Early Occupation in the Far Upper Connecticut River Valley. New Hampshire Archeologist 39(1):70–81. Byers, Douglas S. 1959 Radiocarbon Dates for the Bull Brook Site, Massachusetts. American Antiquity 24(4):​ 427–429. Calogero, B. L. A., and A. R. Philpotts 1995 Rocks and Minerals Used by Tool Knappers in New England. Northeast Anthropology 50:1–17. Cambron, James W., and David C. Hulse 1960 An Excavation on the Quad Site. Tennessee Archaeologist 16:14–26. 1975 Handbook of Alabama Archaeology, Pt. I: Point Types. Revised Edition. Archaeological Research Association of Alabama, Tuscaloosa. Carr, Kurt W. 1986 Core Reconstruction and Community Patterning at the Fifty Site. Journal of Middle ­Atlantic Archaeology 2:79–92.

Northeastern Anthropology, Vol. 7. Department of Anthropology, Franklin Pierce College, Rindge, New Hampshire. Bonnichsen, R., and R. F. Will 1999 Radiocarbon Chronology of Northeastern Paleoamerican Sites; Discriminating Natural and Human Burn Features. In Ice Age Peoples of North America: Environments, Origins, and Adaptations of the First Americans, edited by R. Bonnichsen and K. L. Turnmire, pp. 395– 415. Center for the Study of the First Americans, Oregon State University, Corvallis. Bradley, Bruce, Michael B. Collins, and Andrew Hemmings 2010 Clovis Technology. Archaeological Series 17. International Monographs in Prehistory, Ann Arbor. Bradley, J. W., A. E. Spiess, R. A. Boisvert, and J. ­Broudreau 2008 What’s the Point? Modal Forms and Attributes of Paleoindian Bifaces in the New ­England–​Maritimes Region. Archaeology of Eastern North America 36:119–172. Breitburg, E., J. B. Broster, A. L. Reesman, and R. G. Stearns 1996 The Coats-Hines Site: Tennessee’s First ­Paleoindian–Mastodon Association. Current Research in the Pleistocene 13:6–8. Brennan, Louis A. 1982 A Compilation of Fluted Points of Eastern North America by Count and Distribution: An AENA Project. Archaeology of Eastern North America 10:27–46. Brose, D. S. 1994 Archaeological Investigations at the Paleo Crossing Site, a Paleoindian Occupation in Medina County, Ohio. In The First Discovery of North America: Archaeological Evidence of the Early Inhabitants of the Ohio Area, edited by W. S. Dancey, pp. 61–76. Ohio Archaeological Council, Columbus. Broster, John B., and Mark R. Norton 1996 Recent Paleoindian Research in Tennessee. In The Paleoindian and Early Archaic Southeast, edited by D. G. Anderson and K. E. Sassaman, pp. 288–297. University of Alabama Press, Tuscaloosa. Broster, John B., Mark R. Norton, Bobby Hulan, and Ellis Durham 2008 Paleoamerican and Early Archaic Occupations of the Widimeier Site (40Dv9), Davidson County, Tennessee. Current Research in the Pleistocene 25:64–65. Broyles, B. J. 1966 Preliminary Report: The St. Albans Site 30

Paleoindian Chronology and the Eastern Fluted Point Tradition scendants. In The Archaeological Northeast, edited by M. A. Levine, K. E. Sassaman, and M. S. Nassaney, pp. 3–24. Bergin and Garvey, Westport. Custer, Jay F., Lynn Riley, and Glen Mellin 1994 Final Archaeological Excavations at the Leipsic Site (7KC-194A), State Route 1 Corridor, Kent County, Delaware. Submitted to the Delaware Department of Transportation by the University of Delaware Department of Anthropology Center for Archaeological Research, Newark. Daniel, I. R., Jr. 2001 Stone Raw Material Variability and Early Archaic Settlement in the Southeastern United States. American Antiquity 66:237–265. Daniel, I. R., Jr., and Albert C. Goodyear 2006 An Update on the North Carolina ­Fluted-​ Point Survey. Current Research in the Pleistocene 23:99–101. Daniel, I. R., Jr., Keith C. Seramur, Tara L. Potts, and Matthew W. Jorgenson 2008 Searching a Sand Dune: Shovel Testing the Barber Creek Site. North Carolina Archaeology 57:1–28. Daniel, I. R., Jr., and M. Wisenbaker 1987 Harney Flats: A Florida Paleoindian Site. Baywood, New York. DeJarnette, D. L., and V. J. Knight 1976 LaGrange. Journal of Alabama Archaeology 12:1–60. DeJarnette, D. L., E. Kurjack, and J. Cambron 1962 Excavations at the Stanfield-Worley Bluff Shelter. Journal of Alabama Archaeology ­8(1–2):1–124. Deller, D. Brian, and C. J. Ellis 1988 Early Paleo-Indian Complexes in Southwestern Ontario. In Late Pleistocene and Early Holocene Paleoecology and Archaeology of the Eastern Great Lakes Region, edited by R. S. Laub, N. G. Miller, and D. W. Steadman, Vol. 33:​251–263. Buffalo Society of Natural Sciences, Buffalo. 1992 Thedford II: A Paleo-Indian Site in the ­Ausable River Watershed of Southwestern ­Ontario. Memoirs Series, Museum of Anthropology, University of Michigan, No. 24. Ann Arbor. Dent, Richard J. 1999 Shawnee Minisink: New Dates on the Paleo­ indian Component. Poster presented at the 64th Annual Meeting of the Society for American Archaeology, Chicago. 2002 Paleoindian Occupation of the Upper Delaware Valley: Revisiting Shawnee Minisink

Carty, Frederick M., and Arthur E. Spiess 1992 The Neponset Paleoindian Site in Massachusetts. Archaeology of Eastern North America 20:19–37. Cavallo, J. 1981 Turkey Swamp: A Late Paleo-Indian Site in New Jersey’s Coastal Plain. Archaeology of Eastern North America 9:1–18. Chapman, Jefferson 1973 The Icehouse Bottom Site 40MR23: Report of Investigations No. 13. Department of Anthropology, University of Tennessee, Knoxville. 1976 The Archaic Period of the Lower Little Tennessee River Valley: The Radiocarbon Dates. Tennessee Anthropologist 1:1–12. Childress, W., and D. B. Blanton 1997 A Radiocarbon Date from a Deeply Buried Stratum Yielding a Plano-Like Projectile Point from the Smith Mountain Site in Virginia. Current Research in the Pleistocene 14:12–14. Clausen, C. J., A. D. Cohen, C. Emiliani, J. A. Holman, and J. J. Stipp 1979 Little Salt Spring, Florida: A Unique Underwater Site. Science 203:609–614. Coleman, D. D. 1972 Illinois State Geological Survey Radiocarbon Dates III. Radiocarbon 14(1):149–154. Collins, Michael B. (editor) 1998 Wilson-Leonard, an 11,000-Year Archeological Record of Hunter-Gatherers in Central Texas, Vols. I–V. University of Texas, Austin. Crane, H. R., and J. B. Griffin 1956 University of Michigan Radiocarbon Dates I. Science 124:664–672. 1968 University of Michigan Radiocarbon Dates XII. Radiocarbon 10:61–114. 1972 University of Michigan Radiocarbon Dates XIV. Radiocarbon 14:155–194. Crook, Morgan Ray, Jr. 1990 Rae’s Creek: A Multicomponent Archaeological Site at the Fall Line Along the Savannah River. Department of Anthropology, Georgia State University, Atlanta. Curran, M. L. 1984 The Whipple Site and Paleoindian Tool Assemblage Variation: A Comparison of Intra­ site Structuring. Archaeology of Eastern North America 12:5–40. 1996 Paleoindians in the Northeast: The Problem of Dating Fluted Point Sites. Review of Archaeology 17(1):2–11. 1999 Exploration, Colonization and Settling In: The Bull Brook Phase, Antecedents, and De31

Miller and Gingerich Georgia. Southeastern Archeological Services, Inc., Athens, Georgia. Ellis, C. J. 2004 Understanding “Clovis” Fluted Point Variability in the Northeast: A Perspective from the Debert Site, Nova Scotia. Canadian Journal of Archaeology 28(2):205–253. Ellis, C. J., and D. B. Deller 1997 Variability in the Archaeological Record of Northeastern Early Paleoindians: A View from Southern Ontario. Archaeology of Eastern North America 25(1):1–30. Ellis, C. J., J. Tomenchuk, and J. Holland 2003 Typology, Use and Sourcing of the Late Pleistocene Lithic Artifacts from the Hiscock Site. In The Hiscock Site: Later Pleistocene and Early Holocene Paleoecology and ­Archaeology of Western New York State, edited by R. S. Laub, pp. 221–237. Bulletin of the Buffalo ­Society of Natural Sciences 37. Buffalo. Eren, Metin I., Brian G. Redmond, and Mark A. Kollecker 2004 The Paleo-Crossing (33-ME-274) Fluted Point Assemblage. Current Research in the Pleistocene 21:38–39. Faught, Michael K. 2008 Archaeological Roots of Human Diversity in the New World: A Compilation of Accurate and Precise Radiocarbon Ages from the ­Earliest Sites. American Antiquity 73(4):670– 698. Faught, Michael K., Michael B. Hornum, R. Christopher Goodwin, Brinnen Carter, and S. David Webb 2003 The Earliest-Holocene Tool Assemblages from Northern Florida with Stratigraphically Controlled Radiocarbon Estimates (Sites 8LE2105 and 8JE591). Current Research in the Pleistocene 20:16–18. Feathers, J. K., E. J. Rhodes, S. Huot, and J. M. McAvoy 2006 Luminescence Dating of Sand Deposits Related to Late Pleistocene Occupation at the Cactus Hill Site, Virginia, USA. Quaternary Geochronology 1:167–187. Firestone, R. B., A. West, J. P. Kennett, L. Becker, T. E. Bunch, Z. S. Revay, P. H. Schultz, T. Belgya, D. J. Kennett, J. M. Erlandson, O. J. Dickenson, A. C. Goodyear, R. S. Harris, G. A. Howard, J. B. Kloosterman, P. Lechler, P. A. Mayewski, J. Montgomery, R. Poreda, T. Darrah, S. S. Que Hee, A. R. Smith, A. Stich, W. Topping, J. H. Wittke, and W. S. Wolbach 2007 Evidence for an Extraterrestrial Impact 12,900 Years Ago That Contributed to the Megafaunal Extinctions and the Younger Dryas Cooling. Proceeding of the National Academy of Science 104(41):16016–16021.

and Nearby Sites. In Ice Age Peoples of Pennsylvania, edited by K. W. Carr and James M. Adovasio, pp. 51–78. Pennsylvania Historical and Museum Commission, Harrisburg. Dincauze, Dena 1993 Fluted Points in Eastern Forests. In From Kostenki to Clovis: Upper Paleolithic–­ Paleoindian Adaptations, edited by Olga ­Soffer and N. D. Prazlov, pp. 279–292. Plenum Press, New York. Doran, G. H., and D. N. Dickel 1988 Radiometric Chronology of the Archaic Windover Archaeological Site (8 Br 246). Florida Anthropologist 41:365–380. Dragoo, Donald W. 1965 Investigations at a Paleo-Indian Site in Stewart County, Tennessee. Eastern States Archeological Federation Bulletin 24:12–13. 1973 Wells Creek: An Early Man Site in Stewart County, TN. Archaeology of Eastern North America 1:1–56. Dunbar, James S. 2006a Paleoindian Archaeology. In First Floridians and Last Mastodons: The Page-Ladson Site in the Aucilla River, edited by S. David Webb, pp. 403–435. Springer, Dordrecht. 2006b Pleistocene–Early Holocene Climate Change: Chronostratigraphy and Geoclimage of the Southeast U.S. In First Floridians and Last Mastodons: The Page-Ladson Site in the Aucilla River, edited by S. David Webb, pp. 103–158. Springer, Dordrecht. Dunbar, J. S., M. K. Faught, and S. D. Webb 1988 Page/Ladsen (8JE591): An Underwater Paleo­ indian Site in Northwestern Florida. Florida Anthropologist 41(3):442–452. Dunbar, J. S., S. D. Webb, and D. Cring 1989 Culturally and Naturally Modified Bones from a Paleoindian Site in the Aucilla River, North Florida. In Bone Modification, edited by R. Bonnichsen and M. H. Sorg, pp. 473– 497. Center for the Study of Early Man, University of Maine, Orono. Dunnell, Robert C. 1990 The Role of the Southeast in American Archaeology. Southeastern Archaeology 9:11–22. Ellerbusch, Elijah C., and Stephen J. Yerka 2005 Cumberland and Barnes: Analogous Evolution or Historically Contingent Homology? Paper presented at the Clovis in the Southeast Conference, Columbia, South Carolina. Elliott, Daniel T., R. Jerald Ledbetter, and E. A. Gordon 1994 Data Recovery at Lover’s Lane, Phinizy Swamp, and the Old Dike Sites Bobby Jones Expressway Extension Corridor Augusta, 32

Paleoindian Chronology and the Eastern Fluted Point Tradition Evidence for Pre-Clovis Sites in the Eastern United States. In Paleoamerican Origins: Beyond Clovis, edited by R. Bonnichsen, B. Bradley, D. Stanford, and M. Waters, pp. 103–112. Center for the Study of the First Americans, Texas A&M University, College Station. 2006 Recognizing the Redstone Fluted Point in the South Carolina Paleoindian Point Database. Current Research in the Pleistocene 23:112–114. Goodyear, Albert C., and Tommy Charles 1984 An Archaeological Survey of Chert Quarries in Western Allendale County, South Carolina. Research Manuscript Series 195. South Carolina Institute of Archaeology and Anthropology, Columbia. Goodyear, Albert C., D. Shane Miller, and Ashley M. Smallwood 2007 Introducing Clovis at the Topper Site, 38AL23, Allendale County, South Carolina. Paper presented at the 72nd Annual ­Meeting of the Society for American Archaeology, April 27, Austin. Goodyear, Albert C., III, and Ken Steffy 2003 Evidence of a Clovis Occupation at the Topper Site, 38AL23, Allendale County, South Carolina. Current Research in the Pleistocene 20:23–25. Gramly, Richard Michael 1982 The Vail Site: A Palaeo-Indian Encampment in Maine. Bulletin of the Buffalo Society of Natural Sciences, Vol. 30. Buffalo. Griffin, John W. 1974 Investigations in Russell Cave. National Park Service. Grimes, John R. 1979 A New Look at Bull Brook. Anthropology 3:109–130. Hajic, E. R., R. D. Mandel, and E. A. Bettis, III 2000 Stratigraphic and Paleoenvironmental Investigations. In The 1999 Excavations at the Big Eddy Site (23CE426), edited by N. H. Lopinot, J. H. Ray, and M. D. Conner, pp. 26–36. Southwest Missouri State University Center for Archaeological Research Special Publication 3. Springfield. Hajic, E. R., R. D. Mandel, J. H. Ray, and N. H. Lopinot 1998 Geomorphology and Geoarchaeology. In The 1997 Excavations at the Big Eddy Site (23CE426) in Southwest Missouri, edited by N. H. Lopinot, J. H. Ray, and M. D. Conner, pp. 74–108. Special Publication No. 2. Center for Archaeological Research, Southwest Missouri State University, Springfield. 2005

Fogelman, G. L., and S. W. Lantz 2006 The Pennsylvania Fluted Point Survey. Fogelman Publishing Company, Turbotville, Pennsylvania. Frison, G. C. 1984 The Carter/Kerr-McGee Paleoindian Site: Cultural Resource Management and Archaeological Research. American Antiquity 49(2):​ 218–314. Funk, R. E., and B. Wellman 1984 Evidence of Early Holocene Occupations in the Upper Susquehanna Valley, New York State. Archaeology of Eastern North America 12:81–109. Futato, E. 1977 Radiocarbon Dates from the Tennessee Valley of Northern Alabama. Tennessee Archaeological Society Newsletter 22:35–42. Gardner, W. M. 1974 The Flint Run Complex: Pattern and Process During the Paleo-Indian to Early Archaic. In The Flint Run Paleoindian Complex: A Preliminary Report, 1971–1973 Seasons, edited by William M. Gardner, pp. 5–47. Occasional Publication No. 1. Catholic University of America Archaeology Laboratory, Washington, D.C. 1983 Stop Me If You’ve Heard This One Before: The Flint Run Complex Revisited. Archaeology of Eastern North America 11:49–64. Gingerich, Joseph A. M. 2007 Shawnee-Minisink Revisited: Re-Evaluating the Paleoindian Occupation. Unpublished Master’s thesis, Department of Anthropology, University of Wyoming. 2011 Down to Seeds and Stones: A New Look at the Subsistence Remains from ­Shawnee-​ Minisink. American Antiquity 76(1):127–144. Gingerich, Joseph A. M., and Michael R. Waters 2007 New Dates from the Shawnee-Minisink Site, Pennsylvania. Current Research in the Pleistocene 24:90–92. Goodyear, Albert C. 1974 The Brand Site: A Techno-Functional Study of a Dalton Site in Northeast Arkansas. Research Series 7. Arkansas Archaeological Society, Fayetteville. 1982 The Chronological Position of the Dalton Horizon in the Southeastern United States. American Antiquity 47:382–395. 1999 The Early Holocene Occupation of the Southeastern United States: A Geoarchaeological Summary. In Ice Age Peoples of North America, edited by Robson Bonnichsen and Karen L. Turnmire, pp. 432–481. Oregon State University Press, Corvallis. 33

Miller and Gingerich Stone Age Spear and Arrow Points of the Midcontinental and Eastern United States: A Modern Survey and Reference. Indiana University Press, Bloomington. Kitchel, N. R., and R. Boisvert 2011 New Radiocarbon Date from the Colebrook Paleoindian Site. New Hampshire Archaeological Society Newsletter 27(1):3. Koldehoff, Brad, and Thomas J. Loebel 2009 Clovis and Dalton: Unbounded and Bounded Systems in the Midcontinent of North America. In Lithic Materials and ­Paleolithic ­Societies, edited by Brooke Blades and Brian Adams, pp. 270–288. Wiley-Blackwell, ­Hoboken. Larson, Mary Lou, Marcel Korfeld, and George C. Frison 2009 Hell Gap: A Stratified Paleoindian Campsite at the Edge of the Rockies. University of Utah Press, Salt Lake City. Laub, Richard S. 2002 The Paleoindian Presence in the Northeast: A View from the Hiscock Site. In Ice Age Peoples of Pennsylvania, edited by K. W. Carr and J. M. Adovasio, pp. 105–122. Recent Research in Pennsylvania Archaeology, No. 2. Pennsylvania Historical and Museum Commission, Harrisburg. Laub, Richard S., and Arthur Spiess 2003 What Were Paleoindians Doing at the Hiscock Site? In The Hiscock Site: Late Pleistocene and Holocene Paleoecology and Archaeology of Western New York State, edited by R. S. Laub, pp. 261–271. Bulletin of the Buffalo Society of Natural Sciences 37. Buffalo. Leigh, David S. 2004 Late Pleistocene Braided River of the Atlantic Coastal Plain, USA. Quaternary Science Reviews 23:65–84. 2006 Terminal Pleistocene Braided to Meandering Transition in Rivers of the Southeastern USA. Catena 66:155–160. Lepper, B. T., and L. S. Cummings 1993 Early Holocene Archaeology and Paleoecology of the Manning Site (33CT476) in Southern Ohio. Current Research in the Pleistocene 10:32–33. Levine, M. A. 1990 Accommodating Age: Radiocarbon Results and Fluted Point Sites in Northeastern North America. Archaeology of Eastern North America 18:33–63. Lewis, T. M. N. 1954 The Cumberland Point. Bulletin of the Oklahoma Anthropological Society 11:7–8. 1995

Haynes, C. Vance 1983 Fluted Points in the East and West. Archaeology of Eastern North America 11:24–27. Haynes, C. V., D. J. Donahue, A. J. T. Hull, and T. H. Zabel 1984 Application of Accelerator Dating to Fluted Point Paleoindian Sites. Archaeology of Eastern North America 12:184–189. Haynes, G. 2002 The Early Settlement of North America: The Clovis Era. Cambridge University Press, Cambridge. Hemmings, C. Andrew 2004 The Organic Clovis: A Single Continent-Wide Cultural Adaptation. Ph.D. dissertation, Department of Anthropology, University of Florida. University Microfilms, Ann Arbor. Herbstritt, James T. 1988 A Reference for Pennsylvania Radiocarbon Dates. Pennsylvania Archaeologist 58(2):1–29. Hoffman, Curtiss 1987 A Preliminary Report on the Excavation of Cedar Swamp 4. Archaeological Quarterly of the W. Elmer Ekblaw Chapter, Massachusetts Archaeological Society 9(1):1–25. 1988 Radiocarbon Dates from Massachusetts: An Annotated Listing. Bulletin of the Massachusetts Archaeological Society 39(1):21–39. Holliday, Vance T. 2000 The Evolution of Paleoindian Geochronology and Typology on the Great Plains. Geoarchaeology 15:227–290. Howard, C. D. 1988 Fluting Technology at the Lincoln Hills Site. Plains Anthropologist 33:395–398. 1990 The Clovis Point: Characteristics and Type Description. Plains Anthropologist 35:255–262. Inashima, P. Y. 2008 Establishing a Radiocarbon Date Based Framework for Northeastern Virginia Archeology. Quarterly Bulletin of the Archeological Society of Virginia 63. Richmond. Johnson, Eileen, and Vance T. Holliday 2004 Archeology and Late Quaternary Environments of the Southern High Plains. In Prehistory of Texas, edited by T. K. Perttula, pp. 283–295. Texas A&M University Press, College Station. Jones, B. Calvin, and Louis B. Tesar 2000 The Wakulla Springs Lodge Site (8WA329). Florida Anthropologist 53(2/3):98–114. Justice, Noel 1987 Stone Age Spear Points and Arrow Points of the Midcontinental and Eastern United States. Indiana University Press, Bloomington. 34

Paleoindian Chronology and the Eastern Fluted Point Tradition C. W. McNett, Jr., pp. 3–14. Academic Press, Orlando. McWeeney, Lucinda J. 1994 Archaeological Settlement Patterns and Vegetation Dynamics in Southern New England in the Late Quaternary. University Microfilms, Ann Arbor. Meltzer, D. J. 1988 Late Pleistocene Human Adaptations in Eastern North America. Journal of World Archaeology 2:1–52. 2004 Modeling the Initial Colonization of the Americas: Issues of Scale, Demography, and Landscape Learning. In The Settlement of the American Continents: A Multidisciplinary Approach to Human Biogeography, edited by C. Barton, G. Clark, D. Yesner, and G. Pearson, pp. 123–137. University of Arizona Press, ­Tucson. 2009 First Peoples in a New World: Colonizing Ice Age America. University of California Press, Berkeley. Meltzer, D. J., and J. R. Cooper 2006 On Morphometric Differentiation of Clovis and Non-Clovis Blades. Current Research in the Pleistocene 23:143–145. Meltzer, D. J., and Vance T. Holliday 2010 Would North American Paleoindians Have Noticed Younger Dryas Age Climate Changes? Journal of World Prehistory 23:1–41. Michels, J. W., and I. F. Smith 1967 A Preliminary Report of Archaeological Investigations at the Sheep Rock Shelter Site, Huntingdon, Pennsylvania. Pennsylvania State University, State College. Miller, D. Shane 2010 Clovis Excavations at Topper 2005–2007: Examining Site Formation Processes at an ­Upland Paleoindian Site Along the Middle Savannah River. Occasional Papers 1 of the Southeastern Paleoamerican Survey. South Carolina Institute of Archaeology and Anthropology, Columbia. Moeller, Roger W. 1980 6LF21: A Paleoindian Site in Western Connecticut. American Indian Archaeological ­Institute, Washington, Connecticut. Morrow, Juliet E., and Toby A. Morrow 1999 Geographic Variation in Fluted Projectile Points: A Hemispheric Perspective. American Antiquity 64:215–230. Morse, D. F. (editor) 1997 Sloan: A Paleoindian Dalton Cemetery in Arkansas. Smithsonian Institution Press, Washington, D.C.

Lewis, Thomas M. N., and Madeline Kneberg 1959 The Archaic Culture in the Middle South. American Antiquity 25(2):161–183. Long, A., and B. Rippeteau 1974 Testing Contemporaneity and Averaging Radiocarbon Dates. American Antiquity 39(2):205–214. Lopinot, N. H., J. H. Ray, and M. D. Conner (editors) 2000 The 1999 Excavations at the Big Eddy Site (23CE426). Center for Archaeological Research, Southwest Missouri State University Special Publication No. 3. Springfield. MacDonald, G. F. 1968 Debert: A Paleoindian Site in Central Nova Scotia. Anthropology Papers No. 16. National Museum of Canada, Ottawa. Maslowski, R. F., C. M. Niquette, and D. M. Wingfield 1995 The Kentucky, Ohio and West Virginia Radio­carbon Database. West Virginia Archaeologist 47:1–75. Mason, Ronald J. 1962 The Paleo-Indian Tradition in Eastern North America. Current Anthropology 3:​227– 278. McAvoy, J. M. 1992 Nottoway River Survey, Pt. I: Clovis Settlement Patterns: The 30 Year Study of a Late Ice Age Hunting Culture on the Southern Interior Coastal Plain of Virginia. Archeological Society of Virginia Special Publication No. 28. Dietz Press, Richmond. McAvoy, J. M., and L. D. McAvoy 1997 Archaeological Investigations of Site 48SX202, Cactus Hill, Sussex County, VA. Research Report Series No. 8. Virginia Department of Historic Resources, Richmond. McCary, B. C. 1983 The Paleo-Indian in Virginia. Archeological Society of Virginia, Quarterly Bulletin 38(1):​ 43–70. McDonald, J., and M. Kay 1999 Pre-Clovis Archaeology at SV-2, Saltville, Virginia. Paper presented at the 64th Annual Meeting of the Society for American Archaeology, Chicago. McNett, Charles W., Jr. (editor) 1985 Shawnee-Minisink: A Stratified Paleoindian– Archaic Site in the Upper Delaware Valley of Pennsylvania. Academic Press, Orlando. McNett, C. W., R. J. Dent, J. Evans, S. B. Marshall, and B. A. McMillan 1985 The Upper Delaware Valley Early Man Project. In Shawnee-Minisink: A ­Stratified ­Paleoindian–​Archaic Site in the Upper Delaware Valley of Pennsylvania, edited by 35

Miller and Gingerich Robinson, Brian S., Jennifer C. Ort, William E. Eldridge, Adrian L. Burke, and Bertrand G. Pelletier 2009 Paleoindian Aggregation and Social Context at Bull Brook. American Antiquity 74(3):423– 447. Sellards, E. H. 1952 Early Man in America. University of Texas, Austin. Sellet, Frederic, James Donahue, and Matthew G. Hill 2009 The Jim Pitts Site: A Stratified Paleo­indian Site in the Black Hills of South Dakota. American Antiquity 74(4):735–758. Sherwood, Sarah C., Boyce N. Driskell, Asa R. Randall, and Scott C. Meeks 2004 Chronology and Stratigraphy at Dust Cave, Alabama. American Antiquity 69(3):533–554. Shippee, J. Mett 1966 The Archaeology of Arnold Research Cave, Callaway County, Missouri. Missouri Archaeologist 28. Shott, Michael J., and Jesse A. M. Ballenger 2007 Biface Reduction and the Measurement of Dalton Curation: A Southeastern United States Case Study. American Antiquity 72(1):153–175. Smith, Bruce D. 1986 The Archaeology of the Southeastern United States: From Dalton to De Soto, 10,500–500 Bp. Advances in World Archaeology 5:1–92. Soday, Frank J. 1954 The Quad Site: A Paleo-Indian Village in North Alabama. Tennessee Archaeologist 10:1–20. Spiess, A., and D. Wilson 1987 Michaud: A Paleoindian Site in the New England Maritimes Region. Maine Historic Preservation Commission and the Maine Archaeological Society, Augusta. Spiess, A. E., Deborah Wilson, and James Bradley 1998 Paleoindian Occupation in the New ­England–​Maritimes Region: Beyond Cultural Ecology. Archaeology of Eastern North America 26:201–264. Stanzeski, Andrew J. 1996 Two Decades of Radiocarbon Dating from the New Jersey Shore. Bulletin of the Archaeological Society of New Jersey 51:42–45. Steponaitis, Vincus P. 1986 Prehistoric Archaeology of the Southeastern U.S. Annual Review of Anthropology 15:383–393. Stewart, R. M., Jay Custer, and Donald Kline 1991 A Deeply Stratified Archaeological and Sedimentary Sequence in the Delaware River Valley of the Middle Atlantic Region, United States. Geoarchaeology 6(2):169–182.

Norton, M. R., and J. B. Broster 1992 40HS200: The Nuckolls Extension Site. Tennessee Anthropologist 17:13–32. 1993 Archaeological Investigations at the Puckett Site (40SW228): A Paleoindian/Early Archaic Occupation on the Cumberland River, Stewart County, Tennessee. Tennessee Anthropologist 18(1):45–58. O’Brien, Michael J., John Darwent, and R. Lee Lyman 2001 Cladistics Is Useful for Reconstructing Archaeological Phylogenies: Paleoindian Points from the Southeastern United States. Journal of Archaeological Science 28:1115–1136. Peck, Rodney M. 2002 Eastern Fluted Points. Rodney Peck, ­ annapolis. K Petersen, James B., Robert N. Bartone, and Belinda J. Cox 2000 The Varney Farm Site and the Late Paleo­ indian Period in Northeastern North America. Archaeology of Eastern North America 28:113–140. Prasciunas, Mary M. 2008 Clovis First? A Study of Space, Time, and Technology. Unpublished Ph.D. dissertation, Department of Anthropology, University of Wyoming. 2011 Mapping Clovis: Projectile Points, Behavior, and Bias. American Antiquity 76(1):107–126. Puryear, Kristen Ann 2009 The Late Pleistocene/Early Holocene Vegetational History of the Vail Site. In ­Palaeo-Americans and Palaeo-Environment at the Vail Sites, Maine, edited by R. Michael Gramly, pp. 9–54. Persimmon Press, North Andover. Redmond, B., and K. Tankersley 2005 Evidence of Early Paleoindian Bone Modification and Use at the Sheriden Cave Site (33WY252), Wyandot County, Ohio. American Antiquity 70(3):503–526. Ridge, J. C. 2003 The Last Deglaciation of the Northeastern United States: A Combined Varve, Paleomagnetic, and Calibrated 14C Chronology. In Geoarchaeology of Landscapes in the Glaciated Northeast, edited by L. David-Cremeens and John P. Hart, pp. 15–48. New York State Museum Bulletin, Vol. 497. New York. Robinson, B. S. 1996 Archaic Period Burial Patterning in Northeastern North America. Special issue, “Contributions to the Archaeology of Northeastern North America,” edited by Brian S. Robinson. Review of Archaeology 17(1):33–44. 36

Paleoindian Chronology and the Eastern Fluted Point Tradition Thulman, D. 2008 A Typology of Fluted Points from Florida. Florida Anthropologist 60:63–75. Wagner, D. P. 1994 Pedology and Geomorphology of the Depew Recreation Area. In Geomorphology, Remote Sensing, and Archaeological Monitoring at Depew Recreation Area: Delaware Water Gap National Recreation Area, New Jersey, edited by Paul Y. Inashima, pp. 1–19. Department of the Interior, National Park Service, Denver Service Center, Eastern Team, Eastern Applied Archaeology Center, Silver Spring, Maryland. Wagner, D. P., and J. M. McAvoy 2004 Pedoarchaeology of Cactus Hill, a Sandy Paleo­indian Site in Southeastern Virginia, USA. Geoarchaeology 19:297–322. Wall, R. D. 2008 Deep Testing and Radiocarbon Dates from the Barton Site (18AG3), Allegany County, Maryland. Maryland Archaeology 44:1–4. Waters, M. R., and Thomas W. Stafford, Jr. 2007 Redefining the Age of Clovis: Implications for the Peopling of the Americas. Science 315:1122–1126. Waters, M. R., T. W. Stafford, B. G. Redmond, and K. B. Tankersley 2009a The Age of the Paleoindian Assemblage at Sheriden Cave, Ohio. American Antiquity 74(1):107–112. 2009b Clovis and the American Mastodon at Big Bone Lick, Kentucky. American Antiquity 74(3):558–567. Webb, S. David (editor) 2006 First Floridians and Last Mastodons: The Page-Ladson Site in the Aucilla River. Springer, Dordrecht. Weninger, B., O. Jöris, and U. Danzeglocke 2007 CalPal-2007. Cologne Radiocarbon Calibration and Palaeoclimate Research Package, Cologne. Wood, Raymond W., and Bruce R. McMillan 1976 Ancient Man in North America: A Case Study in the Ozark Highland. Academic Press, New York.

Stewart, R. M., Jeremy Koch, Kurt Carr, Del Beck, Gary Stichcomb, and Frank Vento 2012 The Paleoindian Occupation at Nesquehoning Creek, Carbon County. Paper presented at the Annual Meeting of the Society for American Archaeology, Memphis. Storck, P. L., and J. D. Holland 2003 From Text to Context: Hiscock in the Paleo­ indian World. In Late Pleistocene and Holocene Paleoecology and Archaeology of Western New York State, edited by R. S. Laub, pp. 281– 300. Bulletin of the Buffalo Society of Natural Sciences 37. Buffalo. Surovell, T. A. 2000 Radiocarbon Dating of Bone Apatite by Step Heating. Geoarchaeology 15:591–608. Surovell, T. A., and P. J. Brantingham 2007 A Note on the Use of Temporal Frequency Distributions in Studies of Prehistoric Demography. Journal of Archaeological Science 34:1868–1877. Surovell, T. A., J. Finely, G. M. Smith, P. J. Brantingham, and R. L. Kelly 2009 Correcting Temporal Frequency Distributions for Taphonomic Bias. Journal of Archaeological Science 36:1715–1724. Tankersley, K. B. 1985 The Potential for Early Man Sites at Big Bone Lick, Kentucky. Tennessee Anthropologist 10(1):27–49. 1990 Paleoindian Period. In The Archaeology of Kentucky: Past Accomplishments and Future Directions, edited by David Pollack, pp. 73– 142. Kentucky Heritage Council, Frankfort. Tankersley, K. B., S. Vanderlaan, J. D. Holland, and S. Bland 1997 Geochronology of the Arc Site: A Palaeo­ indian Habitation in the Great Lakes Region. Archaeology of Eastern North America 25:31– 44. Thorbahn, Peter 1982 Interstate 495 Archaeological Data Recovery Project. Manuscript on file at the Massachusetts Historical Commission, Boston.

37

2

Paleoindian Environment and Subsistence Paradigm Case from New England to Virginia and Ohio Lucinda J. McWeeney

Reexamining the environmental conditions that greeted the early settlers in North and South America plays a critical role in how we view the past. Paleoecological studies, oxygen and carbon isotope analyses, temperature and dating extrapolations from tree rings, coral, and the Greenland ice core annular layers (Greenland Ice Core Project [GISP2]) have addressed several significant issues to revise our views on global climate change and its expression in eastern North America. In this chapter, I will briefly summarize the paleoenvironment of several archaeology sites from Maine south to Virginia and west to Ohio (Figure 2.1) and focus on how the use of localized data can create a more holistic view of past environments. Unless otherwise noted dates in this chapter are presented as calendar years before present (cal Bp).

period just before 14,000 cal Bp (McWeeney 1994; Peteet 1995; Shuman et al. 2007). Before 14,600 Bp tundra-like assemblages comprising dwarf shrubs, herbs, and sedges dominated the majority of the Northeast. The warming that occurred after this time was marked by increases in spruce (Picea sp.), larch (Larix sp.), and birch (Betula sp.) pollen throughout most of the region (Gaudreau and Webb 1985; Newby et al. 2005; Peteet 1995). Following the retraction of the ice sheets, areas of the Northeast above 43° latitude were open tundra, mixed spruce parkland, and herbs, with a relatively high percentage of sedge pollen possibly related to the areal extent of wetlands until after 12,000 Bp (Gaudreau and Webb 1985). A recent summary of pollen data for the Northeast by Newby et al. (2005) tracks landscape variations by examining selected vegetation and change on a 1,000-year scale across the region. At 14,000 cal Bp, pollen records suggest a spruce parkland over much of southern New England, with open areas of sedge (>5 percent), herbs, and mixed woodlands from southern Massachusetts northward (Davis 1958; Gaudreau and Webb 1985; McWeeney 2007a; Newby et al. 2005; Sneddon 1987). When humans first arrived sometime around 13,000 Bp, pollen records selected for statistical analysis show an average of 2.5–5 percent sedge pollen throughout central New England, with the most homogeneous and concentrated distribution of sedge beginning in central New Hamp-

Summary of the Northeast Paleoenvironment

By 24,000 cal Bp, the glacial ice sheets began to retract northward (Ridge 2003). When humans first entered the region ca. 13,500–13,000 years ago, the ice sheets had already retracted into Canada above 44.5° latitude (Lothrop et al. 2011; Ridge 2003). Based on this chronology, plant communities would have had ample time to rebound before humans colonized the Northeast. After the Last Glacial Maximum, temperatures began to show a steady increase, reaching the warmest mean July temperatures for the postglacial Pleisto­cene 38

Paleoindian Environment and Subsistence Paradigm Case

Figure 2.1. Map of sites discussed in the chapter.

shire and Vermont (Lothrop et al. 2011; Newby et al. 2005). In portions of Maine and the Canadian Maritimes, sedge pollen levels are above 5 percent (Newby et al. 2005). Pollen records for trees show spruce at levels between 20 and 40 percent over much of New England, except for some coastal areas in southern Maine where spruce pollen levels exceed 40 percent. The additional pine pollen contributes to the evidence of mixed conifer woodlands (Davis and Jacobson 1985; Newby et al. 2005:Figure 1). Around 13,000 Bp, pollen percentages of pine (Pinus spp.) and birch (Betula spp.) are between 5 and 20 percent across New England. Dispersed stands of oak are represented by pollen levels between 5 and 20 percent in Massachusetts, parts of New Hampshire, and northeastern New York (Davis 1958; Gaudreau 1988; Newby et al. 2005). These data may be consistent with a taiga or boreal forest biome, which is usually dominated by coniferous or needleleaf tree species but can have deciduous species such as oak along its southern boundaries (Hare 1994). The spruce woodlands and open tundra-like vegetation that character-

ized the mixed open-forest landscape of northern New England between 14,000 and 13,000 cal Bp differed from that in Massachusetts, Rhode Island, and Connecticut, with major change during the Younger Dryas and the onset of the Holocene (Davis 1958, 1969; Peteet et  al. 1993 [Lindsley Pond]). Other statistical reconstructions of late glacial needleleaf tree cover in New England and the Canadian Maritimes region also suggest 20– 40 percent coverage below 44° latitude between 14,000 and 11,000 Bp, which suggests that a relatively open environment existed into the Holocene (Williams 2003). In parts of southern New England (below 42° latitude), tree cover and forest diversity increase. At 14,000 Bp, parts of Massachusetts show lower pollen percentages of sedge and higher concentrations of spruce and pine ranging between 20 and 40 percent (Leduc 2003; Newby et al. 2005). Dispersed signatures of oak also show up in the same areas, with pollen percentages between 5 and 20 percent (Gaudreau 1986, 1988; Leduc 2003; Newby et al. 2005). In fact, the northern limit of oak may extend farther at 14,000 than at 13,000 39

McWeeney

cal Bp (Leduc 2003; McWeeney 1994; Newby et al. 2005). Pollen cores in southeastern New York (Maenza-Gmelch 1997) and eastern Massachusetts (Sneddon 1987) suggest a more temperate deciduous-boreal-coniferous forest with Quercus, Fraxinus, Ostrya, Pinus, and Abies between 14,700 and 13,040 Bp.1 These data suggest that some parts of New England may have exhibited greater tree cover. For example, there are significant amounts of pine (20–40 percent) and oak (5–20 percent) in pollen cores in eastern and southern New York by 13,000 Bp (Maenza-Gmelch 1997; Newby et al. 2005; Peteet 1995). My own botanical interpretations based on seasonal temperatures and growth ranges for southern New England demonstrate that oak, ash, red maple, and hornbeam would have existed farther north in Connecticut and Massachusetts prior to the Younger Dryas cold event (­McWeeney 1994, 2007a). Percentages of sedge pollen and grasses both in this area and in previously glaciated Pennsylvania were highest before 14,000 cal Bp and appear to have quickly declined over time (Gaudreau and Webb 1985). Evidence of pine pollen at levels above 60 percent in areas of New Jersey and Pennsylvania also suggests that a more temperate forest environment existed well before the onset of the Younger Dryas (Dent 1979; Watts 1979; Yu 2007). The Younger Dryas (YD), 12,900–11,600 cal Bp, marks a rapid climatic reversal when within decades temperatures rapidly became cooler than in the previous millennium (McWeeney 1994; Peteet 1995; Peteet et al. 1990; Shuman et al. 2002). The impact of this climate change is important, as human populations just arriving in the Northeast may have found vegetation and temperatures rapidly changing and responding to climate oscillation (Lothrop et  al. 2011; McWeeney 1994; Newby et al. 2005; Peteet et al. 1990; Shuman et al. 2002; Shuman et al. 2004; Yu 2007). While vegetation changes in selected genera do not appear to have occurred until the end of the YD, we should not view the climate as static during this time period when looking at broader floral changes. For example, the YD may have caused an expansion of more open t­ undra-​ like grasslands vegetation in northern New England, extending the southern ranges of migrating caribou into parts of northern Massachusetts 40

(Newby et al. 2005). However, while grasslands may have extended in portions of northern New England, there is also great heterogeneity in vegetation (see the Willimantic K ­ ettle section in McWeeney 1994). For example, Mayle and Cwynar (1995:815) consider parts of New Brunswick to be a closed spruce forest with spruce pollen percentages as high as 60 percent just prior to the YD. Evidence of a decrease in open landscape or variation around 13,000 Bp is evident in pollen cores from Splan Pond, New Brunswick, and Moulton Pond, Maine, where the percentage of sedge pollen decreases by over 100 percent within the 90 mi that separates these two localities (Newby et al. 2005:​ Figure 3). While the percentage of sedge at Moulton Pond may continue to support the interpretation of an open landscape, the marked reduction in sedge and increase in spruce around 13,000 Bp suggest a rapid reduction in tundra-like conditions long before this farther south. Evidence of lake level fluctuation(s) from throughout the region also suggests potential variation in vegetation and climate. Lake-level data suggest multicentury droughts throughout most of New England. Ground-penetrating ­radar studies at New Long Pond and Rocky Pond in southeastern Massachusetts demonstrate a ­several-​hundred-year period of lower lake levels beginning at the Younger Dryas (Newby et al. 2009). Similar evidence from Davis Pond in western Massachusetts also supports this interpretation, with changes in subsurface facies of pond sediments in layers bracketed between 13,300 and 10,700 cal Bp suggesting dry conditions during the Younger Dryas and late glacial period (Newby et al. 2011). Lake level data from the northern Middle Atlantic region at White Lake, New Jersey, however, show that both lake levels and temperatures fluctuated throughout the YD interval (Yu 2007). Data from cores at White Lake suggest that while vegetation may have quickly responded to climate change in parts of the region, it may have lagged in others. Yu’s work demonstrates that the largest vegetation shift from Picea (spruce) and Alnus (alder) to Quercus (oak) and other hardwoods occurred around the Holo­cene onset. These data reiterate Shuman et al.’s (2002) study, which shows that both climate and moisture appear to have fluctuated throughout the

Paleoindian Environment and Subsistence Paradigm Case

YD, with major changes in vegetation reoccurring in the Holocene as temperatures became warmer and drier (Shuman et al. 2004; Yu 2007), similar to the environmental changes during the Bølling/Allerød warming more than 1,500 years before the YD. To summarize, portions of the eastern U.S. landscape appear to have varied greatly when humans arrived in the region. Around 13,500– 13,000 cal Bp, the uppermost Northeast (>43.5°N latitude) appears to have been relatively open, with significant stands of spruce and other needleleaf species. This environment could have supported large herds of caribou in current-day Maine and New Brunswick and within the vicinity of the Debert site, Nova Scotia. However, while the area from New Hampshire and Vermont to the north is considered to have been relatively open, there is sufficient evidence of increased tree populations. For example, Newby et al. (2005) show concentrations of spruce pollen above 40 percent in some locations and the extension of pine into southern Maine as early as 14,000 cal Bp (Davis and Jacobson 1985; Kellogg 1991). Below 42.5° latitude researchers suggest a moderately open forest dominated by coniferous species in parts of the region (Newby et al. 2005; Williams 2003). Cold-temperate species such as oak appear to have been well established in portions of New York and Massachusetts before 13,000 Bp. In the south, below Massachusetts, the environment appears to be variable, with a greater diversity of plant species represented in pollen cores, including evidence of rather unique and more-temperate-than-expected species in patches across the landscape (Jackson and Williams 2004; McWeeney 1994). Despite the differences in vegetation that may have existed, some researchers suggest that much of the Northeast was under less than 50 percent tree cover until after 11,000 cal Bp (Williams 2003). Whether this hypothesized open environment across a large geographical area is related to low temperatures, low precipitation, low CO2 concentrations (Williams 2003), or just a general moisture balance relationship that prevented the spread and expansion of some taxa (Shuman et al. 2004), the open environments that are posited for this period likely promoted a rich mosaic environment that corresponded to increased opportunities for

humans to exploit a number of different plants and animals. The impact of the Younger Dryas on vegetation is unclear; however, as Yu suggests, the “YD was not uniformly cold” (2007:301). Therefore, we should not expect that human responses were uniform during this time but, rather, were variable as hunter-gatherers adapted to fluctuations in both climate and vegetation. At the least, oscillation in climate likely affected the predictability, availability, and location of seasonal s­ ubsistence items (Kellogg and Custer 1995). For these ­reasons, understanding human responses to environmental changes requires a detailed reconstruction of the locations people occupied. Using Archaeological Data to Reconstruct Local Environments

Improvements in technology now allow for the calibration of radiocarbon dates into calendar years with much tighter control over the s­ tandard deviation and range of dates (Fiedel 1999, 2004; Stafford 1999; Stuiver and Reimer 1993; W ­ aters and Stafford 2007). Refinements of dating methods using accelerator mass spectrometry (AMS [Gove 1992; Stafford 1999]) and optically stimulated luminescence (OSL) have enhanced our ability to distinguish cultural stratification in deeply buried sites with poor organic preservation (Goodyear 2000, 2005; McAvoy et al. 2000). Once it became possible to date milligram-size organic samples, then more of the recovered plant remains could be identified to establish what was growing during the time of interest. Very few sites left a substantial imprint from the first humans entering the eastern North American landscape. Equally, in many instances the botanical record has been difficult to recover or to assess due to poor preservation or earlier recovery techniques. Within the past few decades, advanced paleobotanical techniques have dramatically expanded the recovery and identifica­ tion of botanical assemblages from archaeological sites (McWeeney 2007a; Pearsall 2000; Piperno 1988). Archaeologists employ several techniques to establish vegetation changes. Plant macro­fossils from wood, seeds, stems, and leaves (Hollen­bach 2007; McWeeney 1994, 2007a) combined with microbotanical identification of starch grains, diatoms (McWeeney 1994, 2007a), ­phyto­liths 41

McWeeney

(Madella and Zurro 2007; Piperno 1988), and pollen (Kellogg 1991; Kellogg and Custer 1995; ­McWeeney 2007a; Pearsall 2000) can then be correlated with archaeological components, stratigraphy, and chronologies (McAvoy et al. 2000; McWeeney 2001). By addressing the patterns of human settlement changes and the associated plant remains from archaeological sites from Maine to Virginia we can revise and enhance our image of the environments that confronted Paleoindians entering the Northeast. With the identification of plants available in a specific time period, based on mul­ tiple lines of evidence (macrofossils, pollen, phytoliths, starch grains, and diatoms), we can extrapolate what was available for food, medicine, fuel, travel, and ritual: this is the subsistence paradigm case. I believe that identified plant remains from well-dated proveniences may and should be used to describe and enhance cultural subsistence included within the Paleoindians’ seasonal rounds. This also becomes the subsistence paradigm case. Do we have to prove that Paleoindians consumed the myriad of plants available to them by finding charred seeds and charcoal in hearths? Or can we relate to the diversity of plants available where humans chose to set up camp and trust that they chose to live there based on the variety of food resources? It is no longer necessary to describe the first humans entering North and South America as struggling across tundra environments from Siberia, chasing mammoths and mastodons (Mc­ Weeney 1994, 1996; McWeeney and Kellogg 2001; Meltzer 2009; Walker 2007). While increas­ing evidence of large game associated with Paleoindian tools exists (e.g., Broster et al., this volume; Driskell and Walker 2007; Fisher 2004; Overstreet 2005; Spiess and Wilson 1987; Spiess et al. 1998; Yesner 2007), in many instances bones are found from smaller game animals at Paleo sites (Dent 2007; Dent and Kauffman 1985; Fiedel 2007; Storck and Spiess 1994; Walker 2007; Yesner 2007). For a long time, a dearth of preserved plant remains from Paleoindian sites (more a function of recovery techniques) supported the hypothesis of a specialized big-game-hunting lifestyle. Admittedly, without evidence of food-­processing tools, the processing of plants does not play a

large role in our interpretations and reconstructions. However, we must acknowledge that Paleoindians did not enter the “New World” without a great store of experience on how and what to collect and eat. Ancestral knowledge followed people through evolutionary paths from bipedalism to Homo sapiens sapiens (Grushkin 2010). As noted by Gillam (cited in Anderson 2005:36), social networks, lithic sourcing, and topography within and between drainages shaped the Paleoindian settlement patterns and their search for food in the New World. In this chapter, I describe the late Pleistocene/ early Holocene paleoenvironmental record for southern New England, the Middle Atlantic region, and Ohio based upon plant macrofossils, pollen, and phytoliths (McWeeney 2007a). Those results will then be extrapolated to provide a broader picture of the plants available for Paleoindian subsistence. Sites and Associated Wetlands Neal Garrison Site, Maine

In the late 1990s, an environmental assessment required for a pipeline survey discovered the Neal Garrison site (Kellogg and Simons 2000), located in southwestern Maine in the modern ­conifer-​ hardwood forest region (Braun 2000). Artifacts correlated with a Paleoindian occupation, many of which came from Feature 2, included a channel flake, a graver, and sidescrapers made from Munsungun cherts (Spiess et al. 1998). Excavating the feature in discreet levels for flotation purposes made it possible to recover charcoal and seeds. An off-site control column (a unit devoid of artifacts) was excavated for comparison purposes. Charcoal analyses produced oak and conifer charcoal along with abundant uncharred modern seeds and fungi. An AMS date ca. 10,197 cal Bp (9000 14C  bp) is, at a minimum, nearly 3,000 years too young to be appropriate for a fluted point Paleoindian site. The ligninized conifer charcoal may include spruce and/or pine; however, the structural deterioration made it impossible to identify the material further than the pine family (Pinaceae), though it rules out larch. Based on pollen data, spruce trees reached Maine between 15,570 and 13,870 cal Bp (13,000 and 12,000 14C  bp [Davis and Jacobson 1985]) and had spread into northern New Hampshire by 42

Paleoindian Environment and Subsistence Paradigm Case

12,900 cal Bp (11,000 14C  bp [Spear et al. 1994]) and northeast to Debert by ca. 12,600 cal Bp (10,600 14C  bp [Stea and Mott 1989]). The possibility of pine is a different matter. Red, jack, and white pine grow in Maine today. But red and jack pine were probably the first pines to grow in northern New England following deglaciation. Based on the pollen, white pine probably arrived in coastal Maine around 12,000 calendar years ago (Davis and Jacobson 1985). According to Gaudreau and Webb (1985), the 5 percent isopol indicates that the local presence of oak trees, probably northern red oak, arrived in coastal Maine as early as 13,316 cal Bp (11,400 14C  bp [Davis and Jacobson 1985]), just before the Killarney cold oscillation (Levesque et al. 1993) and possibly concurrent with the entrance of the first humans. The area of impact for the pipeline corridor restricted the land that could be surveyed. If future impact were to threaten the location, then additional artifacts and charred botanical remains could become available for dating and verification of a late Pleistocene occupation.

produced dates around 10,197 cal Bp (9000 14C  bp [Grimes et al. 1984; Robinson et al. 2009]), similar to the radiometric dates from the Neal Garrison site in Maine (McWeeney 2007a) and the Nevers site in New Hampshire (Boisvert 2004). More recently dated white pine charcoal found in the archived site collections at the S­ alem Peabody Museum also returned dates ranging between 11,875 and 10,609 cal Bp (9000 14C  bp [Robinson et al. 2009]). The first radiocarbon assays that are Paleoindian in age come from calcined bone, which yielded dates of 10,410 ± ​60 14C  bp and 10,380 ± ​60 14C  bp (12,350 and 12,310 cal Bp [Robinson and Ort, this volume; Robinson et al. 2009]). Despite these dated samples, the Bull Brook assemblage contains few charred remains to identify to obtain a paleoenvironmental record. Sediment coring was determined to be the best approach for recovering late Pleistocene plant remains close to the Bull Brook sites (McWeeney 1994, 2007a). Nearby Willowdale Swamp, located in the modern hemlock–white pine/hardwood forest region (Braun 2000), provided the best location for the preservation of late Pleistocene plant remains. Preservation of the macrofossil record began approximately 13,000 cal Bp (11,830 ± ​80 14C  bp [McWeeney 1994]). At that time, the coast was several kilometers to the east. After deglaciation of the landscape, incursion by the sea, and subsequent rebound of the land that resulted in the lowering of sea level, the Bull Brook basin was most likely populated by waterfowl bringing in food and discharging remnants of previous meals into the recently deglaciated area, initiating revegetation. The modern proximity of Bull Brook to Plum Island (a barrier island that developed during the Holocene) and its nearness to the mouth of the Merrimack River led to the hypothesis that wildfowl migrations along the Atlantic flyway may have played a role in human settlement selection and repeated visits to Bull Brook. Identification of aquatic plants along with ferns and sedges that grew around and in the head­ waters of Bull Brook around 13,000 cal Bp (11,000 14C  bp) reinforces this scenario. The recovery of spruce and larch needles documents the coniferous presence at nearly 14,000 cal Bp (12,000 14C  bp). Grasses and sedges ­project an image of meadows, wetlands, and open

Bull Brook, Massachusetts

The Bull Brook Paleoindian complex in northeastern Massachusetts lies less than 50 mi southwest of the Neal Garrison site, 175 mi from Jefferson, New Hampshire, and it is only 100 mi west to the Whipple site in Swanzey, New Hampshire. Discovered during a sand mining operation, the Bull Brook Paleoindian sites remain the largest Early Paleoindian period occupation in New England (Grimes et al. 1984; Newby et al. 2005; Robinson and Ort, this volume; Robinson et al. 2009). The latest publication (Robinson et al. 2009) suggests that the artifact and site distributions indicate a place used for large social gatherings. Curran (1994) describes the Bull Brook I site as a central base camp from which groups dispersed for collecting food and other resources. Dincauze (1993) calls it a marshaling location from which splinter groups met and then scouted for new resources. It may have been all of the above, but the context has been destroyed, and interpretations rely on old field notes and the memories of excavators (Robinson and Ort, this volume; Robinson et al. 2009). Absolute dating of the charcoal recovered from the salvage excavations during the ­mining process 43

McWeeney

­ oodlands where the early settlers chose to set up w camp. Herbaceous plants such as cattail, mint, and horehound that do not grow in the modern boreal forest grew at Bull Brook close to 14,000 cal Bp (11,830 14C  bp), indicating a warmer climatic regime than the tundra or open spruce forest originally interpreted from previous data. Burned spruce needles from the Willowdale core suggest that a natural fire may have opened the site up to colonization by new species. If the fire was widespread, then it may have created massive edge effects, encouraging the grazing of animals and attracting humans to the site. Pollen cores from Black Pond, southwest of Bull Brook (Sneddon and Kaplan 1987), and from Bull Brook (McWeeney 1994; Paige Newby, personal communication 2005) indicate the presence of oak, hornbeam, and small amounts of ash, elm, and hazelnut, along with representatives of the walnut family (walnuts) mixed with the spruce, fir, and larch. The conifer presence is reinforced by the presence of spruce needles close to the Bull Brook sites. A drop in water level during the Younger Dryas at Bull Brook, similar to that at Pequot between 13,172 and 11,425 cal Bp (11,200 and 10,000 14C  bp), is inferred based on a decline in the preservation of pollen and macrofossils (McWeeney 1994). Cold and deteriorating environmental conditions appear to have been widespread at that time in southern New England. Israel River Paleoindian Complex

In the late 1990s plant remains were recovered by flotation from one of the Israel River Paleoindian complex of sites near Jefferson, New Hampshire (Boisvert 1999, 2004; Boisvert and Puseman 2001). Surveys and site excavations ­indicate that the Paleoindian components in this area have not been compromised by younger occupations. In addition, the cultural groups inhabit­ ing the northern terraces, above the east-to-west-­ trending Israel River, collected local or Mt. Jasper flow-banded rhyolites as well as Munsungun chert, which may have come from 180 km northeast or from Vermont (Boisvert 2004:51). A complete fluted point was recovered from the Jefferson IV site, and blood analysis identified c­ ervid blood. In 1998, Bosivert and I began a program of column sampling at each locus, removing 25-×44

25-cm units in 2.5-cm increments for flotation. Spruce (Picea sp.) and Canadian hemlock (Taxus canadensis) charcoal have been identified from two of the loci. However, the AMS dates set the spruce presence around 8892 cal Bp (8000 14C  bp), suggesting a fire of natural rather than cultural origin. A significant contributor to the archaeobotanical assemblage, a charred water lily seed (Nymphaea odorata) found in a feature associated with the Paleoindian component from one locus strongly suggests that the inhabitants were collecting this wetland plant for processing back at the campsite, well away from the pond or stream. Ethnographic studies indicate that roasted water lily seeds could be consumed and the floral disk and seeds could be ground to make flour for roasting patties in the smoldering coals (­McWeeney 2001). The water lily plant may be used medicinally, and small amounts of the tubers may be digested without suffering toxic repercussions. The archaeological botanical assemblage was enhanced by the recovery and analysis of a series of sediment cores from Cherry Pond, Martin Meadow Pond, and York Pond to gain a greater understanding of postglacial environment. High-resolution sampling techniques at 1-cm increments have yielded plant macro­fossils, pollen, and phytolith evidence from these cores. Evidence of insect exoskeletons proliferated in the varved sediments that were deposited as the glacier remained close to the ice and still dammed Lake Israel. Soon after the glacier pulled back the lake drained, leaving Cherry Pond as the only extant postglacial open-water basin, while the Israel River channel flowed east to west in the ­valley. Small fragments of spruce and erica­ceous berry plants (Vaccinium sp.) appeared on the land shortly after the glacier withdrew. Water milfoil (Myriophyllum sp.) dominated the aquatic plant remains in the basal organic level and was soon joined by pondweed (Potamogeton sp.). The insects and both aquatic plants are well known as food for waterfowl. The pattern of insects initially attracting the birds to the pond and the water­ fowl initiating the revegetation by depositing seeds from their gullets or from trapped seeds dropping off their feathers and feet is significant. The aquatic weeds and birds would have attracted other animals to the site, followed by human hunters. The plants provide irrefutable evidence

Paleoindian Environment and Subsistence Paradigm Case

for migrating birds arriving at Cherry Pond before 13,000 cal Bp (11,000 14C  bp). The postglacial migration of plants and birds opens a new line of research for the peopling of the Northeast (Dincauze 2000). The area known as Plum Island, a now famous location where migrating fowl stop and breed, is very close to Bull Brook. While Plum Island is a more recent formation, the flyway at the mouth of the ­Merrimack River may be a link between settlement at Bull Brook and hunting/gathering birds in prehistory. The Merrimac Valley flyway would have presented a northward route between the mountains and the shore. The proximity of Bull Brook to the coast in 13,000 cal Bp (11,000 14C  bp) and to the mouth of the Merrimac River creates incentive for further evaluation and archaeological surveys for Paleoindian sites as one travels north along the river. Shared lithic sources plus hunting and gathering patterns indicate potential economic and social relations among the Israel River Paleoindian Complex, the Colebrook site along the Connecticut River (Kitchel 2008), Vail, the Bull Brook sites, interior sites in New Hampshire (Whipple and Tenant Swamp), and other coastal sites in Massachusetts and Maine. The once promising charred seed remains from the Colebrook site are being reevaluated.

the remains of a palisaded fort and postcontact log ­cabins and farms. The charred plant remains complement and enhance the environmental picture described by the anaerobically preserved vegetation. The Late Paleoindian Hidden Creek site is located in a hollow just below the ridge, above the southeast side of the basin (Jones 2000). Analysis of the charred remains from the survey excavations clearly demonstrates the prehistoric cultural use of wetland plants (Perry and Jones 2002). The plant macrofossil investigations at ­Pequot Cedar Swamp followed earlier geological and palynological research (Thorson and Webb 1991; Webb 1990). The initial pollen analysis encountered several difficulties due to a lack of recognition of some taxa as well as large sampling intervals. While bracketing and dating between centimeters to locate the Younger Dryas sediment level, I discovered that only one core produced a date in the YD level (McWeeney 1994, 1998). That date came from a bulk sediment sample from Core C, 254–264 cm below the surface (McWeeney 1994, 1998; Thorson and Webb 1991), that literally contained the thousand years representing the Paleoindian period and the Younger Dryas level and provided an averaged date of approximately 12,500 cal Bp (10,420 ± ​100 14C  bp). By identifying the plant macrofossils it was possible to expand our knowledge of what was growing locally in and around the basin. Because macrofossils do not travel far from their source, they provide a significant documentation of plants that may not be represented by the pollen. The results from the Pequot Cedar Swamp cores (McWeeney 1994, 1998) demonstrate that this part of Connecticut was vegetated with tundra or arctic/alpine-type plants before 18,000 cal Bp (15,210 ± ​80 14C  bp), establishing a minimum time for deglaciation in southeastern Connecticut. Similar tundra plant assemblages from several other sites in Connecticut dating between 16,930 and 16,220 cal Bp (14,100 and 13,540 14C  bp) have been identified (McWeeney 1991; Miller 1994; Stone and Ashley 1992). Following the Bølling/Allerød warming period, vegetation growing in and around the basin began to be preserved under anaerobic and alkaline conditions, as determined by identifying diatoms from the sediment core (McWeeney

Pequot Cedar Swamp, Connecticut

In the mid-1980s, the Mashantucket Pequot Tribal Nation and Dr. Kevin McBride initiated an extensive environmental reconstruction project in the Great Swamp as part of their archaeological survey of the oldest Native American reservation in the United States. Several sediment cores were recovered from the north, south, and east sides of the basin, known as the Pequot ­Cedar Swamp, which enabled us to interpret the last 18,000 (15,000 14c) years of vegetation changes (McWeeney 1994, 1998). The environmental reconstruction project provided significant data that were incorporated into one of the Mashantucket Pequot’s museum displays. Archaeological surveys around the swamp produced evidence of numerous sites beginning with a Late Paleoindian site (Jones 2000) and spanning Early Archaic pithouses (Forrest 1999, 2000) and numerous Middle Archaic occupations, as well as Late Archaic through contact-period sites with 45

McWeeney

1994). Pondweeds, algae, and cattail seeds were preserved. Cattail is well recognized as providing multiple resources for humans besides the food value obtained from the tubers harvested in the late summer. The stalks and leaves were woven into mats for ground cover and wall insulation for wigwams, and the seed heads were used for insulation in clothing. Ericaceous berry–­producing plants such as bilberry grew near the basin, as did sedges and ferns, presenting more available food and medicinal resources. Water lily seeds were AMS dated to 11,425 and 10,190 cal Bp (10,050 and 8890 14C  bp), indicating that the water level rose and formed an open-water pond fringed by a marsh with sedges (Cyperaceae) and ­cattails (Typha sp.) at the end of the Younger Dryas cold period. The pollen and nutshell from Pequot Swamp indicate the local presence of hazel­nut (Corylus sp.) shrubs during the thirteenth millennium cal Bp. Jones (2000) reports an AMS date of 11,886 cal Bp (10,260 ± ​70 14C  bp) on the hazelnut shell from the Hidden Creek Late Paleoindian site. In addition, charred hazelnut shells proliferated at the Sandy Hill site (Forrest 1999, 2000), located several meters from the east side of the swamp, suggesting that hazelnut played an important role in the local prehistoric diet. The charred remains of aquatic plants such as cattail, bulrush (Scirpus sp.), and arrowhead (Saggitaria sp.) found at both Hidden Creek and Sandy Hill sites between 11,886 and approximately 9500 cal Bp (10,200 and 8500 14C  bp) provide clear proof for the prehistoric exploitation of wetland plants (Perry 2012; Perry and Jones 2002; Perry and McBride 2012). The charred remains of oak and pine from fuel wood indicate that the early inhabitants selected from the available terrestrial vegetation. Based on the macrofossil remains recovered at the south end of Pequot Swamp, the tundra plants rapidly disappeared, and trees sprang up on the landscape by 14,000 cal Bp (12,030 14C  bp). Most significantly, white pine needles (Pinus strobus) were preserved, along with the first spruce (Picea sp.), larch (Larix laricina), and fir (Abies balsamea [McWeeney 1994, 1997]). The presence of white pine needles indicates warming temperatures, documenting the local landscape changes during the Bølling/Allerød period. The temperature restrictions for the survival of white pine (Burns

and Honkala 1990) furthermore support the actual presence of a suite of temperate deciduous trees, whose pollen had been reported for the late glacial-period pollen diagrams across northeastern North America (Amundson and Wright 1979; Gaudreau 1988; Gaudreau and Webb 1985) and whose growth ranges overlap with or extend beyond that of white pine (Little 1977). White pine will grow on sterile, dry soils as well as on the edge of wetlands. Willimantic Kettle, Connecticut

Sediments from a postglacial kettle hole in Willimantic (R. Thorson, personal communication 2006), in southeastern Connecticut, received dates prior to the Younger Dryas, at 13,454 cal Bp (11,470 14C  bp), and close to the end of the YD interval, at 12,350 cal Bp (10,440 14C  bp). The plant macrofossils demonstrate a broader biodiversity prior to the Younger Dryas, with spruce, willow (Salix sp.), and brambles (raspberries = Rubus sp.) growing on the land. Marsh species such as cattails and sedges fringed an open-water pond supporting aquatic pondweeds (Najas sp. and Potamogeton sp.) and water lilies (Nymphaea sp.). However, based on the plant macrofossils from the Younger Dryas event, there was a loss of boreal trees and the emergence of a leatherleaf (Chamaedaphne calyculata) bog filled with grasses and sedges. This type of landscape alteration clearly impacted the potential vegetation available for human inhabitants and the prey they hunted. Moose were ready inhabitants of a wetland/bog environment; white-tailed deer would be drawn to graze along the edges where trees met marshes. It is unlikely that caribou migrated southward during the YD; according to Steadman (Funk and Steadman 1994), the caribou were last seen in southeastern New York at Dutchess Quarry Cave around 14,800 cal Bp (12,500 14C  bp), during the Bølling/Allerød warming period. However, abundant small mammals, fish, and birds would have been available for subsistence. Sheriden Cave, Ohio

The Pleistocene faunal assemblage first drew professionals to the Sheriden Cave site in the Indian Trails complex of karst caves in north-­central Ohio. Preserved faunal remains in the cave in46

Paleoindian Environment and Subsistence Paradigm Case

Cactus Hill, Virginia

cluded bones from fish, amphibians, reptiles, ­giant beaver, peccary, short-faced bear, stag moose, and caribou, just to name a few (Wisner 1998). ­Tankersley and Redman (1999) explain that the faunal material indicates potential exploitation of several different environments around the cave by the Paleoindians, who also found shelter there. Rhythmite formations deep inside the cave and below the Paleoindian strata were examined for diatoms, to help understand the sedimentary deposition. However, it was the discovery of a silicified pine needle tracheid, or phytolith, that suggested that some of the prehuman occupation sediments were washed in from outside of the cave, 30 ft below the modern surface. Based on the stomata on the needle phytolith, the pine appears to be jack pine (Pinus banksiana), with a modern-day southern boundary north of Lake Erie. Pollen reports for nearby water bodies such as Neville Marsh indicate that jack pine would have been in north-central Ohio during the late Pleistocene (Shane and Anderson 1993). Some of the charcoal that I analyzed from the cave suggests that water transport was the vehicle of deposition. With dates bracketing 13,000 cal Bp (11,000 14C  bp), the presence of charred willow or poplar, possibly aspen, oak, and hornbeam, describes the immediate vegetation near the site at the cusp of the Younger Dryas. The early presence of oak at Sheriden Cave, along with the documentation of ash wood found in a bog in northern Ohio (Tankersley 1994) dated to 13,316 cal Bp (11,400 14C  bp), further substantiates the paleoenvironmental picture representing open woodland with a mix of conifer and temperate deciduous taxa near the cave (McWeeney 2007a). The temperate climate and mosaic environment, with a diversity of flora and fauna, appealed to the early settlers immediately before the GISP2 date for the major Younger Dryas climate change. We can extrapolate from the physical evidence of nut-producing trees and the mix of animals attracted to the site (Tankersley 1994, 1999) that a diverse assemblage of food resources was available for Paleoindian exploitation. Is this another paradigm subsistence case? Or can we agree that if the temperate plants were present and productive, people would have eaten them?

The discovery and dating of white pine charcoal in a hearth feature at the Cactus Hill site in southeastern Virginia, along the Nottoway River, caused a media and archaeologists’ furor. The AMS date of ca. 18,279 cal Bp (15,070 14C  bp) on white pine charcoal directly associated with cultural artifacts gained nationwide notoriety. The significance of the site, the artifacts, and interdisciplinary scientific exploration funded by the National Geographic Society generated immense media attention for the site’s potential pre-­ Clovis human occupation level: “Humans living in Virginia 18,000 years ago!” Southern hard pine (based on the cross-field pitting and dentate ray tracheids) charcoal recovered from a feature with Paleoindian artifacts above the white pine level had been AMS dated to 12,864 cal Bp (10,920 14C  bp [McAvoy et al. 2000; McWeeney 1997]), placing the fluted point level into what was considered an appropriate date for a Clovis occupation. But what do we do with that level of artifacts and charcoal 7,000 calendar years earlier than the accepted date for fluted point makers settling the Americas? How can that interpretation be validated? White pine does not grow near the site today; in fact it can be found in the Piedmont region of Virginia. The presence of white pine provides documentation for the changing environment and vegetation migration far south of the glaciated region. The temperature range for the growth and survival of white pine indicates a climatic amelioration to –6°C in the winter. Clearly the people chose to live near the river on the Coastal Plain (we acknowledge that the coastline was farther east at the time) in an ameliorated, mosaic environment rather than at higher elevations, where pollen collections report a predominance of spruce. It is incumbent upon us to revise our image of a dark spruce or jack pine forest encompassing the south when humans migrated across the region in search of critical lithic resources and food. McAvoy (McAvoy and McAvoy 1997:182) recently received notice of a nearly 18,000 cal Bp (15,000 14C  bp) date on oak charcoal from the Williamson site, located not too far from Cactus Hill. Once again, the presence of white pine as a thermophilous indicator suggests the opportunity for 47

McWeeney

local representation of temperate deciduous taxa (red oak, ash, red maple, hornbeam) that grow under similar temperature parameters. The AMSdated oak charcoal confirms this scenario. Moreover, ash and hornbeam are also represented in minor amounts by the pollen. Based on a phytolith study done in Great Britain (Powers-Jones et al. 1989), my working hypothesis for establishing human association with the 18,000 cal Bp (15,020 14C  bp) date at Cactus Hill proposed that the phytolith quantities would increase by orders of magnitude in the culturally associated strata. Phosphate analysis was added to the equation as another indicator of human presence. If the quantity of phytoliths and phosphates varied through time parallel with the quantity of human artifacts, I believed that would validate the pre-Clovis human occupation. By establishing overall stratigraphic integrity despite knowing that bioturbation must have occurred to some degree, I could further correlate the site, date, and human settlement. Phytolith quantity was based on actual weight of recovered material. The results of the initial investigation showed that the phytoliths increased in the levels with cultural artifacts and declined in nonoccupied levels (McWeeney 2000, 2007a). The phosphate analyses (Feathers et al. 2006) for each of these levels provided additional evidence of occupation, when mapped along with the quantity of cultural material based on the weight of the lithics. The phosphates increased in strata associated with stone tools, further supporting human presence. The correlation of the phytoliths, phosphates, and lithics produced demonstrable stratigraphic integrity, which adds further validity toward establishing human presence prior to the fluted point level around 12,864 cal Bp (10,920 14C  bp). Additional research involved processing sediments from an off-site column and a second on-site column. The phytoliths and phosphates from the off-site column and the on-site column showed more variance. While the phytoliths, phosphates, and artifacts appeared to closely parallel each other, a statistical model and wiggle matching may improve the results for the phytoliths. The off-site column phytolith quantities appeared to closely match those from the onsite column. The next step will be to identify the

phytoliths from all three columns to determine whether or not different vegetation was represented on-site due to human transport of certain plants to habitation sites. The addition of OSL dating also confirms stratigraphic integrity, and more samples will be dated in correlation with the on-site column. Contrary to analyses performed from another section of the Paleoindian site, soil micromorphology and macro soil analyses (J. McAvoy, personal communication 2006) drawn from the pre-Clovis area clearly indicate human-modified sediments from this portion of the Cactus Hill site, with charcoal and lithic debitage present in the core as well as a buried A-horizon below the pre-Clovis level. The scientific studies performed for Cactus Hill fully document the pre-Clovis occupation at 18,000 cal Bp (see Fiedel, this volume). Analysis of the phytoliths may provide environmental information critical to revising the paradigm subsistence case. Shawnee-Minisink, Pennsylvania

The recent botanical analysis from the Shawnee-​ Minisink Paleoindian site (McWeeney 2007b) provides significant evidence for a variety of food resources from a mixed temperate deciduous tree and shrub environment. The following interpretation will be another example of the “subsistence case.” A recent debate arose over the cultural association for the few charred seeds from edible foods such as berries, grapes, hackberry, and winter cress found on the “fire-floors” in the Paleoindian level (Dent 2007; Gingerich 2011). Interpretation requires us to decide whether these foods were part of the Paleoindian subsistence or random seeds burned in a wildfire. No other evidence suggests that there was a wildfire across the site. Is it necessary to claim the seeds as remnants of food consumed at the site, or is it sufficient to know that these plants were available for Paleoindian consumption? The late Pleistocene stratum containing fluted points (two from both excavations), scrapers (n = >60 from the latest excavations), and ­thousands of lithic artifacts was buried under 3 ft of ­sterile sediments (Dent 2007) or 240 cm of alluvial and eolian sediments that show “minimal disturbance” (Gingerich 2011:128). Thus, should charred seeds found outside of the hearth features from 48

Paleoindian Environment and Subsistence Paradigm Case

the Paleoindian level be considered food remains or seed rain? Who establishes the boundary? How near or far from a feature determines cultural affiliation? Certainly, there are examples where proximity and radiocarbon dates do not assure us of contemporariness (e.g., a Dutchess Quarry Cave fluted point and nearby caribou bone later dated to 12,500 14C  bp [see Funk and Steadman 1994]). All of the radiocarbon dates (Dent 2007; Gingerich 2011, and Chapter 9, this volume) validate human occupation at Shawnee-Minisink by 13,000 cal Bp (11,000 14C  bp), coincidental with rapidly cooling temperatures indicated from the Greenland ice sheets. This correlation with the now recognized global impact of the Younger Dryas cold event (Mayewski et al. 1993) is critical to our understanding of late Pleistocene human events along the Upper Delaware River. Moreover, based on the recently identified hawthorn and sweet gum wood charcoal from S­ hawnee-​ Minisink Hearth 2 (McWeeney 2007b), it appears that a warmer climate conducive to deciduous tree growth survived the Killarney cold episode recorded in Nova Scotia at 13,140 cal Bp (11,200 14C  bp [Levesque et al. 1993; Peteet 1995]). Was this the last hurrah for diverse, thermophilous habitats in the Northeast for the next millennium? Or was the Upper Delaware River a refugium from which deciduous trees migrated northeast, up the major river routes, at the end of the Younger Dryas? The evidence for contemporary or later Paleoindian sites and preserved plant remains along the Delaware River would prove invaluable in making this determination. The remarkable assemblage of charred plant macro­ fossils along with the elusive fish bones (Dent and Kauffman 1985; Gingerich 2011; McWeeney 2007b) provide a rare opportunity to describe a Paleoindian settle­ment site on the brink of rapid climate change (Mayewski et al. 1993; Peteet 1995; Peteet et al. 1993). The botanicals identified from the hearths (Gingerich 2011; McWeeney 2007b) and sediments associated with Paleoindian artifacts at Shawnee-Minisink (Dent and Kauffman 1985) are compatible with a temperate environment. Large quantities of plant remains that clearly fall into the category of “subsistence case” for Paleoindians may be ephemeral or elusive or not asso49

ciated with a hearth feature. Consider that time may have eroded precise feature boundaries from short-term habitation sites. Moreover, the preservation of some plant parts available 13,000 years ago for processing and consumption (think roots and tubers) may have remained below ground, where they grew and subsequently deteriorated (Darby 1996; Dillehay and Rossen 2002). Processing the archaeological sediments for phytoliths or starch grains may increase the evidence necessary to resolve some of these unknowns. By extrapolating from the available charred plant macro­ fossils and animal bones that have been reported in a Paleoindian context, the environment surrounding the site and the potential subsistence pattern may become known. However ephemeral the evidence may be, if dated and contemporaneous with the Paleoindian component, the plant macrofossils do validate what was available in the prehistoric environment. We cannot guarantee what was consumed. Strong documentation for presence improves when charred plant remains can be AMS dated and firmly placed in context and chronology. According to Braun (2000), sweet gum (Liquidambar styriciflua) may be found on moist slopes in the “All-Deciduous Mixed Mesophytic Community,” which interlopes within the oak/ chestnut forest. Braun (2000:53–54) sees sweet gum as a relict from other communities or an earlier developmental stage in the deciduous forest. Hawthorn is an indicator of old fields or when found with ash (Fraxinus sp.) may be part of “pioneer communities initiating the slow return to forest” (Braun 2000:397). In the case of S­ hawnee-​ Minisink, is the hawthorn coming or going in the succession process? There is a disjunct species and a pioneering species identified from the charcoal dating between 13,000 and 12,895 cal Bp (ca. 11,000 14C  bp). This combination of hawthorn and sweet gum is another example of the unique postglacial environment. Yet both trees represent a temperate climate, not necessarily a floodplain forest and clearly not a Boreal forest. Was the sweet gum a relic of the forest created during the Bølling/Allerød warming period only to be driven south during the YD? Or had the climate already turned cold, leaving open patches of meadow to be colonized by hawthorn? Whichever scenario one accepts, charred macrofossils from sweet

McWeeney

gum; oak, willow, and pine charcoal; and possible fragments from hickory nutshells plus the hawthorn charcoal and seeds recovered from Hearths 1 and 2 during the recent excavations at S­ hawnee-​ Minisink provide a unique vegetation community to consider as part of the seasonal round of our subsistence case. Sweet gum, according to King’s American Dispensatory (1898), is a large and beautiful tree, which grows well in moist woods, naturally found in Connecticut and New Jersey. Its range expands southward through the Middle Atlantic states and extends farther south to Mexico. The common name of sweet gum is based upon a sap that drains from cuts made into the tree during the warm seasons. Already of syrup consistency, it will form a soft, resin-like substance. Historically, the sap was incorporated into medicinal uses for several ailments (see Moerman 1998 for Native American uses). However, the five- to ­seven-​lobed leaves, which produce a fragrant scent when bruised, turn red in the fall and along with the red hawthorn fruit may have attracted native travelers to the Shawnee-Minisink site for food and healing purposes. Some species of hawthorn fruits remain on the tree far into winter in southern New England (personal observation; Kricher and Morrison 1988:​226–227). Their bright red color could be easily seen in clearings as one traveled along the river. The fruit would have attracted several bird species and other prey such as deer, all of considerable interest to humans. An online source for herbal remedies describes the multiple healing qualities discovered using hawthorn fruits and flowers (http://www​ .armeniapedia.org/index.php?title=Crataegus). Drawing from millennial-age Greek resources up to modern times, a tea made from hawthorn has been used for heart diseases, dizziness, cough, and sleeplessness. Besides being a source for vitamin C, which is well known to prevent scurvy and contribute to strong gums and teeth, the plant has been helpful for relieving throat illnesses and stomach aches (http://www.armeniapedia.org​ /­index​.php?title=Crataegus). While we will never know if Paleoindians were consciously aware of the benefits derived from the hawthorn tree, as a basic food resource they could have been r­ eadily 50

consumed and provided welcome relief from many ailments. Regarding the charred seeds recovered by the American University crew, the variety and low number are similar to those found in the Late Paleoindian component at Dust Cave, Alabama (Hollenbach 2007). The charred amaranth, grape, blackberry/raspberry, hackberry, and bogbean/ buckbean all represent available food and medicinal resources. Significantly, these plants also provide evidence for what grew near to the S­ hawnee-​ Minisink site and clues to the environmental conditions when people lived there around 13,000 calendar years ago. Conclusion

The botanical data recovered from several well-recognized Paleoindian sites spanning from Maine, east to Ohio, and south to Virginia allow us to interpret the paleoenvironmental changes that occurred in the east at the end of the Pleistocene. Initially, after the Laurentide glacier withdrew from southern New England well before 18,000 cal Bp (15,000 14C  bp), a tundra-like environment dominated the vegetation for more than 5,000 calendar years (McWeeney 1994). Arctic/alpine plants such as dwarf birch (Betula michauxii or nana), dwarf willow (Salix herbacea), driads (Dryas integrifolia), and bilberry (Vaccinium uliginosum) blanketed the landscape. During the next 2,000 years, several environmental changes occurred. The temperature warmed, and the first conifer complex of trees appeared between 15,000 and 13,200 cal Bp (12,500 and 12,000 14C  bp). Preserved needles prove that spruce (Picea sp.), larch (Larix laricina), fir (Abies balsamea), and white pine (Pinus strobus) colonized the land along with the temperate deciduous species identified from the regional pollen studies: oak (Quercus sp.), hornbeam (Carpinus carolinia/Ostrya virginiana), ash (Fraxinus sp.), and red maple (Acer rubrum) appear in several pollen spectra reported for southern New England (Gaudreau 1988). Postglacial lakes (Lake Israel, Lake Hitchcock) and ponds (Willowdale Swamp, Pequot Swamp, and Willimantic Kettle) either drained or began to fill in by natural vegetative succession (McWeeney 1994, 1998, 1999, 2007b). Swamps and bogs developed, yielding a diverse assemblage of aquatic

Paleoindian Environment and Subsistence Paradigm Case

and moist-adapted vegetation highly touted for the significant bioresources available (Nicholas 1988). Temperate-adapted animals most likely inhabited southern New England and the Middle Atlantic region at that time (Guilday 1982; Guilday and Parmalee 1982; McWeeney 1994). All of this occurred while we believe the Paleoindians were migrating throughout eastern North America. However, close to 13,000 cal Bp (11,000 14C  bp) numerous climatic indicators point to colder winters and possibly hotter summers. During the interlude known as the Younger Dryas period, 12,900 to 11,900 cal Bp (11,000 to 10,100 14C  bp), a climatic change of global proportions enveloped the region, with a decrease in temperature of at least 3–4°C or more (Peteet et al. 1993; Shuman et al. 2002). Regionally, a lag in vegetation responses such as migration and reproductive changes may have imposed a delay on the dated evidence for this cold regime. The dominant impact experienced by humans, other animals, and plant life alike was a period of cold so severe and extended that it may have forced many animals, such as whitetailed deer, to migrate southward. Despite the cold, there is no evidence of caribou being hunted in southern New England or the Middle Atlantic regions during this period. Fluctuations and warming midway through the Younger Dryas indicated by the GISP2 may correlate with periods when humans may have temporarily expanded their settlement and hunting patterns into northern New England. For instance, Bull Brook and the Hedden site in Maine were occupied around 10,600 14C  bp, and Vail, ca. 10,700 14C  bp (Newby et al. 2005; Robinson et al. 2009; Spiess et al. 1998); and the same holds true for the Maritime provinces as determined from the human occupation dates from Debert in Nova Scotia, which cluster around the middle of the millennium (Levine 1990). However, the cold temperatures resumed for several hundred more years, leaving, to date, no evidence of people who manufactured fluted points until the climatic warming began and temperate species migrated northward. The terminus of the YD is validated by the discovery of white oak charcoal from the Templeton site that was AMS dated to 11,928 ± ​229 cal Bp (10,215 ± ​90 14C  bp, overlapping Moeller’s 1980 date

of 11,895 ± ​497 cal Bp or 10,190 ± ​300 14C  bp). The significant AMS radiocarbon dates from thermophilous botanical remains (oak and white pine) establish the chronology for Late Paleoindian settle­ment (Templeton, Neponset) in southern New England (McWeeney 1994, 2007; Newby et al. 2005). It is my conjecture that the return of people and deciduous trees into southern New England clearly demonstrates that temperatures had ameliorated before 10,000 14C  bp. With more diverse vegetation growing in the region, a re­ advance of Paleoindian settlement began along the coastal and riverine sites in southern New E ­ ngland. The results from this synthesis of plant data from several archaeological sites have illuminated the necessity of recovering plant remains from Paleoindian sites, ideally in features. We can improve our recovery of organic remains by anticipating the need for flotation and by systematically collecting column samples from each unit, whether we see a feature or not (Dent 2007; Dent and Kauffman 1985; Moeller 1980). Only by due diligence will we expand the environmental picture for the late Pleistocene, highlight the cultural settlement patterns, and resolve the subsistence paradigm case. Examination of the charred remains before AMS dating will provide valuable information by documenting precisely what vegetation was growing near a site at a specific time. I­ dentifying plant remains to genus and species level will help establish postglacial rates of vegetation migration and changes in growth ranges that document climatic fluctuations for the last 18,000 years. Examination of the organic remains prior to dating also allows us to rule out contamination from roots and allows incompletely carbonized material from the organics to be radiocarbon dated, as seen at Cactus Hill (McWeeney 2000). In the absence of charred organics or bone, it is possible to obtain an AMS date from phytoliths processed specifically for dating or by using OSL techniques. Employing more than one dating method ensures more reliable chronologies. As science and technology improve we will be able to expand our knowledge of the past and continue to correlate the environmental variables associated with cultural patterns and subsistence. 51

McWeeney

Acknowledgments

New Hampshire. Current Research in the Pleistocene 18:8–10. Braun, Lucy 2000 Deciduous Forests of Eastern North ­America. 2nd ed. Blackburn Press, Caldwell, New ­Jersey. Burns, R. M., and B. H. Honkala 1990 Silvics of North America, Vol. 1: Conifers. Agri­culture Handbook 654. Washington, D.C.: U.S. Department of Agriculture Forest Service. Curran, Mary Lou 1994 Paleoindians in the Northeast: The Problem of Dating Fluted Point Sites. Review of Archaeology 17(1):2–10. Darby, Melissa Cole 1996 Wapato for the People: An Ecological Approach to Understanding the Native American Use of Sagittaria latifolia on the Lower Columbia. Unpublished Master’s thesis, Portland State University. Davis, Margaret B. 1958 Three Pollen Diagrams from Central Massachusetts. American Journal of Science 256:​ 540–570. 1969 Climatic Changes in Southern Connecticut Recorded by Pollen Deposition at Roger’s Lake. Ecology 50(3):409–432. Davis, Robert B., and George L. Jacobson, Jr. 1985 Late Glacial and Early Holocene Landscapes in Northern New England and Adjacent Areas of Canada. Quaternary Research 23:​ 341–360. Dent, Richard J. 1979 Ecological and Sociocultural Reconstruction in the Upper Delaware Valley. Ph.D. dissertation, American University, Washington, D.C. University Microfilms, Ann Arbor. 2007 Seed Collecting and Fishing at the ­Shawnee-​ Minisink Paleoindian Site: Everyday Life in the Late Pleistocene. In Foragers of the Terminal Pleistocene in North America, edited by Renee B. Walker and Boyce N. Driskell, pp. 116–131. University of Nebraska Press, Lincoln. Dent, Richard J., and B. E. Kauffman 1985 Aboriginal Subsistence and Site Ecology as Interpreted from Microfloral Remains. In Shawnee-Minisink: A Stratified Paleoindian– Archaic Site in the Upper Delaware Valley of Pennsylvania, edited by Charles W. McNett, pp. 55–79. Academic Press, New York. Dillehay, Tom D., and Jack Rossen 2002 Plant Food and Its Implications for the ­Peopling of the New World: A View from

I appreciate my colleagues who provided botanical remains or helped me to collect and identify organic remains from archaeological sites and wetland cores while I was a Ph.D. student at Yale University. D ­ orothy Peteet, one of the first American scientists to argue for North American evidence for the Younger Dryas climate change, deserves a huge applause for guiding my work with plant macrofossils and scientific writing. Thank you to those who shared their own botanical results. Not enough can be said to the editor for his knowledge and patience with bringing this information up to date; he deserves more than thanks. I do take responsibility for the interpretations.

Note This chapter is adapted from McWeeney 2007a. According to The Random House Dictionary of the English Language, paradigm case, in the field of logic, means “a perfectly clear and uncontroversial use of a word or expression the meaning of which is being investigated,” and subsistence is the “means of supporting life, a living or livelihood; the source from which food and other items necessary to exist are obtained.” 1. Dates are presented as accelerator mass spectrometry dates in Maenza-Gmlech 1997 and were calibrated using CalPal 3.2.

References Cited Amundson, D. C., and Herbert E. Wright, Jr. 1979 Forest Changes in Minnesota at the End of the Pleistocene. Ecological Monographs 49(1):1–16. Anderson, David G. 2005 Pleistocene Human Occupation of the Southeastern United States: Research Directions for the Early 21st Century. In ­Paleoamerican Origins: Beyond Clovis, edited by ­Bradley T. Lepper, Robson Bonnichsen, Dennis Stanford, and Michael R. Waters, pp. 29–42. Center for the Study of the First Americans, Texas A&M University Press, College Station. Boisvert, Richard A. 1999 Paleoindian Occupation of the White Mountains, New Hampshire. Geographie Phyique et Quaternaire 53(1):159–174. 2004 Clovis-Era Archaeology in Northern New Hampshire: The Israel River Complex. In New Perspectives on the First Americans, edited by Bradley T. Lepper and Robson Bonnichsen, pp. 49–54. Center for the Study of the First Americans, Texas A&M, College Station. Boisvert, Richard A., and K. Puseman 2001 The Jefferson IV Site (27-CO-45), Jefferson, 52

Paleoindian Environment and Subsistence Paradigm Case South America. In The First Americans: The Pleisto­cene Colonization of the New World, edited by Nina Jablonski, pp. 237–254. Memoirs of the California Academy of Science, San Francisco. Dincauze, D. F. 1993 Pioneering in the Pleistocene: Large Paleoindian Sites in the Northwest. In Archaeology of Eastern North America: Papers in Honor of Stephen Williams, edited by James B. Stoltman, pp. 43–60. Archaeological Report No. 25. Mississippi Department of Archives and History, Jackson. 2000 The Northeast. Common Ground, Spring/ Summer: 34–43. Driskell, Boyce N., and Renee B. Walker 2007 Making Sense of Paleoindian Subsistence Strategies. In Foragers of the Terminal Pleistocene in North America, edited by Renee B. Walker and Boyce N. Driskell, pp. 226–238. University of Nebraska Press, Lincoln. Feathers, J. K., E. J. Rhodes, S. Huot, and J. M. McAvoy 2006 Luminescence Dating of Sand Deposits Related to Late Pleistocene Human Occupation at the Cactus Hill Site, Virginia, USA. Quaternary Geochronology 1:167–187. Fiedel, Stuart J. 1999 Older Than We Thought: Implications of Corrected Dates for Paleoindians. American Antiquity 64(1):95–115. 2004 Clovis Age in Calendar Years: 13,500–13,000 CALYBP. In New Perspectives on the First Americans, edited by Bradley T. Lepper and Robson Bonnichsen, pp. 73–80. Texas A&M, College Station. 2007 Quacks in the Ice: Waterfowl, Paleoindians, and the Discovery of America. In Foragers of the Terminal Pleistocene in North America, edited by Renee B. Walker and Boyce N. Driskell, pp. 1–14. University of Nebraska Press, Lincoln. Fisher, Daniel C. 2004 Mastodons, Mammoths, and Humans in the North American Mid-Continent. In New Perspectives on the First Americans, edited by Bradley T. Lepper and Robson Bonnichsen, pp. 81–86. Center for the Study of the First Americans, Texas A&M, College Station. Forrest, Daniel T. 1999 Preliminary Results from Recent Excavations at the Sandy Hill Site. Paper presented at the 39th Annual Meeting of the Northeastern Anthropological Association, March 20, Rhode Island College. 2000 Beyond Presence and Absence: Establishing

Diversity in Connecticut’s Early Holocene Archaeological Record. Bulletin of the Archaeological Society of Connecticut 62:79–100. Funk, Robert E., and David W. Steadman 1994 Archaeological and Paleoenvironmental Investigations in the Dutchess Quarry Caves, Orange County, New York. Persimmon Press, Buffalo. Gaudreau, Denise R. 1986 Late-Quaternary Vegetational History of the Northeast: Paleoecological Implications of Topographic Patterns in Pollen Distributions. Unpublished Ph.D. dissertation, Department of Biology, Yale University. 1988 The Distribution of Late Quaternary Forest Regions in the Northeast: Pollen Data, Physiography, and the Prehistoric Record. In Holocene Human Ecology in Northeaster North America, edited by George P. Nicholas, pp. 215–256. Plenum Press, New York. Gaudreau, Denise R., and Thompson Webb, III 1985 Late-Quaternary Pollen Stratigraphy and Isochrone Maps for the Northeastern United States. In Pollen Records of Late-Quaternary North American Sediments, edited by Vaughn M. Bryant, Jr., and R. C. Holloway, pp. 247– 280. American Association of Stratigraphic Palynologists Foundation, Dallas. Gingerich, Joseph A. M. 2011 Down to Seeds and Stones: A New Look at the Subsistence Remains from ­Shawnee-​ Minisink. American Antiquity 76(1):127–144. Goodyear, Albert C. 2000 Topper Site: Results of the 2000 Allendale Paleoindian Expedition. Legacy 5(2):18–25. 2005 Evidence for Pre-Clovis Sites in the Eastern United States. In Paleoamerican Origins: Beyond Clovis, edited by Robson Bonnichsen, Bradley T. Lepper, Dennis Stanford, and Michael R. Waters, pp. 103–112. Texas A&M University Press, College Station. Gove, H. E. 1992 The History of AMS, Its Advantages Over Decay Counting: Applications and Prospects. In Radiocarbon After Four Decades: An Interdisciplinary Perspective, edited by R. E. Taylor, A. Long, and R. S. Kra, pp. 214–229. ­Springer-​ Verlag, New York. Grimes, John R., W. Eldridge, B. G. Grimes, A. Vaccaro, F. Vaccaro, J. Vaccaro, N. Vaccaro, and A. Orsini 1984 Bull Brook II. Archaeology of Eastern North America 12:159–183. Grushkin, Dan 2010 Footsteps of Our Ancestors. Archaeology 63​ (4):13. 53

McWeeney ument, http://www.henriettesherbal.com​ /­eclectic/kings/liquidambar.html. Kricher, J. C., and G. Morrison 1988 Field Guide to Eastern Forests of North America. Houghton-Mifflin Co., Boston. Leduc, Phil 2003 Pollen Viewer 3.2. Electronic document, http://www.ncdc.noaa.gov/paleo/pollen​ /­viewer/webviewer.html. Levesque, Andre, J. F. E. Mayle, Ian R. Walker, and Leslie C. Cwynar 1993 A Previously Unrecognized Late-­Glacial Cold Event in Eastern North America. ­Nature 361:623–626. Levine, M. A. 1990 Accommodating Age: Radiocarbon Results and Fluted Point Sites in Northeastern North America. Archaeology of Eastern North America 18:33–63. Little, Elbert L., Jr. 1977 Atlas of United States Trees. Vol. 1: Conifers and Important Hardwoods. Miscellaneous Publication 1146. U.S. Department of Agriculture, Forest Service, Washington, D.C. Lothrop, Jonathan C., Paige Newby, Arthur Spiess, and James Bradley 2011 Paleoindians and the Younger Dryas in the New England–Maritimes Region. Quaternary International 242(2):546–569. Maenza-Gmelch, T. 1997 Vegetation, Climate, and Fire During the Late Glacial–Holocene Transition at Spruce Pond, Hudson Highlands, Southeastern New York. Journal of Quaternary Science 12(1):​ ­15–24. Madella, Marco, and Debora Zurro (editors) 2007 Plants, People and Places: Recent Studies in Phytolith Analysis. Oxbow Books, Oxford. Mayewski, P. A., L. D. Meeker, S. Whitlow, M. S. Twickler, M. C. Morrison, R. B. Morrison, R. B. Alley, P. Bloomfield, and K. Taylor 1993 The Atmosphere During the Younger Dryas. Science 261:195–197. Mayle, F., and L. C. Cwynar 1995 A Review of Multi-Proxy Data for the Younger Dryas in Atlantic Canada. Quaternary Science Reviews 14:813–821. McAvoy, Joseph M., James C. Baker, James K. ­Feathers, R. L. Hodges, Lucinda J. McWeeney, and Thomas R. Whyte 2000 Summary of Research at the Cactus Hill Archaeological Site, 44SX202, Sussex County, Virginia. Report to the National Geographic Society in compliance with Grant #6345-98, on file with an author.

Guilday, John E. 1982 Appalachia 11,000–12,000 Years Ago: A Biological Review. Archaeology of Eastern North America 10:22–25. Guilday, John E., and P. W. Parmalee 1982 Vertebrate Faunal Remains from Meadowcroft Rockshelter, Washington County, Pennsylvania: Summary and Interpretation. In Meadowcroft Rockshelter and the Cross Creek Drainage, edited by Ronald C. Carlisle and James M. Adovasio, pp. 163–174. Department of Anthropology, University of Pittsburg, Pittsburg. Hare, T. 1994 Habitats. Macmillan Press, New York. Hollenbach, Kandace D. 2007 Gathering in the Late Paleoindian Period: Archaeobotanical Remains from Dust Cave, Alabama. In Foragers of the Terminal Pleistocene in North America, edited by Renee B. Walker and Boyce N. Driskell, pp. 132–147. University of Nebraska Press, Lincoln. Jackson, S. T., and J. W. Williams 2004 Modern Analogs in Quaternary Paleoecology: Here Today, Gone Yesterday, Gone Tomorrow? Annual Review of Earth and Planetary Sciences 32:495–537. Jones, Brian D. 2000 The Late Paleoindian Hidden Creek Site in Southwestern Connecticut. Archaeology of Eastern North America 25:45–80. Kellogg, D. C. 1991 Prehistoric Landscapes, Paleoenvironments, and Archaeology of Western Muscongus Bay, Maine. University Microfilms, Ann Arbor. Kellogg, D. C., and Jay Custer 1995 Paleoenvironmental Studies of the State Route 1 Corridor: Contexts for Prehistoric Settlement. Delaware Department of Transportation Archaeology Series No. 114. Dover. Kellogg, D. C., and K. Simons 2000 Phase III Date Recovery Excavations at the Neal Garrison Paleoindian Site (ME 1.8), ­Eliot, York County, Maine. Report prepared for the Portland Natural Gas Transmission System. John Milner Associates, Inc., West Chester, Pennsylvania. Kitchel, Nathaniel 2008 Charred Plant Remains from the Colebrook Site: Paleoindian Paleoethnobotany on New Hampshire’s Northwest Coast. Unpublished Master’s thesis, Northern Arizona State University. King’s American Dispensatory Liquidambar. — ​Sweet-Gum. Electronic doc1898 54

Paleoindian Environment and Subsistence Paradigm Case Boyce N. Driskell, pp. 148–166. University of Nebraska Press, Lincoln. 2007b Shawnee Minisink Site — A Revival of Botanical Excavations. Report prepared for Joseph A. M. Gingerich. Manuscript on file, Seneca Falls, New York. McWeeney, Lucinda J., and Douglas Kellogg 2001 Early and Middle Holocene Climate Changes and Settlement Patterns Along the Eastern Coast of North America. Archaeology of Eastern North America 29:187–211. Meltzer, D. J. 2009 First Peoples in a New World: Colonizing Ice Age America. University of California Press, Berkeley. Miller, Norton G. 1994 Snapshot Paleobotanical Analyses of Late-​ Glacial Sediments in the Connecticut River Valley, Northern Vermont to Central Connecticut. Geological Society of America, Abstracts with Programs, Northeastern Section 27:69. Moeller, Roger W. 1980 6LF21: A Paleoindian Site in Western Connecticut. American Indian Archaeological ­Institute, Washington, Connecticut. Moerman, D. E. 1998 Native American Ethnobotany. Timber Press, Portland, Oregon. Newby, P., James Bradley, Arthur Spiess, Bryan Shuman, and Phillip Leduc 2005 A Paleoindian Response to Younger Dryas Climate Change. Quaternary Science Reviews 24:141–154. Newby, P., J. P. Donnelly, and B. Shuman 2009 Evidence of Centennial-Scale Drought from Southeastern Massachusetts During the Pleistocene/Holocene Transition. Quaternary Science Reviews 28:1675–1692. Newby, P., B. N. Shuman, J. P. Donnelly, and D. MacDonald 2011 Repeated Century-Scale Droughts Over the Past 13,000 yr near the Hudson River ­Watershed, USA. Quaternary Research 75:523–530. Nicholas, George E. 1988 Ecological Leveling: The Archaeology and Environmental Dynamics of Early Post Glacial Land Use. In Holocene Human ­Ecology in Northeastern North America, edited by George P. Nicholas, pp. 257–296. Plenum Press, New York. Overstreet, D. F. 2005 Late Glacial Ice Marginal Adaptation in Southeastern Wisconsin. In ­Paleoamerican

McAvoy, Joseph M., and Lynn D. McAvoy 1997 Archaeological Investigations of Site 44SX202, Cactus Hill, Sussex County, Virginia. Nottoway River Survey Archaeological Research Report No. 2 and Research Report Series No. 8. Department of Historic Resources, Richmond, Virginia. McWeeney, Lucinda J. 1991 Plant Macrofossil Identification as a Method Toward Archaeo-Environmental Reconstruc­ tion. Bulletin of the Archaeological Society of Connecticut 54:87–97. 1994 Archaeological Settlement Patterns and Vegetation Dynamics in Southern New England in the Late Quaternary. University Microfilms, Ann Arbor. 1996 Charcoal Evidence for Human Modification of the Landscape in Southern New England. Paper presented at the 61st Annual Meeting of the Society for American Archaeology, New Orleans. 1997 Report on the Charcoal Remains from the Cactus Hill Site, Nottoway River, Virginia. In Archaeological Investigations of Site 44SX202, Cactus Hill, Sussex County, Virginia, edited by Joseph M. McAvoy and Lynn D. McAvoy, Appendix D(1). Nottoway River Survey Archaeological Research Report No. 2 and Research Report Series No. 8. Department of Historic Resources, Richmond, ­ irginia. V 1998 Pequot Cedar Swamp Paleoenvironment. In Coastally Restricted Forests, edited by Aimlee Laderman, pp. 124–141. Oxford University Press, New York. 1999 A Review of Late Pleistocene and Holocene Climate Changes in Southern New England. Bulletin of the Archaeological Society of Connecticut 62:213–218. 2000 Comparing Feature and Non-Feature Botanical Remains from the Neal Garrison Site. In Phase HI Data Recovery Excavations at the Neal Garrison Paleoindian Site (ME 1.8), ­Eliot, York County, Maine, Appendix XL. Report prepared for the Portland Natural Gas Transmission System. John Milner Associates, Inc., West Chester, Pennsylvania. 2001 The Role of Sediment Analyses in Establishing Human Presence at Early Archaeological Sites. Current Research in the Pleistocene 18:44–46. 2007a Revising the Paleoindian Environmental Picture in Northeastern North America. In Foragers of the Terminal Pleistocene in North America, edited by Renee B. Walker and 55

McWeeney Origins: Beyond Clovis, edited by R. Bonnichsen, B. T. Lepper, D. Stanford, and M. R. Waters, pp. 183–195. Center for the Study of the First Americans, Texas A&M University Press, College Station. Pearsall, Deborah 2000 Handbook of Paleoethnobotany. 2nd ed. ­Academic Press, New York. Peteet, D. M. 1995 Global Younger Dryas? Quaternary International 28:93–104. Peteet, D. M., R. A. Daniels, L. E. Heusser, J. S. Vogel, J. R. Southon, and D. Nelson 1993 Late Glacial Pollen, Macrofossils, and Fish Remains in Northeastern U.S.A.: The Younger Dryas Oscillation. Quaternary Science Review 12:597–612. Peteet, D. M., J. S. Vogel, D. E. Nelson, J. R. Southon, R. J. Nickmann, and L. E. Heusser 1990 Younger Dryas Climatic Reversal in Northeastern USA? AMS Ages for an Old Problem. Quaternary Research 33:219–230. Perry, David 2012 Understanding Plant Use: The Importance of Parenchymatous Tissues. Manuscript on file, Mashantucket Pequot Museum and Research Center, Mashantucket, Connecticut. Perry, David, and Brian Jones 2002 A Synopsis on Recent Identification of Soft Plant Tissues from Prehistoric Contexts at Mashantucket. Paper presented at the 25th Annual Ethnobiology Conference, University of Connecticut, Storrs. Perry, David, and Kevin A. McBride 2012 Vegetative Plant Remains from Site 72. Manu­script on file with author. Piperno, Dolores R. 1988 Phytolith Analysis: An Archaeological and Geological Perspective. Academic Press, ­London. Powers-Jones, A. H., J. Padmore, and D. D. Gilbertson 1989 Studies of Late Prehistoric and Modern Opal Phytoliths from Coastal Sand Dunes and Machair in Northwest Britain. Journal of ­Archaeological Science 16:27–45. Ridge, J. C. 2003 The Last Deglaciation of the Northeastern United States: A Combined Varve, Paleo­ magnetic, and Calibrated 14C C ­ hronology. In Geoarchaeology of Landscapes in the Glaciated Northeast, edited by L. David-­Cremeens and John P. Hart, pp. 15–48. New York State Museum Bulletin, Vol. 497. New York.

Robinson, Brian S., Jennifer C. Ort, William E. ­ ldridge, Adrian L. Burke, and Bertrand G. Pelletier E 2009 Paleoindian Aggregation and Social Context at Bull Brook. American Antiquity 74(3):423– 447. Shane, Linda C. K., and K. H. Anderson 1993 Intensity, Gradients, and Reversals in Late Glacial Environmental Change in East-­ Central North America. Quaternary Science Reviews 12:307–320. Shuman, B., P. J. Bartlein, and T. Webb, III 2007 Response to “Comments on: ‘The Magnitude of Millennial- and Orbital-Scale Climatic Change in Eastern North America During the Late-Quaternary’ by Shuman et al.” Quaternary Science Reviews 26:268–273. Shuman, B., P. Newby, Y. Huang, and T. Webb, III 2004 Evidence for the Close Climatic Control of New England Vegetation History. Ecology 85(5):1297–1310. Shuman, B., Thompson Webb, III, Patrick Bartlein, and John W. Williams 2002 The Anatomy of a Climatic Oscillation: ­Vegetation Change in Eastern North America During the Younger Dryas Chronozone. Quaternary Science Reviews 21:1777–1791. Sneddon, L. A. 1987 A Late-Glacial and Post-Glacial Vegetation History of Black Pond, Norwell, Massachusetts. Unpublished Master’s thesis, University of Massachusetts, Boston. Sneddon, L., and Lawrence Kaplan 1987 Pollen Analysis from Cedar Swamp Pond, Westborough, Mass. Archaeological Quarterly of the W. Elmer Ekblaw Chapter, Massachusetts Archaeological Society 9(2):1–15. Spear, R. W., M. B. Davis, and L. C. K. Shane 1994 Late Quaternary History of Low- and Mid-Elevation Vegetation in the White Mountains of New Hampshire. Ecological Monographs 64:85–109. Spiess, Arthur E., and Debra Wilson 1987 Michaud: A Paleoindian Site in the New ­England–Maritimes Region. Occasional Publication in Maine Archaeology, No. 6. Spiess, Arthur E., Debra Wilson, and James W. ­Bradley 1998 Paleoindian Occupation in the New ­England–​Maritimes Region: Beyond Cultural Ecology. Archaeology of Eastern North America 26:201–264. Stafford, Thomas W., Jr. 1999 AMS Dating: Better Methods to Achieve Bet-

56

Paleoindian Environment and Subsistence Paradigm Case nal Remains from Dust Cave, Alabama. In Foragers of the Terminal Pleistocene in North America, edited by Renee B. Walker and Boyce N. Driskell, pp. 99–115. University of Nebraska Press, Lincoln. Waters, Michael R., and Thomas W. Stafford, Jr. 2007 Redefining the Age of Clovis: Implications for the Peopling of the Americas. Science 315(5815):1122. Watts, William A. 1979 Late Quaternary Vegetations of Central Appalachia and the New Jersey Coastal Plain. Ecological Monographs 49(4):427–469. Webb, Robin S. 1990 Late Quaternary Water-Level Fluctuations in the Northeastern United States. Unpublished Ph.D. dissertation, Department of Geology, Brown University, Providence. Williams, J. W. 2003 Needleleaved and Broadleaved Tree Cover Distributions in North America Since the Last Glacial Maximum. Global and Planetary Change 35:1–23. Wisner, George 1998 Ohio Cave, Sealed Since Ice Age, Yields Data on Paleo-Americans. Mammoth Trumpet 13(1):1, 8–11. Yesner, David R. 2007 Faunal Extinction, Hunter-Gatherer Foraging Strategies, and Subsistence Diversity Among Eastern Beringian Paleoindians. In Foragers of the Terminal Pleistocene in North America, edited by Renee B. Walker and Boyce N. Driskell, pp. 15–31. University of Nebraska Press, Lincoln. Yu, Z. 2007 Rapid Response of Forested Vegetation to Multiple Climatic Oscillations During the Last Deglaciation in the Northeastern United States. Quaternary Research 67(2):297–303.

ter Dates. Paper presented at the “Clovis and Beyond” Conference, Santa Fe, New Mexico. Stea, Ralph R., and Robert J. Mott 1989 Deglaciation Environments and Evidence for Glaciers of Younger Dryas Age in Nova Scotia, Canada. Boreas 18:169–187. Stone, Janet R., and Gail M. Ashley 1992 Ice-Wedge Casts, Pingo Scars and the Drainage of Glacial Lake Hitchcock. New England Intercollegiate Geological Conference, 84th Annual Meeting, Guidebook for Field Trips in the Connecticut Valley Region of Massachusetts and Adjacent States. Contributions of the University of Massachusetts Department of Geology 66:305–331. Storck, Peter L., and Arthur E. Spiess 1994 The Significance of New Faunal Identifications Attributed to an Early Paleoindian (Gainey Complex) Occupation at the Udora Site, Ontario, Canada. American Antiquity 59(1):​121–142. Stuiver, M. T., and P. J. Reimer 1993 Extended 14C Data Base and Revised CALIB3.0 14C Age Calibration Program. ­Radiocarbon 35(1):215–230. Tankersley, Kenneth B. 1994 Sheriden: A Clovis Cave Site in Eastern North America. Geoarchaeology: An International Journal 12:713–724. 1999 Sheriden: A Stratified Pleistocene–Holocene Cave Site in the Great Lakes Region of North America. BAR International Series 800:67–75. Tankersley, Kenneth B., and B. Redman 1999 Radiocarbon Dating of a Projectile Point from Sheriden Cave, Ohio. Current Research in the Pleistocene 16:76–77. Thorson, Robert M., and Robyn Webb 1991 Postglacial Development of the Cedar Swamp. Journal of Paleolimnology 6:17–35. Walker, Renee B. 2007 Hunting in the Late Paleoindian Period: Fau-

57

3

Reconstructing the Pleistocene Environment of the Greater Southeast Jessi J. Halligan

Although controversy surrounds the earliest human arrival in the Southeast (Bradley and Stanford 2004; Faught 2008; Goodyear 2005; Haynes 2005), people were well established during the terminal Pleistocene. Paleoindians would have seen changes in the entire landscape caused by sea level rise, water table fluctuation, changes in floral biomes, and a complete transformation of available fauna. This chapter discusses the paleoenvironmental record from the Southeast during the terminal Pleistocene and early Holocene, spanning the period from the Last Glacial Maximum (LGM) at approximately 24,000–20,000 cal Bp (Ehlers and Gibbard 2004; McKay et al. 2008; but see Peltier and Fairbanks 2006) to 10,000 cal Bp, when modern or nearly modern conditions were established in most of the Southeast. In this chapter, “the Southeast” refers to the area of the modern states of Louisiana, Arkansas, Mississippi, Georgia, Alabama, Florida, North Carolina, South Carolina, Tennessee, Kentucky, Virginia, West Virginia, Maryland, and Delaware (Figure 3.1). This division is somewhat arbitrary but reflects modern climate patterns and general usage of the term. Paleoenvironment, as used below, will be as inclusive as possible, including climate, geography, flora, and fauna. Throughout this chapter, ages are presented in calendar years, as several proxy records and dating methods only give calendar ages. Thus, if publications presented calendar ages, these were used; if only uncalibrated radiocarbon ages were available, ages

were calibrated using OxCal version 4.1.4 (IntCal 09 curve) to give an approximate age before present (Bronk Ramsey 2009, 2010). Paleoenvironmental Problems in the Southeast

Paleoenvironmental studies are always difficult: Data sets are spotty, chronological control is often poor, proxy records may or may not represent past environments, and interpretations of past floral and faunal records are frequently contro­ versial or contradictory (Dahms 1998; Dahms and Holliday 1998; Delcourt and Delcourt 1991; Dincauze 2000; McDonald and Bryson 2010). The Southeast provides an extra challenge to those who would look to reconstruct its environmental history (Delcourt and Delcourt 1985). The Southeast’s highly varied geography has resulted in many different ecological zones, including plains and forests, mountain streams and coastal estuaries, rocky mountaintops and alluvial floodplains. Evidence suggests that the past was similarly complex, including scattered refugia where the local environment may have changed little during the past 20,000 years (Delcourt and Delcourt 1985; Gonzales et al. 2009). This heterogeneous geography has also led to a very spotty paleoenvironmental record. There are rare areas with good preservation in some areas, but most of the terminal Pleistocene and early Holocene record is missing in many places. Northern Florida is pocked with sinkholes con58

Reconstructing the Pleistocene Environment of the Greater Southeast

Figure 3.1. Southeastern area during the terminal Pleistocene including currently submerged land and locations named in the text. The dark gray band refers to landscape actively submerged during the Clovis time period (Waters and Stafford 2007) using the sea level curves in Figure 3.2.

taining late Pleistocene fauna and deep sediment records, the Atlantic Coastal Plain lakes contain their own plant fossils, and there are scattered bogs and peat lenses throughout the Southeast. Many upland and inland areas, however, do not have preserved landforms or fossils, so paleoenvironmental reconstructions are based largely on extrapolation from nearby areas. A sequence from a single v­ alley, thus, may become the proxy record for an entire state due to lack of data, meaning that both precision and accuracy are compromised. Nevertheless, more than 50 years of research have resulted in a general understanding of past environments in the Southeast on a coarse scale. The Southeast during the terminal Pleistocene was unlike anywhere today. The landscape was a patchy mosaic of rapidly changing flora and

fauna, with coastal areas being flooded by incoming seas and climates more variable than any seen today. All of these environmental variables interacted with and impacted one another, creating a complicated system of change and response. Paleoclimatology Glacial Ice Volume: Glacials, Interglacials, and Heinrich Events

The Pleistocene is defined by the presence of glaciation; the beginning of the Holocene at approximately 12,000 cal Bp is thus defined by the ending of continent-wide glaciers and the ­warming of the climate to near-modern conditions. ­Glacial retreat was not a linear process (Figure 3.2); from the LGM onward, there were several g­ lacial advances, which corresponded to lowered sea levels 59

Figure 3.2. Terminal Pleistocene and Holocene temperature proxy records showing notable climatic events, plotted using original data from the references cited in the figure.

Reconstructing the Pleistocene Environment of the Greater Southeast

and generally cooler and moister ­climates in large portions of the Southeast (Clauzet et al. 2007; Peltier and Fairbanks 2006). Glacial advances and retreats were probably caused by a complicated interaction of oscillation in the Earth’s orbit (Milankovitch cycles) and rotation of the Earth’s axis (see Hays et  al. 1976 for a c­ lassic discussion). Glaciations were highly cyclical, but within full-​­glacial times, there were cooler periods followed by abrupt warming, known as Dansgaard– Oeschger events (Dansgaard et  al. 1984). This warming would lead to massive glacial calving in the North Atlantic, causing Heinrich events (Heinrich 1988), increased percentages of icerafted glacial debris in ocean sediments occurring at approximately 10,000-year intervals. Because of this melting, large amounts of freshwater were discharged into sea currents, disturbing circulation patterns and possibly causing the next cold event (Bond et al. 1992), although see Bigg and colleagues (2011) for a summary of complications with these models. Most important for Paleoindian archaeology, glacier extents impacted when and how people colonized the Americas and probably affected when and how people arrived in the Southeast (Anderson and Gillam 2000; Dixon 1999, 2001; Faught 2008; Fiedel 2000; Kelly 2003; Mandryk et al. 2001; Yesner 2001). The most recent climatic reversal, the Younger Dryas, is especially relevant to the study of southeastern Paleoindians because it may have been observable by people and has been argued to have caused the end of Clovis and/ or the Pleistocene megafauna in North America (Dunbar 2006; Firestone et al. 2007; Scott 2010; Semken et al. 2010). This period occurred roughly between 12,900 and 11,500 cal Bp and was marked by significant glacial advance and evidence for a return to full glacial conditions in northern climes, but there has been much recent debate about its duration, intensity, causation, and impact on humans (Anderson et al. 2011; Firestone et al. 2007; Holliday and Meltzer 2010 and comments; Meltzer and Holliday 2010; Newby et al. 2005; Straus and Goebel 2011; Surovell et al. 2009).

related to glacial advance and retreat, including the aforementioned Heinrich events. Isotopic data from extant Pleistocene ice can be used to infer environmental conditions when the ice was formed. Relative deuterium percentages (heavy hydrogen) collected from Antarctic ice cores and oxygen-18 percentages from Greenland ice cores have been correlated to glacial ice advance and retreat (Johnsen et al. 2001; Jouzel and EPICA Community Members 2004; NGRIP Members 2004; Petit et al. 1999). These curves can be very accurately read, but as shown in Figure 3.2, they can be somewhat difficult to interpret. Relative sea level curves also can help give an approximate indication of the amount of ice contained in glaciers (Balsillie and Donoghue 2004; Peltier and Fairbanks 2006). These proxy records only give information on total ice volume; it then becomes necessary to interpret how advances and retreat applied to paleo­ climate and people. Dunbar (2006) has written an excellent summary of the many types of issues that archaeologists need to consider when interpreting paleoenvironmental reconstructions. He discusses how global circulation and ice volume models can be applied to local climatic reconstructions of the Southeast during the terminal Pleistocene. For instance, lake core microbotanical research (Grimm et al. 2006) demonstrates warmer, wetter climate regimes in Florida during North Atlantic cold phases, possibly due to thermohaline circulation patterns. Geological Records Sea Level and Hydrologic Records

Sea level rise caused by glacial melting had a dramatic effect on the Southeast during the terminal Pleistocene. Sea levels were approximately 120 m lower during the LGM and 50–75 m lower during the Clovis period (12,800–13,250 cal Bp [Balsillie and Donoghue 2004; Peltier and Fairbanks 2006; Waters and Stafford 2007]). Roughly 25 percent of the land area of the Southeast has been lost due to sea level rise since the LGM (2.3 million km2 at the beginning of the LGM to approximately 1.75 million km2 today; Figure 3.1). Numerous sea level curves for the area show sea level rise occurring in punctuated bursts (Figure 3.2), with the fastest rate of rise occurring when the Mississippi River was an outlet for meltwater

Proxy Records

There is no direct way to measure the past environment or past glacial ice volume, but Figure 3.2 shows several proxy records that have been 61

Halligan

pulses from glacial Lake Agassiz prior to 12,900 cal Bp (Balsillie and Donoghue 2004; Fisher 2003; Peltier and Fairbanks 2006), although this rapid rise was probably not entirely due to Mississippi drainage (Clark et al. 1996; Donnelly et al. 2005). Inland areas were affected by these rising sea levels in more ways than simple submergence. Rising water levels refilled deep aquifers, stream channels, and lake beds. This increased available surface water and dramatically transformed many river channels from high-energy braided patterns to gently meandering systems, modifying entire riparian zones (Blum et al. 2008; Leigh et al. 2004; Rittenour et  al. 2007). Dry inland and upland areas became coastal swamps, such as the current Delmarva Peninsula (Lowery et al. 2010). Some low-lying inland areas became freshwater lakes or swamps, while others dried out from changing hydrologic forces (Dunbar 2006; Grimley et al. 2009; Grimm et al. 2006). Farther upland areas also experienced indirect impacts from changing moisture regimes in the lowlands and changing rainfall patterns (Dunbar 2002). Many areas experienced drier conditions due to lowered local water tables, resulting in common valley incision and erosion sequences, fluctuating lake levels, and missing sediment packages (Driese et al. 2005; Faure et al. 2002; Sherwood et al. 2004; Webb and Dunbar 2006). Recent research also indicates that several droughts in the lower Southeast had major impacts on local water tables and local plant and animal communities, even with an overall trend of sea level rise and increasing moisture (Dunbar 2002, 2006; Newsom 2006; Thulman 2009).

cal Bp. This soil was covered by an overstory similar to boreal forest and experienced commonly fluctuating water tables during its formation; the next stratum dates to before 7000 cal Bp. They interpret this nearly 15,000-radiocarbon-year hiatus in the sediment package as representing a geomorphically active period with common incision, erosion, and downstream aggradation of sediment. Leigh and colleagues (2004) have discovered convincing evidence of braided stream sequences with eolian dunes on the Atlantic Coastal Plain of Georgia dating from 30,000 to 17,000 cal Bp. These sequences may indicate cooler and drier paleoclimates, probably with thinner vegetation. Stream systems in Georgia and the Carolinas changed from braided channels to meandering streams coincident with expanding forests and rapid climate warming from about 16,000 to 11,000 cal Bp (Leigh 2008). This is supported by missing sediment sequences at the Sandy Run Creek site, on the Upper Coastal Plain of Georgia, where there is an erosional unconformity from approximately 25,000 to 13,000 cal Bp (LaMoreaux et al. 2009). Dust Cave in northern Alabama contains one of the most intact and best-dated cultural sequences from the latest Pleistocene and early Holocene in the Southeast (Sherwood et al. 2004). Prior to 17,700 cal Bp, water levels were very high, and the cave was filled with sediment. At this time, base level began to lower on the Cumberland Plateau, leading to major erosional events in many of the local caves. Starting at approximately 13,000 cal Bp until approximately 11,000 cal Bp, spring activity within the cave flushed sediments toward the mouth of the cave, mixing extinct Pleistocene and extant Holocene fauna, while periodic flooding brought in fine-grained alluvial sediments, causing accumulation of deposits. The Aucilla River in northern Florida contains numerous sinkholes with sediment sequences and cultural materials from the terminal Pleistocene. Several researchers (Donoghue 2006; Dunbar 2002, 2006; Thulman 2009; Webb 1998; Webb and Dunbar 2006) have proposed that during the lower sea levels and congruent lower water tables of the Pleistocene–Holocene transition, the Aucilla and Wacissa rivers did not flow. However, many of the deeper sinks in the modern channel contained springs, making them conve-

Sediment Records

The paleoclimatology and paleohydrology discussions above are generalized for the entire Southeast. It is not possible to generalize sediment records to such an extent. Highly local variables such as slope, vegetation, grain size, bedrock type, and local climate determine sediment characteristics and whether sediments are even preserved. In general, however, the sediment records from the terminal Pleistocene and early Holocene indicate a period of great geomorphic activity. Driese and colleagues (2005) have recorded a stable, full-glacial paleosol in southeastern West Virginia dating to approximately 28,000–27,000 62

Reconstructing the Pleistocene Environment of the Greater Southeast

nient water sources that attracted people and animals. More than 90 radiocarbon dates have been collected from geological sections in the Aucilla River that indicate at least three different cycles of sinkhole infilling dating to 46,000–40,000 cal Bp, 37,000–28,000 cal Bp, and 18,600–9500 cal bp (Dunbar 2006; Webb 1998). These are inferred to indicate drier climates with less vegetation cover, allowing for sinkhole exposure and sediment i­ nflux.

individually to changing climate, with some responses more rapid than others, so plant communities were virtually always in flux (Gonzales et al. 2009; Williams et al. 2004; Williams et al. 2010). However, this same research has given generalized plant associations for eastern biomes in 1,000-calendar-year intervals for the past 21,000 years (Leduc 2003; Williams et al. 2004), which have been adapted in Figures 3.3 and 3.4 to display what is currently known about regional plant communities during the terminal Pleistocene, with the modern environment shown for reference. These figures show generalized biome associations throughout the late Pleistocene. Maps within Williams et al. (2004) show similar plots for individual plant species, which graphically display the dynamism of forest response to climate change. The biome maps in Figure 3.3 mask much of this dynamic response by focusing on plant function within the biome rather than individual species composition, but the biomes shown help explicate large-scale environmental change. Williams and colleagues (Williams et al. 2000; Williams et al. 2004; Williams et al. 2010) define these biomes in detail and justify the plant functional associations used, which is beyond the scope of this chapter. For instance, cold deciduous forests include boreal summergreen trees, eurythermic conifers, and heaths. The “boreal summergreen” category includes Alnus, Betula, Corylus, Larix/Pseudotsuga, Myricaceae, Populus, and Salix pollens. These very coarse-grained discussions can be refined in some areas by more detailed studies. Boreal forest sequences were recorded in White Pond, South Carolina, prior to 14,800 cal Bp, followed by oak-hickory-beech associations from approximately 14,800 cal Bp to approximately 11,000 cal Bp (Watts 1980). Jackson Pond in central Kentucky also was covered by boreal forests during the terminal Pleistocene (Wilkins et al. 1991). Prior to 16,800 cal Bp, the closed overstory was dominated by spruce and jack pine, followed by open boreal woodland dominated by spruce and sedges from approximately 16,800 to 11,300 cal Bp. This was slowly replaced by mesic deciduous forest by 10,000–7300 cal Bp. The Fulton section of the Mississippi River in western Tennessee contained a multiproxy

Biotic Records Paleobotanical Record

Much of the paleobotanical research in the Southeast was completed by Paul and Hazel Delcourt (Delcourt 1979, 2002; Delcourt 1980; Delcourt and Delcourt 1985, 1991; Delcourt and Delcourt 1998, 2004; Delcourt et al. 1981; Delcourt et al. 1983), focusing on reconstructing major forest b ­ iomes throughout the past 20,000 years. During the LGM, boreal forests extended as far south as 34°N. By 16,500 cal Bp, these were replaced by mixed conifer-deciduous cool temperate forests between 34°N and 37°N, with spruce and jack pine disappearing in these latitudes by approximately 12,500 cal Bp (Delcourt 1979; Delcourt et al. 1980). Three different bog sites in West Virginia and Maryland may have been covered by tundra from the LGM until approximately 13,000 cal Bp; this was followed by a boreal forest biome until approximately 10,500 cal Bp, which was succeeded by mixed deciduous forests (Delcourt and Delcourt 1998). Possibly tundras retreated much sooner in most areas, however, as Williams and colleagues (Williams 2003:14; Williams et al. 2004) recorded high values of tree pollen in the Southeast by 14,000 Bp. Locations in Virginia and Tennessee were boreal forests from roughly 20,000 cal Bp to approximately 12,500 cal Bp, never ­experiencing tundra. Moving south, Georgia, Florida, and Louisiana probably never had boreal forests or tundra in the late Quaternary. They were always covered with a mosaic of mixed deciduous forests and more open park- and scrublands (Delcourt 2002; Delcourt and Delcourt 1998). Recent research with more precise dating and more complete plant community data is showing that biome associations during the late Pleistocene were complicated. Each species responded 63

Figure 3.3. Reconstructed biomes for the Southeast from 21,000 BP to 10,000 BP (Simplified from Leduc 2003; Williams et al. 2004.)

Reconstructing the Pleistocene Environment of the Greater Southeast

Figure 3.4. Modern biomes for the southeastern United States. (Associations from Leduc 2003; Williams et al. 2004.)

­ aleoenvironmental record indicating ­boreal p forest conditions around a slack-water lake prior to 19,300 cal Bp, followed by lake infilling by 19,000 cal Bp (Grimley et al. 2009). At Nonconnah Creek in southwestern Tennessee, ­mesic deciduous forest survived throughout the late Pleistocene due to climate amelioration caused by local bluff topography, while much of the surrounding territory was boreal forest (Delcourt et al. 1980). Goshen Springs in southeastern Alabama contains a nearly continuous sediment sequence for the late Quaternary. Paul Delcourt (1980) demonstrated that from approximately 26,000 cal Bp to approximately 5000 cal Bp the area was warm with dry summers, leading to an oak–hickory– southern pine biome, with no evidence of boreal forest or open prairie. Seasonal droughts, however, were common. Sandy Run Creek, in central

Georgia, was covered by a grassland with pine and spruce stands prior to 25,000 cal Bp (La­Moreaux et al. 2009). Sediments are absent for the period 25,000–13,000 cal Bp, but terminal Pleistocene pollen (13,000–11,000 cal Bp) indicates cool, moist conditions with common oak and less pine. During the early Holocene, pine was even less common, and tupelo and oak pollen indicate a moist, warm climate. The Page-Ladson site in northern Florida shows a well-established mesic hardwood f­ orest between approximately 15,500 and 13,300 cal Bp (Han­sen 2006). The Lake Annie sequence in south-central Florida was dominated by rosemary, indicative of drier climate, prior to 13,000 cal Bp (Watts 1975). Sequences from several other lakes in Florida show that the Floridian peninsula has biotic responses linked closely in time to northern Heinrich events, with peaks in pine 65

Halligan

pollen occurring during times of ice advance and moister climates, and a general trend toward oak and hickory during drier ice retreat events (Grimm et al. 2006; Watts and Hansen 1994). Recent plant genetics studies (Gonzales et al. 2008; Gonzales et al. 2009) indicate that there may have been many more pockets of mesic deciduous forest throughout the late Pleistocene Southeast than originally thought. Jackson and Overpeck (2000) further point out that climate constantly changes, and plants respond to this change as individuals, meaning that plant communities are also constantly changing. This means that it is difficult to make meaningful paleobotanical reconstructions, but these reconstructions are a necessary starting place for discussing human activities in the terminal Pleistocene and early Holocene.

topic analyses that help explicate chronologies, animal behaviors, and local environments (Fisher and Fox 2006; Hoppe and Koch 2007; McDonald and Bryson 2010; Semken et al. 2010). Because the Rancholabrean record is so extensive in the Southeast, it is occasionally possible to track environmental change by marking the local disappearance of certain small species (Mihlbachler et al. 2002). Unfortunately, the extinction dates of larger fauna are still heavily debated (Faith and Surovell 2009; Fiedel 2008; Grayson 2007; Haynes 2008:2; Scott 2010). In general, the fauna of the Southeast indicate no-analogue environments during the terminal Pleistocene. While some plant communities may have been similar to those found in modern environments, the faunal communities at the end of the Pleistocene were remarkably unlike modern assemblages, with greater species richness than can be found anywhere today (Morgan and Emslie 2010; Semken et al. 2010). For instance, at times, the Florida peninsula maintained populations of animals now found only in western arid environments associated with animals now found only in tropical habitats and also associated with species still found in Florida. Morgan and Emslie (2010) interpret these associations, which can be found as far back as the late Pliocene, as coinciding with glacial intervals when Florida climates were drier with milder winters; higher sea levels and cooler winters probably marked interglacials, restricting the presence of tropical and western fauna in the Florida record. Several sinkholes in the Lower Aucilla River, including the archaeological sites Page-Ladson and Sloth Hole, contain numerous fauna from approximately 18,000–11,000 cal Bp that include California condors, porcupines (Erethizon dorsatum) only found today in western and northern locales, margays, and several extinct species: giant land tortoises, pampatheres, glyptodonts, the bear Tremarctos floridanus, the capybara Hydrochoerus holmesi, and the tapir Tapiris veroensis (Webb and Simons 2006). The uniqueness of late Pleistocene fauna holds true even when excluding extinct species and focusing upon small mammals whose modern environmental preferences are well known (Semken et al. 2010 and references within). Semken and colleagues (2010) have recently dated a number

Paleofaunal Record

One of the most striking aspects of the Pleistocene– Holocene transition in North America is the faunal extinction that occurred throughout the late Pleistocene and essentially ended by the early Holocene. During these several millennia, over 30 genera became extinct or died out locally in North America (Faith and Surovell 2009; Grayson 2007; Haynes 2008:2; Holliday and Meltzer 2010; Scott 2010). Many of these genera were large mammals such as sloths, camels, horses, and elephants and became extinct at approximately the same time as Clovis ended and the Younger Dryas occurred, leading to questions about the role humans had in these extinctions (Faith and Surovell 2009; Haynes 2008; Hemmings 2004; Martin 1984; Surovell and Waguespack 2008; Waguespack and Surovell 2003). Less discussed by archaeologists, many smaller mammals also became extinct locally (Mihlbachler et al. 2002; Semken et al. 2010). The Southeast has one of the most intact late Pleistocene (Rancholabrean) faunal records in the world, with well-preserved specimens known from many locations, especially Big Bone Lick, Kentucky, and rivers in Florida but also including numerous caves and rockshelters (Dunbar et al. 2006; Fisher and Fox 2006; Hedeen 2008; Hemmings 2004; Semken et al. 2010; Waters et al. 2009; Webb and Simons 2006; Webb et al. 1984). While many of these fossils were found in secondary context or in private collections, these bones sometimes can be dated or used in iso­ 66

Reconstructing the Pleistocene Environment of the Greater Southeast

of microfaunal cave assemblages from the greater Southeast in order to determine if the faunas from cave deposits were palimpsests or represented actual species richness. They discovered that a number of animals were contemporaneous, with radio­carbon dates overlapping at two standard deviations. All of these dated specimens were then categorized by modern environment and grouped by 500-radiocarbon-year increments, showing that boreal forest animals, desert animals, prairie animals, and plains animals could all be found in association in the Southeast prior to the Younger Dryas (Semken et al. 2010:252). They interpret this to be indicative of patchy environments during the late glacial period. Strontium isotope research by Hoppe and Koch (2007) indicating that late Pleistocene megafauna were migrating farther than their full-glacial counterparts may support this conclusion. In short, the animal world of the Paleoindians was more complex with greater species ­richness. Even as genera were becoming extinct, ­animal associations remained unlike modern ones. When trying to understand Paleoindian subsistence strategies, thus, it is important to consider both how these different associations may have changed animal behaviors and how people were able to respond to the changing faunal associations during the terminal Pleistocene.

ing people and animals to adjust to new plant associations. One particular type of climate change would have been perceivable by the individual, however. Not only were there approximately 25 m of water level rise during the 450 years of Clovis (Waters and Stafford 2007), the shoreline moved as much as 80 km landward in the Gulf of Mexico during the Clovis period alone (Figures 3.1 and 3.2). If Clovis people were exploiting the coasts, Gulf Coast groups would have seen the shoreline move inland approximately 100–200 m every year, even though actual sea level rise may have only been millimeters. This rise would have flooded estuaries and bays, drowning fragile coastal ecosystems and very likely preventing coastal biomes from re-forming due to the continued rapid submergence. If people were using coastal resources, those resources would have been flooded and destroyed, and coastline and landscape knowledge would have needed updating on a yearly basis, requiring a great deal of flexibility. Flexibility probably would have been the trait most advantageous for late Pleistocene people in the Southeast. Their landscape was in flux: Coastlines were moving, animals were dying, deserts were becoming marshes, and marshes were becoming lakes, only to become deserts again. ­Animals also may have been changing their migration patterns due to changing climate and floral patterns. Possibly the only observable aspect of the environment that remained unchanging over generations would have been the location of toolstone, and even many of these quarries became flooded over time (e.g., Hemmings 1999). On the other hand, much of the Pleistocene was marked by rapid change, so perhaps adaptation to fluctuation was an important part of Paleoindian lifeways in the Southeast. This will become clearer as we learn more about both Paleoindians and their environments. Unfortunately, the current data sets are somewhat spotty for Southeastern paleoenvironments and have numerous problems with scale and resolution. While these problems must be acknowledged, the transition from a glacial world to our modern interglacial planet is becoming better understood almost daily. As more data are collected, our knowledge of Paleoindian environmental constraints and cultural choices will also improve.

Southeastern Environments During the Terminal Pleistocene and Human Implications

The late Pleistocene and early Holocene of the Southeast were marked by relatively rapid change on a scale potentially observable by an i­ ndividual. Changing climate completely altered plant and animal communities and would have had profound impacts upon human lifeways. C ­ hanging fauna may have required adopting different subsistence strategies; changing flora would have required different landscape knowledge. It is impor­ tant to remember, however, that not all of these changes would have been observable by people. All of the animal species did not become extinct in a single generation, and it may or may not have been important to people that armadillos and northern wood rats stopped inhabiting the same areas. The change from a cool forest to a warmer one would have occurred over generations, allow67

Halligan

References Cited

Dahms, Dennis E. 1998 Reconstructing Paleoenvironments from ­Ancient Soils: A Critical Review. Quaternary International 51–52:58–60. Dahms, Dennis E., and Vance T. Holliday 1998 Soil Taxonomy and Paleoenvironmental Reconstruction: A Critical Commentary. Quaternary International 51–52:109–114. Dansgaard, W., S. Johnsen, H. Clausen, D. Dahl-­ Jensen, N. Gundestrup, C. Hammer, and H. Oeschger 1984 North Atlantic Climatic Oscillations Revealed by Deep Greenland Ice Cores. In Climate Processes and Climate Sensitivity, edited by J. E. Hansen and T. Takahashi, pp. 288–298. Geophysical Monographs Series, Vol. 29. American Geophysical Union, Washington, D.C. Delcourt, Hazel R. 1979 Late Quaternary Vegetation History of the Eastern Highland Rim and Adjacent Cumberland Plateau of Tennessee. Ecological Monographs 49(3):255–280. 2002 Forests in Peril: Tracking Deciduous Trees from Ice-Age Refuges into the Greenhouse World. McDonald and Woodward, Blacksburg. Delcourt, Hazel R., and Paul A. Delcourt 1985 Quaternary Palynology and Vegetational History of the Southeastern United States. In Pollen Records of Late-Quaternary North American Sediments, edited by Vaughn M. Bryant, Jr., and Richard G. Holloway, pp. 1–​ 37. American Association of Stratigraphic Palynologists Foundation, Dallas. 1991 Quaternary Ecology: A Paleoecological Perspective. Chapman and Hall, London. Delcourt, Hazel R., Paul A. Delcourt, and Elliott C. Spiker 1983 A 12,000-Year Record of Forest History from Cahaba Pond, St. Clair County, Alabama. Ecology 64(4):874–887. Delcourt, Hazel R., Darrell C. West, and Paul A. ­Delcourt 1981 Forests of the Southeastern United States: Quantitative Maps for Abovegound Woody Biomass, Carbon, and Dominance of Major Tree Taxa. Ecology 62(4):879–887. Delcourt, Paul A. 1980 Goshen Springs: Late Quaternary Vegetation Record for Southern Alabama. Ecology 61(2):​ 371–386. Delcourt, Paul A., and Hazel R. Delcourt 1998 Paleoecological Insights on ­Conservation of Biodiversity: A Focus on Species, ­Ecosystems, and Landscapes. Ecological Applications 8(4):​​ 921–934. 2004 Prehistoric Native Americans and Ecological

Anderson, David G., and J. Christopher Gillam 2000 Paleoindian Colonization of the Americas: Implications from an Examination of Physiography, Demography, and Artifact Distribution. American Antiquity 65(1):43–66. Anderson, David G., Albert C. Goodyear, James Kennett, and Allen West 2011 Multiple Lines of Evidence for Possible Human Population Decline/Settlement Reorganization During the Early Younger Dryas. Quaternary International 242(2):570–583. Balsillie, James H., and Joseph F. Donoghue 2004 High Resolution Sea-Level History for the Gulf of Mexico Since the Last Glacial Maximum. Report of Investigations 103. Florida Geological Survey, Tallahassee. Bigg, Grant R., Richard C. Levine, and Clare L. Green 2011 Modelling Abrupt Glacial North Atlantic Freshening: Rates of Change and Their Implications for Heinrich Events. Global and Planetary Change 79(3–4):176–192. Blum, Michael D., Jonathan H. Tomkin, Anthony Purcell, and Robin R. Lancaster 2008 Ups and Downs of the Mississippi Delta. ­Geology 36(9):675–678. Bond, Gerard, Hartmut Heinrich, Wallace Broecker, Laurent Labeyrie, Jerry McManus, John Andrews, Sylvain Huon, Ruediger Jantschik, Silke Clasen, Christine Simet, Kathy Tedesco, Mieczyslawa Klas, Georges Bonani, and Susan Ivy 1992 Evidence for Massive Discharges of Icebergs into the North Atlantic Ocean During the Last Glacial Period. Nature 360(6401):245–249. Bradley, Bruce, and Dennis Stanford 2004 The North Atlantic Ice-Edge Corridor: A Possible Palaeolithic Route to the New World. World Archaeology 36(4):459–478. Bronk Ramsey, Christopher 2009 Bayesian Analysis of Radiocarbon Dates. ­Radiocarbon 5(1):337–360. 2010 OxCal Version 4.1.4. Vol. 2010. Oxford ­University, Oxford. Clark, Peter U., Richard B. Alley, Lloyd D. Keigwin, Joseph M. Licciardi, Sigfus J. Johnsen, and Huaxiao Wang 1996 Origin of the First Global Meltwater Pulse Following the Last Glacial Maximum. Paleo­ ceanography 11(5):563–577. Clauzet, Gabriel, Ilana Wainer, Alban Lazar, Esther Brady, and Bette Otto-Bliesner 2007 A Numerical Study of the South Atlantic Circulation at the Last Glacial Maximum. Palaeo­geography, Palaeoclimatology, Palaeoecology 253(3–4):509–528. 68

Reconstructing the Pleistocene Environment of the Greater Southeast Change: Human Ecosystems in Eastern North America Since the Pleistocene. Cambridge University Press, Cambridge. Delcourt, Paul A., Hazel R. Delcourt, Ronald C. Brister, and Laurence E. Lackey 1980 Quaternary Vegetation History of the Mississippi Embayment. Quaternary Research 13(1):​ 111–132. Dincauze, Dena Ferran 2000 Environmental Archaeology: Principles and Practice. Cambridge University Press, Cambridge. Dixon, E. James 1999 Bones, Boats, and Bison: Archaeology and the First Colonization of Western North America. University of New Mexico Press, Albuquerque. 2001 Human Colonization of the Americas: Timing, Technology, and Process. Quaternary Science Reviews 20:277–299. Donnelly, Jeffrey P., Neal W. Driscoll, Elazar ­Uchupi, Lloyd D. Keigwin, William C. Schwab, E. Robert Thieler, and Stephen A. Swift 2005 Catastrophic Meltwater Discharge Down the Hudson Valley: A Potential Trigger for the Intra-Allerød Cold Period. Geology 33(2):​ ­89–92. Donoghue, Joseph F. 2006 Geography and Geomorphology of the Aucilla River Region. In First Floridians and Last Mastodons: The Page-Ladson Site in the Aucilla River, edited by S. David Webb, pp. 31–49. Springer, Dordrecht. Driese, Steven G., Zheng-Hua Li, and Sally P. Horn 2005 Late Pleistocene and Holocene Climate and Geomorphic Histories as Interpreted from a 23,000 14c yr Bp. Paleosol and Floodplain Soils, Southeastern West Virginia, USA. Quaternary Research 63(2):136–149. Dunbar, James S. 2002 Chronostratigraphy and Paleoclimate of Late Pleistocene Florida and the Implications of Changing Paleoindian Land Use. Unpublished Master’s thesis, Department of Anthropology, Florida State University. 2006 Pleistocene–Early Holocene Climate Change: Chronostratigraphy and Geoclimate of the Southeast U.S. In First Floridians and Last Mastodons: The Page-Ladson Site in the Aucilla River, edited by S. David Webb, pp. 103–158. Springer, Dordrecht. Dunbar, James S., C. Andrew Hemmings, Pamela K. Vojnovski, S. David Webb, and William M. Stanton 2006 The Ryan/Harley Site 8JE1004: A Suwannee Point Site in the Wacissa River, North Florida. In Paleoamerican Origins: Beyond ­Clovis,

edited by Robson Bonnichsen, ­Bradley T. Lepper, Dennis Stanford, and Michael R. ­Waters, pp. 81–96. Texas A&M University Press, College Station. Ehlers, Jürgen, and Philip L. Gibbard 2004 Quaternary Glaciations: Extent and Chronology. Pt. 2: North America. Developments in Quaternary Science, 2. Elsevier, Amsterdam. Faith, J. Tyler, and Todd A. Surovell 2009 Synchronous Extinction of North ­America’s Pleistocene Mammals. Proceedings of the National Academy of Sciences 106(49):20641– 20645. Faught, Michael K. 2008 Archaeological Roots of Human Diversity in the New World: A Compilation of Accurate and Precise Radiocarbon Ages from Earliest Sites. American Antiquity 73(4):670–698. Faure, Hugues, Robert C. Walter, and Douglas R. Grant 2002 The Coastal Oasis: Ice Age Springs on Emerged Continental Shelves. Global and Planetary Change 33:47–56. Fiedel, Stuart J. 2000 The Peopling of the New World: Present ­Evidence, New Theories, and Future Directions. Journal of Archaeological Research 8(1):​ 39–103. 2008 Sudden Deaths: The Chronology of Terminal Pleistocene Megafaunal Extinction. In American Megafaunal Extinctions at the End of the Pleistocene, edited by Gary Haynes, pp. 21–37. Springer, Dordrecht. Firestone, R. B., A. West, J. P. Kennett, L. Becker, T. E. Bunch, Z. S. Revay, P. H. Schultz, T. Belgya, D. J. Kennett, J. M. Erlandson, O. J. Dickenson, A. C. Goodyear, R. S. Harris, G. A. Howard, J. B. Kloosterman, P. Lechler, P. A. Mayewski, J. Montgomery, R. Poreda, T. Darrah, S. S. Que Hee, A. R. Smith, A. Stich, W. Topping, J. H. Wittke, and W. S. Wolbach 2007 Evidence for an Extraterrestrial Impact 12,900 Years Ago That Contributed to the Megafaunal Extinctions and the Younger Dryas Cooling. Proceedings of the National Academy of Sciences 104(41):16016–16021. Fisher, Daniel C., and David L. Fox 2006 Five Years in the Life of an Aucilla River Mastodon. In First Floridians and Last Mastodons: The Page-Ladson Site in the Aucilla River, edited by S. David Webb, pp. 343–378. Springer, Dordrecht. Fisher, Timothy G. 2003 Chronology of Glacial Lake Agassiz Melt­ water Routed to the Gulf of Mexico. Quaternary Research 59:271–276. 69

Halligan Pleistocene, edited by Gary Haynes, pp. 1–20. Springer, Dordrecht. Hays, J. D., John Imbrie, and N. J. Shackleton 1976 Variations in the Earth’s Orbit: Pacemaker of the Ice Ages. Science 194(4270):1121–1132. Hedeen, Stanley 2008 Big Bone Lick: The Cradle of American Paleontology. University of Kentucky Press, Lexington. Heinrich, Hartmut 1988 Origin and Consequences of Cyclic Ice Rafting in the Northeast Atlantic Ocean During the Past 130,000 Years. Quaternary Research 29(2):142–152. Hemmings, C. Andrew 1999 Fossil Hole Site Update. Aucilla River Times, Vol. 12. Florida Museum of Natural History, Gainesville. 2004 The Organic Clovis: A Single ­Continent-​ Wide Cultural Adaptation. Unpublished Ph.D. dissertation, University of Florida. Holliday, Vance T., and David J. Meltzer 2010 The 12.9-ka ET Impact Hypothesis and North American Paleoindians. Current Anthropology 51(5):575–607. Hoppe, Kathryn A., and Paul L. Koch 2007 Reconstructing the Migration Patterns of Late Pleistocene Mammals from Northern Florida, USA. Quaternary Research 68(3):​ 347–352. Jackson, Stephen T., and Jonathan T. Overpeck 2000 Responses of Plant Populations and Communities to Environmental Changes of the Late Quaternary. Paleobiology 26(4):194–220. Johnsen, S. J., D. Dahl-Jensen, N. Gundestrup, J. P. Steffensen, H. B. Clausen, H. Miller, V. Masson-­ Delmotte, A. E. Sveinbjorndottir, and J. White 2001 Oxygen Isotope and Palaeotemperature Records from Six Greenland Ice-Core Stations: Camp Century, Dye-3, GRIP, GISP2, Renland and NorthGRIP. Journal of Quaternary Science 16(4):299–307. Jouzel, J., and EPICA Community Members 2004 EPICA Dome C Ice Cores Deuterium Data. Data Contribution Series #2004-038. NOAA/ NGDC Paleoclimatology Program, Boulder. Kelly, Robert L. 2003 Maybe We Do Know When People First Came to North America; and What Does It Mean If We Do? Quaternary International 109–110:133–145. LaMoreaux, Heidi K., George A. Brook, and John A. Knox 2009 Late Pleistocene and Holocene E ­ nvironments of the Southeastern United States from the

Gonzales, Eva, James L. Hamrick, and Chang ­Shu-Mei 2008 Identification of Glacial Refugia in South-​ Eastern North America by Phylogeographical Analyses of a Forest Understorey Plant, Trillium cuneatum. Journal of Biogeography 35(5):844–852. Gonzales, Leila M., Eric C. Grimm, John W. Williams, and Erik V. Nordheim 2009 A Modern Plant-Climate Research Dataset for Modeling Eastern North American Plant Taxa. Grana 48(1):1–18. Goodyear, Albert C. 2005 Evidence for Pre-Clovis Sites in the Eastern United States. In Paleoamerican Origins: Beyond Clovis, edited by Robson Bonnichsen, Bradley T. Lepper, Dennis Stanford, and Michael R. Waters, pp. 103–112. Center for the Study of the First Americans, Texas A&M University Press, College Station. Grayson, Donald K. 2007 Deciphering North American Pleistocene Extinctions. Journal of Anthropological Research 63:185–214. Grimley, David A., Daniel Larsen, Samantha W. ­Kaplan, Catherine H. Yansa, B. Brandon Curry, and Eric A. Oches 2009 A Multi-Proxy Palaeoecological and Palaeo­climatic Record Within Full Glacial ­Lacustrine Deposits, Western Tennessee, USA. Journal of Quaternary Science 24(8):​ 960–981. Grimm, Eric C., William A. Watts, George L. Jacobson, Jr., Barbara C. S. Hansen, Heather R. Almquist, and Ann C. Dieffenbacher-Krall 2006 Evidence for Warm Wet Heinrich Events in Florida. Quaternary Science Reviews 25(17– 18):2197–2211. Hansen, Barbara C. S. 2006 Setting the Stage: Fossil Pollen, Stomata, and Charcoal. In First Floridians and Last Mastodons: The Page-Ladson Site in the Aucilla River, edited by S. David Webb, pp. 159–181. Springer, Dordrecht. Haynes, C. Vance, Jr. 2005 Clovis, Pre-Clovis, Climate Change and Extinction. In Paleoamerican Origins: Beyond Clovis, edited by Robson Bonnichsen, Bradley T. Lepper, Dennis Stanford, and Michael R. Waters, pp. 113–132. Center for the Study of the First Americans, Texas A&M University Press, College Station. Haynes, Gary 2008 Introduction to the Volume. In American Megafaunal Extinctions at the End of the 70

Reconstructing the Pleistocene Environment of the Greater Southeast Muridae, Rodentia) from Northern Florida. Quaternary Research 58(3):289–295. Morgan, Gary S., and Steven D. Emslie 2010 Tropical and Western Influences in Vertebrate Faunas from the Pliocene and Pleistocene of Florida. Quaternary International 217(1–2):143–158. Newby, Paige, James Bradley, Arthur Spiess, Bryan Shuman, and Phillip Leduc 2005 A Paleoindian Response to Younger Dryas Climate Change. Quaternary Science Reviews 24(1–2):141–154. Newsom, Lee A. 2006 Paleoenvironmental Aspects of the Macrophytic Plant Assemblage from Page-Ladson. In First Floridians and Last Mastodons: The Page-Ladson Site in the Aucilla River, edited by S. David Webb, pp. 181–215. Springer, ­Dordrecht. NGRIP Members 2004 High Resolution Climate Record of the Northern Hemisphere Back into the Last Inter­glacial Period. Nature 431:147–151. Peltier, W. R., and R. G. Fairbanks 2006 Global Glacial Ice Volume and Last Glacial Maximum Duration from an Extended Barbados Sea Level Record. Quaternary Science Reviews 25:3322–3337. Petit, J. R., J. Jouzel, D. Raynaud, N. I. Barkov, J. M. Barnola, I. Basile, M. Bender, J. Chappellaz, J. D ­ avis, G. Delaygue, M. Delmotte, V. M. Kotlyakov, M. Legrand, V. Lipenkov, C. Lorius, L. Pépin, C. Ritz, E. Saltz­ man, and M. Stievenard 1999 Climate and Atmospheric History of the Past 420,000 Years from the Vostok Ice Core, Antarctica. Nature 399:429–436. Rittenour, Tammy M., Michael D. Blum, and Ronald J. Goble 2007 Fluvial Evolution of the Lower Mississippi River Valley During the Last 100 k.y. Glacial Cycle: Response to Glaciation and Sea-Level Change. Geological Society of America Bulletin 119(5–6):586–608. Scott, Eric 2010 Extinctions, Scenarios, and Assumptions: Changes in Latest Pleistocene Large Herbivore Abundance and Distribution in Western North America. Quaternary International 217(1–2):225–239. Semken, Holmes A., Jr., Russell W. Graham, and Thomas W. Stafford, Jr. 2010 AMS 14C Analysis of Late Pleistocene Non-​ Analog Faunal Components from 21 Cave Deposits in Southeastern North America. Quaternary International 217(1–2):240–255.

Stratigraphy and Pollen Content of a Peat Deposit on the Georgia Coastal Plain. Palaeogeography, Palaeoclimatology, Palaeoecology 280(3–4):300–312. Leduc, Phillip 2003 Pollen Viewer 3.2. Vol. 2010. National Climatic Data Center, Asheville. Leigh, David S. 2008 Late Quaternary Climates and River Channels of the Atlantic Coastal Plain, Southeastern USA. Geomorphology 101:90–108. Leigh, David S., Pradeep Srivastava, and George A. Brook 2004 Late Pleistocene Braided Rivers of the Atlantic Coastal Plain, USA. Quaternary Science Reviews 23:65–84. Lowery, Darrin L., Michael A. O’Neal, John S. Wah, Daniel P. Wagner, and Dennis J. Stanford 2010 Late Pleistocene Upland Stratigraphy of the Western Delmarva Peninsula, USA. ­Quaternary Science Reviews 29(11–12):1472– 1480. Mandryk, Carole A. S., Heiner Josenhans, Daryl W. Fedje, and Rolf W. Mathewes 2001 Late Quaternary Paleoenvironments of Northwestern North America: Implications for Inland Versus Coastal Migration Routes. Quaternary Science Reviews 20:301–314. Martin, Paul S. 1984 Prehistoric Overkill: The Global Model. In Quaternary Extinctions, edited by Paul S. Martin and R. G. Klein, pp. 354–403. University of Arizona Press, Tucson. McDonald, H. Gregory, and Reid A. Bryson 2010 Modeling Pleistocene Local Climatic Parameters Using Macrophysical Climate Modeling and the Paleoecology of Pleistocene Megafauna. Quaternary International 217(1–2):​ 1­ 31–137. McKay, R. M., G. B. Dunbar, T. R. Naish, P. J. Barrett, L. Carter, and M. Harper 2008 Retreat History of the Ross Ice Sheet (Shelf) Since the Last Glacial Maximum from Deep-Basin Sediment Cores Around Ross I­ sland. Palaeogeography, Palaeoclimatology, Palaeoecology 260(1–2):245–261. Meltzer, David, and Vance Holliday 2010 Would North American Paleoindians Have Noticed Younger Dryas Age Climate Changes? Journal of World Prehistory 23(1):​ 1–41. Mihlbachler, M. C., C. Andrew Hemmings, and S. David Webb 2002 Morphological Chronoclines Among Late Pleistocene Muskrats (Ondatra zibethicus: 71

Halligan Webb, S. David 1998 Two Cycles of Late Pleistocene Sinkhole Filling in the Middle Aucilla River, Jefferson County, Florida. Paper presented at the Wakulla Springs Woodville Karst Plain Symposium, Wakulla Springs, Florida. Webb, S. David, and James S. Dunbar 2006 Carbon Dates. In First Floridians and Last Mastodons: The Page-Ladson Site in the Aucilla River, edited by S. David Webb, pp. 83–101. Springer, Dordrecht. Webb, S. David, Jerald T. Milanich, Roger Alexon, and James S. Dunbar 1984 A Bison antiquus Kill Site, Wacissa River, Jefferson County, Florida. American Antiquity 49(2):384–392. Webb, S. David, and E. Simons 2006 Vertebrate Paleontology. In First Floridians and Last Mastodons: The Page-Ladson Site in the Aucilla River, edited by S. David Webb, pp. 215–246. Springer, Dordrecht. Wilkins, Gary R., Paul A. Delcourt, Hazel R. Delcourt, Frederick W. Harrison, and Manson R. Turner 1991 Paleoecology of Central Kentucky Since the Last Glacial Maximum. Quaternary Research 36(2):224–239. Williams, John W. 2003 Variations in Tree Cover in North America Since the Last Glacial Maximum. Global and Planetary Change 35(1–2):1–23. Williams, John W., B. Shuman, P. J. Bartlein, N. S. Diffenbaugh, and T. Webb 2010 Rapid, Time-Transgressive, and Variable Responses to Early Holocene Midcontinental Drying in North America. Geology 38(2):135– 138. Williams, John W., B. Shuman, Thompson Webb, III, Patrick J. Bartlein, and Phillip Leduc 2004 Late Quaternary Vegetation Dynamics in North America: Scaling from Taxa to ­Biomes. Ecological Monographs 74(2):309–334. Williams, John W., Thompson Webb, III, Pierre H. Richard, and Paige Newby 2000 Late Quaternary Biomes of Canada and the Eastern United States. Journal of Biogeography 27(3):585–607. Yesner, David R. 2001 Human Dispersal into Interior Alaska: Antecedent Conditions, Mode of Colonization, and Adaptations. Quaternary Science Reviews 20:315–327.

Sherwood, Sarah C., Boyce N. Driskell, Asa R. Randall, and Scott C. Meeks 2004 Chronology and Stratigraphy at Dust Cave, Alabama. American Antiquity 69(3):533–554. Straus, Lawrence, and Ted Goebel 2011 Humans and Younger Dryas: Dead End, Short Detour, or Open Road to the Holocene? Quaternary International 242(2):259–261. Surovell, Todd A., Vance T. Holliday, Joseph A. M. Gingerich, Caroline Ketron, C. Vance Haynes, Ilene Hilman, Daniel P. Wagner, Eileen Johnson, and Philippe Claeys 2009 An Independent Evaluation of the Younger Dryas Extraterrestrial Impact Hypothesis. Proceedings of the National Academy of Sciences 106(43):18155–18158. Surovell, Todd A., and Nicole M. Waguespack 2008 Human Prey Choice in the Late ­Pleistocene and Its Relation to Megafaunal ­Extinctions. In American Megafaunal Extinctions at the End of the Pleistocene, edited by Gary Haynes, pp. 77–105. Springer, Dordrecht. Thulman, David K. 2009 Freshwater Availability as the Constraining Factor in the Middle Paleoindian Occupation of North-Central Florida. Geoarchaeology 24(3):243–276. Waguespack, Nicole M., and Todd A. Surovell 2003 Clovis Hunting Strategies, or How to Make Out on Plentiful Resources. American Antiquity 68(2):333–352. Waters, Michael R., and Thomas W. Stafford, Jr. 2007 Redefining the Age of Clovis: Implications for the Peopling of the Americas. Science 315:1122–1126. Waters, Michael R., Thomas W. Stafford, Jr., Brian G. Redmond, and Kenneth B. Tankersley 2009 Clovis and the American Mastodon at Big Bone Lick, Kentucky. American Antiquity 74(3):558–567. Watts, W. A. 1975 A Late Quaternary Record of Vegetation from Lake Annie, South-Central Florida. ­Geology 3(6):344–346. 1980 Late-Quaternary Vegetation History at White Pond on the Inner Coastal Plain of South Carolina. Quaternary Research 13(2):187–199. Watts, W. A., and B. C. Hansen 1994 Pre-Holocene and Holocene Pollen ­Records of Vegetation History from the Florida Peninsula and Their Climatic Implications. Palaeo­geography, Palaeoclimatology, Palaeoecology 109(2–4):163–176.

72

II

Reinvestigations of Classic Sites

4

A Report on the 2008 Field Investigations at the Shoop Site (36DA20) Kurt W. Carr, J. M. Adovasio, and Frank J. Vento

Located in central Pennsylvania, the Shoop Paleoindian site was first reported by John Witthoft in 1952. His research and several subsequent analyses (Carr 1989; Cox 1972, 1986; Krieger 1954; Wilmsen 1970) were based on an unmapped surface collection conducted over an 8-ha (20-ac) area. Witthoft identified 11 artifact concentrations within the site, although these were never mapped. Shoop is one of the largest Paleoindian sites in the east, and its size and artifact assemblage are presently unique in the Middle Atlantic region. There are, however, numerous questions concerning its function, its role within a settlement system, and its chronological placement within the Paleoindian period. With the assistance of a grant from the National Park Service (Challenge Cost Share Program), a field-testing program was conducted in 2008 to address some of these issues. The goals of this work were to systematically sample and map one of the artifact concentrations, search for subsurface features for possible radiometric dating and environmental data, conduct more detailed lithic sourcing ­studies, develop a comprehensive artifact catalog for the site, conduct microwear analysis on a sample of the tools, and develop an accurate map for the entire site. Ultimately, this led to a National Historic Landmark nomination. The following will summarize the previous research at the site, present the results of the 2008 fieldwork, and ­present a preliminary analysis of all artifacts currently available at the State Museum of Penn-

sylvania, at the Smithsonian Institution, and in two private collections. Previous Research

When John Witthoft reported on the Shoop site in 1952, it was only the third major publication on a Paleoindian site east of the Mississippi — ​Parrish (Webb 1951) and Williamson (McCary 1951) preceded it — ​and the first in the Northeast. His hypothesis concerning the nature of Paleoindian culture was controversial at the time, but some of Witthoft’s ideas are still part of our model for Paleoindian behavior. This research had a major impact on Paleoindian studies in the east, but it is ironic that this site was analyzed at such an early date. In the early 1950s, so little else was known about other contemporary sites that Witthoft’s analysis essentially existed in a vacuum for decades. There were few if any published sites similar to Shoop, and his intensive level of analysis was a landmark. In retrospect, it is difficult to compare this site with others in the Middle Atlantic region, and in many ways it does not fit our current model for Paleoindian settlement patterns. A blade tool–producing site in the uplands of the Ridge and Valley Province 350 km (200 mi) from its lithic source (Figure 4.1) is a very controversial proposal for a site in the Middle Atlantic region (see Carr and Adovasio 2002). The Shoop site has been surface collected by professionals but mostly avocational archaeologists since it was discovered in the 1930s. Based 75

Carr et al.

Figure 4.1. Map of the Middle Atlantic region illustrating the sites and quarries discussed in the text.

on several informant interviews, collectors have been picking up every visible artifact for the past 20 years or more. These collections typically include numerous pieces of microdebitage that can measure to less than 5 mm in size. There are four major collections known that probably include 80 percent or more of all known artifacts from the site. These are the State Museum of Pennsylvania (n = 4,171), the Arthur Shertzer collection (n = 2,063), the Smithsonian Institution (n = 321), and the Wise collection (n = 190). In addition, the Frank Soday collection at one time contained approximately 120 “carefully collected” artifacts gathered during the 1940s. This was analyzed by

Cox in 1972. The collection included both tools and debitage, and Cox (1986:102) felt that this was a representative sample of both. Unfortunately, the last known address for Soday was in Tulsa, Oklahoma, and the whereabouts of the collection are currently unknown. Shoop is a multicomponent site with bifurcate (Middle Archaic) through triangular (Late Woodland) projectile point types recovered from unprovenienced surface collecting. Based on our examination of local collections, there are approximately 100 non-Paleoindian artifacts, of which the majority are projectile points chipped from local lithic materials. No pottery has been 76

Report on the 2008 Field Investigations at the Shoop Site

recovered from the site, and fire-cracked rock is rare or nonexistent. Over 97 percent of the assemblage has been identified as Onondaga chert, and 90 percent of the fluted points are in the same material. Six of the fluted points (four jasper and two non-Onondaga chert) are not manufactured from Onondaga chert. The remainder of the collection consists of patinated jasper, chert, and chalcedony. For this analysis, only Onondaga and heavily patinated, presumably older, material was s­ tudied.

He argued that because the first inhabitants of the Shoop site were new arrivals putatively migrating from the north, they were not ­familiar with the lithic sources of central Pennsylvania. Therefore, the more proximal jasper outcrops at Vera Cruz (­Lehigh County) and Houserville (Centre County) were not extensively utilized. Witthoft compared the Shoop technology with the Denbigh Flint Complex and asserted that they were not the same but very similar. In comparison to other eastern Paleoindian sites, the Williamson site in Virginia and Coe’s early sites in the Carolina Piedmont, the Shoop site showed the greatest dependence on blades. He also believed that the jasper from Williamson originated in Pennsylvania, although, in reality, it was probably Flint Run jasper from the Shenandoah Valley. Relying upon (1) the presence of blades, (2) the exclusive use of a lithic source from the north, and (3) similarities with the Denbigh Flint Complex, he tentatively dated the site as one of the earliest in the east. Because the fluting technique was “crude” and there was a greater frequency of blade usage, he postulated that the Shoop site was earlier than the western Clovis tradition, that is, pre-Clovis in age. Although in 1952 there were very few radiocarbon dates, Witthoft estimated the age of the Shoop site between 18,000 and 8000 Bp. Dating has continued to be a problem with this site. Witthoft (1952) identified the Enterline Chert Industry as one of the earliest Paleoindian industries in the New World. The distinctive flaking techniques of blade production carried over into the use of guide flakes in the fluting of projectile points, involving a triple fluting process that is diagnostic of the Enterline Chert Industry. Similar principles were used in the final sharpening and the reworking of tools. The removal of flakes was not sequential

Overview of Witthoft’s Research

Shoop was discovered by amateur archaeologist George Gordon in the 1930s. Frank Soday and Gerald Fenstermaker were also early collectors on the site. While at the University of Pennsylvania, Edgar Howard (1942) conducted limited testing at the site, although the results and location of these excavations are unknown. Assisted by Sam Farver in the 1950s, John Witthoft conducted an intensive although unprovenienced surface collection of an area about 8 ha (20 ac) in size. Witthoft (1952) published a landmark article describing the site and analyzing a collection of approximately 2,000 artifacts, including more than 1,500 pieces of debitage, 400 tools, and 53 projectile points. Although many of his ideas have been revised, they played a significant role in interpretations of the Paleoindian period for more than 50 years. Witthoft defined the lithic technology used at Shoop as the Enterline Chert Industry. Enterline is the name of a local hamlet. One of his most controversial conclusions was that this assemblage consisted of blade tools made from blade cores in the tradition of the Old World Upper Paleo­lithic. The bases for this identification were the relatively narrow dimensions of the flakes and the fact that the dorsal flake scars were frequently aligned ­parallel rather than perpendicular to the long axis of the flake. Barry Kent et al. (1971:8) noted that J. L. Giddings (1951) had just published a report on the Denbigh Flint Complex, and he argued that the presence of blades and the absence of microliths (an alleged Mesolithic influence) suggested a very early date. Witthoft used the same arguments but also emphasized that all the lithic material was derived from the Onondaga chert outcrops of western New York, approximately 350 km (200 mi) north of the site (Figure 4.1).

nor were alternate chips drawn from opposite faces of a bifacial tool. Instead two chips were removed along an edge, with a space of an eighth to a third of an inch between the striking platforms and then a chip drawn between them removing part of the two earlier flake scars [Witthoft 1952:32]. However, in a subsequent analysis, Cox (1986:102) disagrees with this description and describes the sequential removal of flakes. 77

Carr et al.

Figure 4.2. Physiographic zones of Pennsylvania, highlighting the Shoop site and the Houserville and Vera Cruz jasper quarries.

The proposed function and location of this site set the model for Paleoindian site distributions in the east. The site was noted as a most unlikely place for an Indian camp, but it did offer a commanding view of Armstrong Valley, implying that it was a lookout station for hunting. Witthoft (1952:493) noted that to the east of the site there is a cul-de-sac area that might have been suitable for trapping big game animals. For the following two decades, until Gardner’s (1977) publication on the Flint Run Complex, archaeologists in the east searched for Paleoindian sites in hill/ridgetop settings. Based on a surface survey, and assisted by Sam Farver, Witthoft (1952:467) identified 11 artifact concentrations scattered over 20 ac. He characterized these as “slightly elevated areas . . . often more than a hundred yards apart and generally less than thirty feet in diameter” (1952:467). Unfortunately, none of these concentrations were mapped. Since Witthoft’s work, Wilmsen (1970), Cox (1972), Fogelman (1986), and Carr (1989) have conducted major analyses using the Shoop material. The reanalysis suggests that this is not a

blade core technology but, rather, the inhabitants of Shoop used a combination of bifacial and polyhedral cores. These authors also concluded that Witthoft’s Enterline Chert Tradition is not distinctive from other fluted point production strategies and this site is not pre-Clovis in age but more or less contemporary with other eastern Clovis sites. These studies will be reviewed below and incorporated into a comprehensive analysis of artifact function and the site’s placement within a regional settlement system. Environment Setting

The Shoop site (36DA20) is situated in the Ridge and Valley Physiographic Province, Susquehanna Lowlands Section (Figure 4.2). It is positioned between the fourth and fifth ridge (Peters Mountain and Berry Mountain) and 16 km from the southern margin of this section. It is located 24.5 km (15.2 mi) north of Harrisburg, Pennsylvania. It is 9.8 km (6.1 mi) east of the Susquehanna River in what could be characterized as a small valley on top of Dividing Ridge (Figure 4.3). This ridge is situated between two third-order streams. P ­ owell Creek lies 2.1 km (1.3 mi) to the south. The site 78

Report on the 2008 Field Investigations at the Shoop Site

Figure 4.3. (a) Close-up of the area shown in Figure 4.3b, illustrating the cross section between Third Mountain in the south and Berry Mountain in the north.

is drained by Armstrong Creek, which lies to the north. Several small springs exist along the southeastern edge. These first flow into an artificial pond and eventually create a first-order stream, Conley’s Creek East (Figure 4.4). Another first-order stream, Conley’s Creek, lies at the base of a very steep slope 536 m (.3 mi) north of the site, and it merges with Armstrong Creek 1,421 m (.9 mi) to the north, where it becomes a third-order stream. The soils are members of the Calvin–Leck Kill–Klinesville association (U.S. Department of Agriculture 1972). They are characterized as being situated in upland areas between the mountains. This association makes up about 30 percent of Dauphin County. The soils are shaley silt loams, overlying gray sandstone or acid red shale (U.S. Department of Agriculture 1972). The bedrock is Devonian in age and consists of the Duncannon Member of the Catskill Formation. It is bordered by the Clarks Ferry Member and the Irish Valley Member, all of which are siltstones, mud-

stones, sandstones, and conglomerates. The soils are very shallow, and bedrock frequently lies immediately below the plow zone. In addition, bedrock is exposed on the surface in several locations. As might be expected, the soils are rocky and well drained. During archaeological testing, using ¼-inch mesh, approximately 30 percent of the material would not pass through the screen. Based on limited testing across the entire site and intensive testing in one corner of the site, all artifacts are found in a plow-zone context or in the very top of the B-horizon. It is assumed that all of these have been disturbed by historic agricultural activity. However, there is at least one section of the site that has the potential for undisturbed cultural material. We had an opportunity to auger test the swale situated between two concentrations (the Lesher West and the Lesher North concentrations — ​see Figure 4.5). Soils in this swale were a meter deep and consisted of a thick plow zone to 30 cm and a colluvial B-​­horizon extending to a depth of 90 cm. Bedrock was encountered 79

Figure 4.3. (b) The Shoop site situated in the Ridge and Valley physiographic zone. The line marks the geographic cross section of the dividing ridge on which the Shoop site is located.

Report on the 2008 Field Investigations at the Shoop Site

Figure 4.4. Topographic map illustrating streams, cul-de-sac, and Steigman’s Bog.

at 110 cm. The B-­horizon from 40 to 90 cm was highly developed with clay skins and may well be late Pleistocene in age. Surrounded on three sides by artifact concentrations, this area may contain buried Paleoindian material. However, we were not able to conduct further testing as the field was under active cultivation. Based on our additional field survey, the Shoop site has now been determined to cover at least 15.2 ha (37.4 ac). Situated along rolling terrain on a north-facing slope, elevations range from 253 m (760 ft) to 276 m (820 ft) above sea level. In evaluating Witthoft’s “overlook” description and assuming there were no obstructions such as tall trees, the maximum view to the north when looking through the narrow stream gorge is 4.7 km (3.2 mi). Views in the other directions are 700 m to the west, 250 m to the south, and 1.9 to 2.9 km to the east (i.e., the cul-de-sac). The site is exposed to the elements and is protected only from southerly winds. The site has been farmed for over 200 years and was first occupied by a family named Shoop in the eighteenth century (Paul Shoop, personal communication 2008). Currently, four houses, two barns, and several outbuildings occupy the property. A paved road crosses from east to west along the northern part of the site, and an unimproved farm road traverses the site from north to south. A small woodlot separates the southern artifact concentration (Shertzer South) from the rest of the site, but otherwise the site is actively farmed. The woodlot measures approximately

Figure 4.5. Idealized stratigraphic profile of ­Steigman’s Bog.

2 ha and has not been plowed in the twentieth century. It offers one of the best areas for relatively undisturbed cultural deposits. The original Witthoft/Farver collection is curated at the State Museum of Pennsylvania. Additional materials have been added to the State Museum holdings, including artifacts collected in the first systematic testing of the site in 2008. The collection currently contains 4,171 tools, utilized flakes, and debitage (Table 4.1). The Smithsonian Institution holds the Gordon collection, and this contains 320 tools, utilized flakes, and a low frequency of debitage. The Wise collection was personally recovered by Jan and Warren Wise, who still retain the collection. This assemblage contains very little debitage, and it is assumed that only tools were picked up. Other than the 2008 State Museum material, these collections cannot be located to any specific concentration. The most productive fields are at the north end of the site on the Lesher property and formerly the Shertzer 81

Carr et al. Table 4.1. Tool Categories from the Shoop Site by Collection.

Type

Debitage Cores Utilized flakes Retouched flakes Sidescrapers Endscrapers Other scrapers Bifaces Wedges Gravers and concavities Fluted points Total

State Museum of Pennsylvania Collection

3,008 7 258 185 167 244 63 110 77 17 35 4,171

Shertzer Smithsonian Wise Collection Collection Collection

Total

1,663 8 58 91 43 62 17 45 52 24 2 2,065

4,700 20 332 335 261 514 106 214 143 43 77 6,745

property, now the Wise property. It is assumed that the State Museum and Smithsonian collections mostly came from the favored northern areas but could include material from all concentrations. Since the late 1980s, Mr. Shertzer, one of the property owners, has been collecting from a previously unknown concentration at the south end of the site. He does not allow anyone else to collect from this site, and he has kept this material separate from artifacts he has collected from other areas of the site. His collection from a single concentration consists of 2,063 tools, utilized flakes, and debitage.

26 1 6 26 38 130 6 48 4 0 34 319

3 4 10 33 13 78 20 11 10 2 6 190

125 km (70 mi) of the Shoop site, but by 14,500 14c yr Bp, they had moved into central New York (Ridge 2003:30). With the warm Bølling climatic episode, their retreat accelerated, and by 12,600 14c yr Bp, they were across the St. Lawrence River (Sirkin 1977:212). The Onondaga chert quarries of western New York would have been accessible by this time, which is probably at least 2,000 years before Shoop was occupied. Tundra probably would have existed directly in front of the glaciers but also would have quickly moved northward with the retreating front. Based on his analysis of Longswamp and Criders Pond, Guilday (1984:255) suggests that the tundra was less than 100 km wide. The term tundra, especially in the Middle Atlantic region, would characterize a low-latitude tundra (receiving much more sunlight than arctic tundra) and would only be superficially similar to the arctic tundra of the present. The Bølling is followed by the Older Dryas cooling event, which slows the retreat of the glaciers and the amelioration of the environment. By 12,000 14c yr Bp, this episode ends, and the Allerød warming begins. Glaciers retreat, and cold-adapted flora and fauna migrate north. Toward the end of this episode, at 11,000 14c yr Bp, humans were definitely occupying the region. The Shawnee-­Minisink site, with two Clovis points and dates of almost 11,000 14c yr Bp, may well be an example of Clovis in the Northeast during the Allerød climatic episode (see Gingerich, Chap-

Paleoenvironmental Reconstruction

Much of Carbone’s (1976) data on late glacial and early Holocene environments is derived from sites in Pennsylvania and, therefore, is applicable to this region. Davis (1983), Lundelius et al. (1983), Sirkin (1977), and Watts (1979) have also contributed to the understanding of past environments in this region. The most comprehensive work can be found in Vento et al. 1994, and their work will also be incorporated below. It should be noted at this juncture that all dates used here are presented as uncorrected radiocarbon years Bp. At the Last Glacial Maximum, 18,000 14c yr Bp, glaciers covered the northern part of Pennsylvania in an arc stretching from the Delaware Water Gap in the east to 15 km north of Williamsport in central Pennsylvania to 8 km south of New Castle at its western end. Glaciers existed within 82

Report on the 2008 Field Investigations at the Shoop Site

Figure 4.6. Location of the 2008 excavation within the site boundaries using the 2008 mapping.

ter 9, this volume). The Younger Dryas climatic episode begins at 10,950 14c yr Bp (Isarin and Bohncke 1999) and represents a return to a cold, glacial-like climate. This was the last Pleisto­cene cooling episode, and it stabilized the glacial retreat until approximately 10,150 14c yr Bp (Mayew­ ski et al. 1993). Contrary to previous cold events, this change occurred abruptly within 40–100 years. In addition, this episode was cold and dry rather than cold and wet as in previous episodes. Under these conditions, it is assumed that there was an overall reduction in food resources compared with previous times. The Younger Dryas in the Northeast created an unusual mosaic of ecological settings not found in the region today. Based on Carbone’s (1976) reconstruction for the Ridge and Valley provenience, this area of Pennsylvania would have been characterized by a combination of spruce, pine, oak, and nonarboreal species resulting in what is defined as an open boreal forest or parkland. The higher ridges to the east and south of the site were possibly spruce parkland, and the slopes were coniferous forest. The valley floors to the north and south were probably a combination of grassland

and mixed coniferous and deciduous forests. Situated on a low ridge, the site is at the edge of these two zonal types, although the relatively level topography and the well-drained soils suggest that the environment resembled the valley floors more than the ridge slopes. As noted by Cox (1972), the site itself must have been open grassland at the time of occupation, otherwise it would not have been selected for occupation. The area around the springs and along the low-order streams may have been occupied by an open deciduous woodland or grassland adjacent to stands of spruce and pine on the relatively shallow shaley soils. Witthoft (1952:493) felt that the Shoop site was a hunting camp and the focus of attention was a cul-de-sac situated approximately 1,170 m to the east of the site at the base of Broad M ­ ountain (see Figure 4.4). In 1988, a geomorphological investigation was conducted in a wet section of the culde-sac (identified as Steigman’s Bog). Three backhoe trenches were placed with the assistance of F. J. Vento, A. T. Bouldurian, and J. Herbstritt. The first meter of soils is a combination of colluvial and alluvial deposits (Figure 4.6). Below this is a very thin pebble layer, followed by a­ pproximately 83

Carr et al.

.5 m of alluvial soils resulting from vertical accretion. At an average depth of 1.7 m, discontinuous, light blue, gleyed clay was encountered, ­probably representing a marsh or bog. Immediately under this horizon, we encountered the present water table and a layer of pebbles of undetermined depth (although it is assumed that bedrock was not far below). Two humic acid dates were assayed from the gleyed clay layer, producing dates of 12,000 ± ​120 Bp (Pitt 273A) and 11,615 ± ​85 Bp (Pitt 273B). The samples represent the entire thickness of the gleyed horizon, and, thus, it is not possible to determine whether these dates are associated with the beginning or end of the deposit. Vento’s analysis (Carr 1989:11) of the environmental implications of this profile is summarized as follows. The heavy deposits at the bottom are the result of the high competence of Pleistocene Armstrong Creek when precipitation was high and temperatures were lower. There were decreased levels of vegetation and low rates of evapotranspiration. The gleyed level represents a late Pleistocene episode of slightly increased temperature and/or slightly decreased precipitation (i.e., the Allerød climatic episode). The broad stream was then replaced with a narrower channel during the Younger Dryas. Due to a raised water table, the floodplain was bog-like in ­nature. The topography and a minimal amount of ­sampling with an auger suggest that this feature was about 8 ha (20 ac) in size. Unfortunately this horizon does not contain any identifiable pollen. Unlike the case for many Paleoindian sites that are adjacent to these types of settings, this swamp is approximately 1,170 m (.073 mi) east of the site. Several authors have commented on the site’s unlikely location; however, with additional testing, other swampy areas could be discovered along this stream and closer to the site. During the time of the Paleoindian occupation, the sloped areas around the Shoop site were probably covered with spruce and pine; the more level areas, with grasses and deciduous elements such as oak; and the buried swamp to the east, with birch and other hydrophytic species. The exact dating of the occupation is important in reconstructing the vegetational patterns, as these would have changed between 11,100 and 10,200 14c yr Bp. During the Allerød episode (up to 10,950 14c yr Bp), the forest would have been more closed,

and deciduous species would have been more common along with temperate faunal species. With the onset of the Younger Dryas, the overall environment would be more open, with a higher frequency of grasses, nondeciduous species, and more cold-adapted fauna (Vento et al. 2008). The southeast corner of the site has a series of springs that would have been attractive to Paleoindians. These flow all or most of the year, and they are associated with the Shertzer South ­locus, where over 2,000 Paleoindian artifacts were recovered. Based on informant interviews, the north end of the site contains the highest concentration of artifacts. Although there is no water in this area of the site, there is, however, a good view of the valley for hundreds of meters in all directions (assuming it was not covered in large trees). The focus of this area of the site may have been the bog to the east and the low-order streams to the north and east. This appears to be a major “upland” Paleoindian residential hunting camp. If it were not for the intensity of the occupation, one could hypothesize that Paleoindians occasionally utilized upland environments. However, the Shoop site probably represents several recurrent visits, each consisting of a relatively large population. This is the largest known site of this age in the state, and it is unique in location and size. 2008 Field Project

The goals of the 2008 field investigation were to (1) systematically test and map one of the Shoop artifact concentrations, (2) search for subsurface features that may be suitable for ­radiometric dating, (3) conduct more detailed lithic s­ ourcing s­ tudies, (4) conduct microwear analysis on a sample of the tools, and (5) develop an ­accurate map for the entire site. Much of the site has been heavily collected, and there was a desire to choose a location that had escaped intensive looting. With the gracious assistance of the property owner, Mr. Ron Lesher, we identified a small, 1-ac field in the northern section of the site that had not been plowed since the early 1970s. The pristine condition of this area was supported by a­ erial photographs. In addition, according to Barry Kent, former state archaeologist, John Witthoft had previously identified this field to him as one of his “concentrations.” Bordered by a tree line to the west, a regularly cultivated field to the north, a 84

Report on the 2008 Field Investigations at the Shoop Site

paved road to the south, and Mr. Lesher’s vegetable garden to the east, it was designated Lesher West (Figure 4.5). It is situated approximately 120 m southwest of the largest and most productive concentration, designated Lesher North. The latter is separated by a swale from Lesher West. Across the tree line to the west is the Allman concentration, and this may have been part of the Lesher West concentration. The field was plowed, disked, and allowed to be rainwashed for over a month, during which it received about 4 cm of rain. Four controlled surface collections were conducted in May 2008. It rained again between the second and third collection. The first three collections consisted of ­individuals slowly walking, spaced at 3-m intervals, across the field. Multiple collections were conducted because artifact numbers and artifact sizes were very small on each transect. The final collection was conducted by surveyors literally crawling on the surface, elbow to elbow. A total of 60 prehistoric artifacts were recovered, and these were concentrated in the northern half of the field covering an area about 40 m in diameter. Based on the surface collection, shovel probes (50 cm in diameter) were placed at 5-m intervals. A total of 72 probes were excavated, and the fill was subsequently screened using ¼-inch screen followed by a ⅛-inch screen. These produced a total of 180 artifacts. The units contained between 0 and 20 artifacts and averaged 2.5 artifacts per probe. Each probe was taken to 20 cm below the topsoil except where precluded by bedrock. The shovel probes revealed areas of high artifact concentrations but also areas where bedrock was just below the plow zone and, in some cases, was probably being plowed. As noted above, locating Paleoindian features was a primary goal. Therefore, 1-m units were placed in areas associated with high artifact densities and where bedrock was not immediately below the surface. The fill was subsequently screened following a procedure similar to the shovel probes. In the time allowed, 42 units were excavated, including a contiguous block of 27 units. Six hundred forty-four prehistoric artifacts were recovered from these units, with a unit average of over 15 artifacts and a range from 0 to 47. We were able to trench 8 m to the west of the concentration and found a sharp decrease in arti-

fact density. Based on shovel probes, this concentration could have extended 6 m to the north, 6 m to the east, and possibly as much as 10 m to the south. This would produce a concentration that measures approximately 12 × 25 m. Several soil stains were identified. These contained charcoal and in some cases, what appeared to be thermally altered soil. However, all could be parsimoniously explained as being the result of tree roots or animal burrowing. The investigation of the Lesher West loci produced a total of 844 prehistoric artifacts. The vast majority of these were from the plow zone or the first 10 cm below the plow zone. Over 90 percent of these were less than 1 cm in length, and half were recovered from the ⅛-inch screen. The assemblage consists of 98 percent Onondaga chert and 98 percent are unmodified debitage. Only one unbroken tool was recovered, and the rest of the tools are fragmentary. The microdebitage appears to represent a combination of pressure flakes with both steep and acute striking platform angles and flake fragments. Based on other collections from the site, there should have been more large pieces from this location. The very high frequency of microdebitage in this area suggests that it has been heavily collected. Although the results of this fieldwork are somewhat disappointing, the distribution of the microdebitage is indicative of an occupation or an activity area approximately 12 × 25 m or 300 m2. This is within the size range reported by Witthoft (1952). Lithic Technology

The analysis of Paleoindian technology has been a major theme with the Shoop artifact assemblage. The presence of blades has been discussed by several researchers, and considering current discussions (Collins 1999; Dickens 2005; Goebel et al. 2008; Haag 2004; Haynes 2002; Steffy and Goodyear 2006), we will begin with that topic here. Except for the bifacial fluted points, Witthoft (1952) defined the Enterline Chert Industry as a blade industry. This has implications for the age of the site, its relationship with Old World technologies, and the settlement system. Krieger (1954) pointed out that there are also other bifacial artifacts, and all the collections from this site contain a variety of broken late-stage bifaces. Because some of these are broad or irregular in outline, 85

Carr et al.

but well thinned, they all do not appear to be projectile point preforms. Whatever their function, they represent 10.6 percent of all worked artifacts. Projectile points represent an additional 3.8 percent of all tools and worked artifacts. According to Witthoft, the remaining 85.6 percent are blade tools or debitage from blade cores. Although blade tools and a blade tool technology are found on western Paleoindian sites, eastern Paleoindian sites are generally characterized by a flake tool tradition (Collins 1999). Blades may be present on eastern sites, but they do not dominate the tool assemblages and never to the extent as found on European Upper Paleolithic sites. The exceptions to this are generally in regions such as the Upper Mississippi Valley that are rich in high-quality, homogeneous cherts available in a large package size (Carr et al. 2010; Haag 2004). According to Bordaz (1970), blade core techniques increase the possible edge area per unit of core material fivefold over Mousterian flake core techniques. Callahan (1979) has demonstrated that blade cores require more time to produce; however, once the core has been established, there is very little waste of raw material. This has obvious advantages for a nomadic hunting and gathering society that is generally quite selective in choosing lithic material. However, without belaboring the topic, there is a growing body of literature that suggests that blade tool technology is not necessarily more economical or efficient than flake tools (Carr et al. 2010; see also Jennings et al. 2010; Prasciunas 2007). We do not have the time to discuss this issue here. Krieger (1954) questioned the identification of blades at Shoop, and Wilmsen (1970) and Cox (1972) used analytical methods to reach the same conclusion. In contrast, Bouldurian (1985) has argued that blades were being produced on a regular basis at Shoop. There are several different definitions of the term blade, although they all use the two-to-one length/width ratio. Other variables include (1) the number of parallel flake scars on the dorsal surface, (2) a triangular or trapezoidal cross section, (3) the striking platform angle, and (4) the spine angle. Witthoft used the first two, and Wilmsen and Cox emphasized the latter two attributes. Subsequent work by Collins (1999) has emphasized that Clovis blades also have a pronounced curve.

Wilmsen (1970) compared eight Paleoindian sites with the Denbigh site material and found marked differences. He considered the Denbigh assemblage a true core and blade technology, and, contrary to Witthoft, none of the sites including Shoop has the same degree of blade utilization. Cox (1972) found that a relatively low striking platform angle and a triangular cross section characterized the Shoop material. He felt that these were not blades but simply large thick flakes struck from the ends of bifacial cores. There are no complete cores or exhausted cores from the site, although there are approximately 20 fragments. These do not always illustrate parallel flaking as would be expected from blade cores. Striking platform angles on flakes were analyzed in order to characterize the types of cores used at Shoop. In this analysis, it was found that 28 percent of the artifacts with striking platforms had low to moderate striking angles (25–59°). These platforms were frequently ground or pressure flaked, and they frequently retained parts of a biface edge. Using Callahan’s (1979) bifacial reduction sequence model, these traits suggest that the Shoop flakes resulted from a stage 3 or 4 bifacial core (Carr 1987). Flat (relatively unworked) striking platforms or platforms with a few large hinge fracture scars and steep striking platform angles (75–90°) made up 39 percent of the assemblage. It is believed that these flakes resulted from angular polyhedral cores. Thirty-three percent could not be confidently assigned to either group. Contrary to Cox’s conclusions, this suggests a slight preference for polyhedral cores over bifacial cores. As discussed by Callahan (1979), Carr (1992), and Verrey (1986), bifacial cores are the dominant type in the Flint Run Complex found at the Thunderbird site in Virginia. These are useful and efficient in that they serve as both a core and a tool. In a comparison of Flint Run cores with Shoop cores, Carr (1986) concluded that the differential core type at these two sites was partially related to the package size and the nature of the bedrock. The Flint Run Complex is associated with jasper outcrops that are massive in size, and there are no practical restrictions to the size and shape of cores. However, the Flint Run jasper is not homogeneous in quality and contains impurities and voids. Onondaga chert is tabular or nodular but found in layers no more than 13 cm (5 inches) 86

Report on the 2008 Field Investigations at the Shoop Site

and their triangular cross section was too difficult to work. Finally, as noted above, Collins (1999) described the western Clovis blades as having a pronounced curve. Shoop flakes are not generally curved and are flat. A distinguishing characteristic of the Enterline Chert Industry was the use of guide flakes in the fluting process. However, subsequent analyses by Callahan (1979), Cox (1972), and Gardner and Verrey (1979) all concluded that multiple fl ­ uting is common for fluted points in the Middle Atlantic region. In general, this characteristic is no more distinctive to the Shoop material than to any other Paleoindian assemblage. They also suggest that the so-called guide flutes just as frequently preceded the central flute as they came after the central flute. Callahan argues that the goal was thinning and as many flutes were used as necessary. In contrast to Callahan, and related to ­multiple fluting, D. Stanford (personal communication 2010) believes that small multiple flutes are not a characteristic of the western Clovis technology and represent a later fluted point technological trait. Investigators of the Shoop assemblage (Carr 1989; Cox 1972; Witthoft 1952) have observed that a relatively high percentage of the lithics have been sufficiently thermally altered to cause potlid fractures. In our analysis, 36 percent of all artifacts had evidence of pot-lid fractures. We believe that this has little to do with intentional heat treatment but relates to aboriginal fires or later natural fires. Based on experiments conducted by the senior author, once pot-lid fractures have occurred, the advantages of heat treatment are lost and further successful knapping will be difficult or impossible. Considering that the artifacts have been very close to the ground surface for probably 11,000–12,000 years, they must have been exposed to numerous fires. However, this is true of all sites in the region, and artifacts from other sites do not seem to exhibit as many pot-lid fractures on flakes as found in the Shoop collection. Although we have not examined other sites systematically, the frequency of this trait appears to be greater at Shoop. This same observation was made by Witthoft (1952). If this is true, it suggests that the heating relates to the Paleoindian occupation rather than later cultural or natural activities. In addition, and also observed by Cox (1986:108),

Figure 4.7. Schematic of an idealized Shoop site lithic core and the evolution of an endscraper.

thick (Wray 1948:41). It is also more homogeneous in quality than the Flint Run material. The relatively thin, blocky nature of the Onondaga Formation may be the reason for using angular polyhedral cores. The desired shape of the flakes (tool blanks) produced from cores is also influenced by the shape of the core. The most common prepared tools in the Shoop assemblage are endscrapers (n = 514). Cox (1972) characterizes these as being produced on thick flakes. We suspect that angular cores from tabular bedrock are a better source for this type of flake than bifacial cores (Figure 4.7). The Shoop cores were prepared in that they were shaped to produce a flake of predetermined dimensions; however, they were not blade cores. Along with preforms for endscrapers, many other flake types were produced from Shoop cores, and it would seem that the cores were modified to produce different types of tool blanks. According to Callahan (1979), polyhedral blade cores are associated with eastern Clovis; however, he believes that the vast majority of long narrow flakes result from bifacial cores. They may or may not be intentionally produced, but they are not regularly produced on blade cores, and the high-efficiency lithic reduction discussed by Bordaz is clearly not being achieved (or sought after for that matter) at Shoop. Although it is not necessarily applicable, Callahan (1979) found that blades were not the best blank for the production of fluted points. They were usually too narrow, 87

Carr et al.

the pot-lid fractures are patinated, suggesting that the thermal alteration dates to Paleoindian times. Another distinguishing characteristic of the Shoop assemblage is that it is dominated by a single lithic type. There are few if any other eastern Paleoindian sites of this size or artifact number that are characterized by a single lithic type so distant from its source. The stone for over 98 percent of the artifacts originates in the Divers Lake Onondaga chert quarries in western New York. The other lithic types mainly consist of jasper from Pennsylvania and a black chert, probably of local origin. One of the fluted points is bluish green in color and may originate in the Normanskill (Coxsackie) quarries of eastern New York. There have been questions as to Witthoft’s identification of the Onondaga Formation of western New York as the source of the Shoop lithic material. Holland and Dincauze (1999) and Moeller (1989) have argued on theoretical grounds that there must be a closer source. Topping (1995) has claimed that quantities of cobble-size Onondaga chert were available in the Lower Susquehanna Valley; however, these claims were investigated by the senior author and Dennis Stanford and were found to be not credible. There are no other published references identifying cobble sources for this material, and therefore we believe that this explanation can be ruled out. Quantifiable comparisons using trace element studies have yet to be conducted, but based on observable character­ istics, the outcrops at Divers Lake in western New York contain material that is essentially the same as the Shoop assemblage. Donald Hoff (senior geologist for the State Museum of P ­ ennsylvania in 1989) used the description developed by Lavin and Prothero (1981) and low magnification to compare Shoop artifacts with samples from ­Divers Lake. He concluded that the artifacts in the State Museum came from the chert outcrops at Divers Lake. More recently, Frank Vento compared hand samples of specimens from Divers Lake and the State Museum collection. In his opinion, the color, texture, and blue/tan banding are the same. He has also performed a thin section analysis using one artifact from the State Museum collection. This was compared with samples from Divers Lake. Both samples contained the same microfossils (i.e., sponge spicules) and limonite staining. They

also shared minute fractures that resemble desiccation cracks. Further, the size and frequency of dolomite and calcite rhombs are the same. Finally, both samples are fine grained and very homogeneous in character. Vento acknowledges that the Old Port/Onondaga Formation has been mapped within 40 km of the site; however, the chert in the local formation is black in color and contains high levels of detrital quartz not found in the Shoop artifacts. He feels strongly that the Shoop artifacts originate in the Divers Lake area, at a distance of over 350 km from the site. In further support of the long-distance movement of this chert, less than 4.0 percent of the assemblage retains cortex or a pebble rind. It is generally agreed that the frequency of cortex on flakes indicates the relative distance to the lithic source. The use of Onondaga chert throughout prehistory is common in the Susquehanna Valley, and much of it has been attributed to p ­ ebble sources. However, there is no other site in the valley where Onondaga chert makes up more than 20 percent of the lithic material, let alone 98 percent. It could be argued that the material was part of a trade and exchange system; however, this means that they only traded with groups near the Divers Lake area and the occupants of Shoop did not use any local material. In addition, the fact that all the material is heavily curated and the flakes are very small strongly suggests that the source is a great distance from the site. The overriding preference for cherts by Paleoindian people is clearly demonstrated at Shoop. Even though their tools were near exhaustion, they were not picking up any of the diabase, quartz, or quartzite that is common throughout the Susquehanna Valley. The fact that they did not use local cherts or jaspers is probably related to their unfamiliarity with the region and their short occupation. This suggests that Shoop could represent a colonizing group. In summary, the technological system observed at Shoop appears similar to other early Paleoindian components in the Middle A ­ tlantic region. There was an emphasis on high-quality cherts, and bifacial and polyhedral cores were used to produce flake blanks for tools. The Shoop assemblage is distinctive in that the tools are heavily reworked, the debitage is generally small, and tool maintenance, rather than tool produc88

Report on the 2008 Field Investigations at the Shoop Site

tion, was likely the most common lithic reduction activity. Finally, there are few eastern Paleoindian sites that illustrate the exclusive use (98 percent) of a single lithic type that has traveled 350 km from its source. Other sites such as Bull Brook in Massachusetts (Spiess et al. 1998:204) and Varney Farm in Maine (Peterson et al. 2002:128) are dominated by a single source that has traveled 200 to 300 km, but these situations are rare. Tool Analysis

Along with the focus on debitage and reduction strategies, several functional analyses were conducted on the formal tools and utilized flakes from the Shoop site. Witthoft (1952), Wilmsen (1970), and Cox (1972) have divided the tools into different morphological and/or functional classes. Witthoft separated the tools into six cate­gories, which included (1) endscrapers, (2) sidescrapers, (3) projectile points, (4) gravers, (5) knives and other bifacial tools, and (6) utilized flakes. He further subdivided the scraper group into triangular endscrapers, elongated endscrapers, and pointed sidescrapers. Cox recognized similar groupings but subdivided them to include (1) a greater variety of unifaces with lateral retouch, (2) drills, and (3) wedges (of which he identified 70). Cox (1972, 1986) presented both summaries and detailed metric data on all tools and projectile points. Wilmsen (1970) emphasized edge angels on tool bits and also developed metric data on the tool assemblage. The emphasis on tool morphology and edge angles is understandable for the 1970s. These tool types are useful for inter­site comparisons, and similar tool types were used in the analysis discussed below. However, artifact morphology is not directly related to tool function, and an intensive microwear analysis is planned for the future. For those interested in a metric analysis of tools, consult Cox 1986. Focusing on the edge angles of tool bits, Wilmsen (1970) performed a relatively i­ntensive analysis of the site’s function. He concluded that the range of artifacts was relatively small and that there was very little on-site tool p ­ roduction. Some broken bifaces may have been point blanks, though these were not produced on site but, rather, brought in nearly finished. Based on edge angles, he defines two categories of tools; one is represented by a small number of low-edge-­ 89

Figure 4.8. Shoop endscrapers from the State ­Museum collection.

angled endscrapers, and the other includes cutting tools that Wilmsen feels were used to process animal foods. The vast majorities of tools were steep-angled endscrapers, sidescrapers, and utilized flakes. Wilmsen believes that these were “indicative of plant processing operations” and “the manufacture of wooden implements” (1970:79). All researchers noted that the tools are heavily reworked and generally very small. Almost all flakes over 3 cm show evidence of use, and many of the smaller retouch flakes clearly exhibit parts of tool edges. The results of our preliminary analysis (of the four major collections) are presented in Table 4.2. In a total of 6,745 artifacts, 2,029 (30.1 percent) were prepared tools or utilized flakes. Endscrapers, as has been noted by all previous researchers, are the most common prepared tool category (Figure 4.8). They represent 25.4 percent of all tools and 7.6 percent of all artifacts. These are most frequently triangular to parallel sided in cross section and less frequently trapezoidal in cross section (Cox 1986:118). The bulb of percussion is usually reworked, obliterating the striking platform. Sometimes it appears that a burin blow has been used to remove an edge and develop a triangular shape. Of note, the

Carr et al.

In addition, there were black micro-inclusions that could be interpreted as charcoal. Striations were observed but were not extensive. There is a % of Total Type Count % of Tools Artifacts ­possibly that the scrapers were also used on dry hides, but “wood, and possibly charred wood, was Debitage 4,700 69.68 the dominate contact material” (Pope 2010:4). Cores 20 .30 Using the edge angles of tool bits, both Cox Utilized flakes 332 16.40 4.92 and Wilmsen felt that the working of wood, bone, Retouched flakes 335 16.54 4.97 and antler was a common activity at Shoop. It Sidescrapers 261 12.89 3.87 would seem that the microwear analysis generally Endscrapers 514 25.38 7.62 supports this hypothesis. Although more tools Other scrapers 106 5.23 1.57 need to be tested, endscrapers are the most comBifaces 214 10.57 3.17 mon tool from the site, which suggests that woodWedges 143 7.06 2.12 working, possibly the production of tool handles, Gravers and 43 2.12 .64 concavities spear shafts, or other composite tool parts, was a Fluted points 77 3.80 1.14 common activity at Shoop. Well-preserved haft-­ Total 6,745 100.00 100.00 related polishes were also observed by Pope on all specimens: “Many showed microscopic traces from having been inserted into a haft, including Shertzer collection contains a double-ended end- probable adhesives, striations and bright polish scraper (an endscraper bit on both the proximal caused by movement and friction” (2010:7). and distal end). These are found in Old World The 514 endscrapers are surprising for their Upper Paleolithic assemblages but are very rare consistent shape and small size. A sample of nearly in Paleoindian assemblages. 200 endscrapers from the site ­exhibits minimal In an effort to elucidate the activities that took deviation between artifact means: mass, 5.35 ± ​ place at Shoop, 16 endscrapers were submitted 3.10 g; length, 28.66 ± ​6.76 mm; width, 21.84 ± ​ to Melody Pope (2010) at the University of Iowa 3.70 mm; and thickness, 8.04 ± ​1.82 mm. The confor microwear analysis. The specimens were ex- sistent size of scrapers from the site could be the amined at magnifications between 50 and 400×. product of systematic blank production or speThey were examined in two stages. In the first, cialized manufacture for hafting. Cox and Witthey were all observed at high and low magnifi- thoft claimed that there was little or no evidence cation to characterize the nature of the use traces. for hafting, and especially Cox believed that they The second phase involved recording more de- were “hand held.” However, to use such small tailed observations on a subset of the sample tools unhafted defies both logic and functional (n = 8). Generally, all the pieces exhibited well-­ considerations. In addition, the limited micro­ developed micropolish, implying that the tools wear analysis documents haft-related polishes. were heavily used prior to discard. The detailed Although a very small number have symmetrianalysis suggests a “complex use history of cycles cally placed concavities along the lateral edges of use, maintenance and probably reuse in differ- that suggest some type of hafting, almost all are ent functional contexts” (Pope 2010:3). thinned on the proximal ends and shaped so that According to Pope, use polish was observed the entire piece is triangular. The bulb of percuson both dorsal and ventral surfaces. The distribu- sion and the striking platform are sufficiently altion and appearance of polishes on all “scrapers tered so that they are difficult to identify, and we suggest they were used on a hard material, simi- believe that this was an attempt to develop a stanlar to that seen with experimental wood polishes, dardized shape to fit existing implements. perhaps a dry, seasoned wood or, in some cases, “Retouched and utilized flakes” is the second fire-hardened wood” (Pope 2010:4). In particular, most common tool category (16.5 percent). This they appear to represent scraping, slicing, or shav- is a catchall category that includes drills, awls, and ing motions. The polish was bright and highly amorphous retouched tools (16.4 percent). Utireflective, especially suggesting charred wood. lized flakes are third, and approximately 65 perTable 4.2. Tool Categories and Frequencies from the Shoop Site.

90

Report on the 2008 Field Investigations at the Shoop Site

cent of these have low edge angles. Sidescrapers are the fourth most common prepared tool (12.9 percent). These are frequently thin flakes compared with endscrapers, and, generally, they do not retain any evidence of the bulb of percussion. Pointed sidescrapers are a subset of this category. These are usually thick, with steep retouching on two sides. Witthoft (1952:478) believed that they were used for scraping and gouging. Nonfluted late-stage bifaces make up approximately 11 percent of all tools. Many of these are wide and not systematically flaked. They do not appear to be blanks for fluted projectile points. Until a wear pattern analysis can be conducted, it is assumed that these functioned as general cutting tools — ​ that is, knives. Cox (1972) was the first to identify wedges at Shoop, and he felt that they were used to work bone. These implements number 143 and constitute 7 percent of all tools. Many of these are thick with crushed edges at both ends and appear to have been used to work a hard material such as wood, bone, or antler. However, some are very thin, less than 1 cm thick, and these may have functioned on some other, softer material. The function of “wedges” is problematic, and at some sites these may have been used as cores (Goodyear 1993). Jeff Kaline has examined the sample of wedges from the State Museum, and he is confident that they exhibit the same wear patterns as items that he has used in splitting bone and wood. Although some flake scars extend from end to end (suggesting that they functioned as small cores), the majority of damage is concentrated along the edges. We agree that many of these items appear to have functioned as wedges or splitting tools. There are 77 fluted points and point fragments in the collection of the State Museum, the Smithsonian Institution, the Wise collection, and the Shertzer collection. This represents 3.8 percent of all tools. Adding specimens from private collections, there may be as many as 96 fluted points that have been recovered from this site. In the four collections, there were 24 that were classified as entire, and nine were so fragmentary that measurements could not be taken. Proximal fragments (n = 24) are approximately equal to distal and medial fragments (n = 23). A concerted effort was made to refit fragments from the different collections but with little success. Cox (1986) col91

Table 4.3. Metrics of a Sample of 72 Fluted Bifaces from the Shoop Site.

Metric

Length from base Maximum width Maximum thickness Width at base Length of grinding Length of flute Width at top of flute Width at top of grinding Depth of basal concavity

Mean (mm)

Count

Range (mm)

44.94 23.22 7.02 22.4 21.22 20.25 21.58 22.72

23 24 33 47 41 37 49 52

32.9–68.9 18–29.9 5–9.2 15.6–27.2 14–29.9 11.6–37.2 14.1–31 17.2–31.2

3.7

23

1.2–6.1

lected a variety of metric data from 53 fluted projectile points. We were able to measure 72 specimens (Table 4.3). Needless to say, the results are similar to the metric analysis conducted by Cox. The collection is notable for the numbers of heavily reworked specimens. As discussed above, most researchers believe that the site was occupied for a relatively short time, early in the Paleoindian sequence. The metric data are probably not very useful in defining a temporal/cultural typology, but these data are useful in documenting the effects of projectile point rejuvenation and distance from the raw material source. Of note are two possible chipped adzes (Figure 4.9). These were identified in the Shertzer collection and were not previously reported. Both are made from a rough jasper that is more similar to the Houserville source in Centre County than to the Hardyston quarries in eastern P ­ ennsylvania. This artifact type has been reported from the Clovis component at the Topper site (Smallwood et al., this volume) and also from a Late Paleoindian context (Dalton phase) in Arkansas (Morse 1971). Several authors have commented on the number of multiple or multifunctional tools. These mainly consist of graver spurs or concavities on scrapers. Cox (1972) estimated that 20 to 40 percent of the scrapers were multiple tools. Many of these spurs or concavities are very small, and the spurs frequently occur at the distal corner of endscrapers. According to J. Witthoft (personal communication 1983), this may be a function of resharpening rather than utilization. Depending on

Carr et al. Table 4.4. Cataloged Artifacts from the Shoop Site by Weight (Onondaga Chert Only).

Type

Weight (g)

Debitage Retouched and utilized flakes Utilized flakes Endscrapers Sidescrapers General scrapers Wedges Early-stage bifaces Late-stage bifaces Other tools Total

682 700 590 561 375 314 293 161 151 62 3,889

% of Total Artifacts

17.5 18.0 15.2 14.4 9.6 8.1 7.5 4.1 3.9 1.5 100

Figure 4.9. Two jasper adzes from the Shertzer

collection.

debitage, with 82.3 percent of the ­assemblage by weight falling into the nondebitage ­category. These figures also de-emphasize the distance to the Onondaga chert quarries. Initially, the thought of thousands of prepared tools being carried hundreds of kilometers seems rather extraordinary and worthy of a special explanation. Considering that the total weight of the State Museum collection is 3.889 kg (8.6 lbs) and that of the Shertzer collection is 3.274 kg (7.2 lbs), this type of movement would not have represented any burden to a nomadic society. These two collections represent approximately 70 percent of the total known material, although this is an unknown percentage of the total number of artifacts remaining at the site. Nevertheless, it would seem unlikely that the total original weight of all Shoop artifacts could equal hundreds of kilograms. In summary, there is a large number and wide variety of tools from the Shoop site. Scrapers of all types and finished projectile points are ­notable and contribute nearly half of the assemblage. However, the remaining tools include wedges, bifaces, gravers, awls, drills, flakeshavers (limaces), and a large number of utilized flakes. Clearly, a wide variety of activities took place at Shoop.

the way they were hafted, the corners may have been differentially reworked and thus protrude beyond the scraper bit. The concavities along the lateral margins of scrapers could be part of the hafting mechanism. However, some of these are well shaped and probably functioned as gravers. The frequency of tools and utilized flakes to debitage is very high, with an overall frequency of 30 percent. This is a very high number for eastern Paleoindian sites. Calculated at 40 ­percent, only the Vail site in Maine has a greater frequency of tools (Gramly 1982). Certain individual loci at Debert (MacDonald 1968) and Bull Brook (Grimes et al. 1984) yielded approximately a 50 percent tool-to-debitage ratio. Admittedly, much of the Shoop collection is the result of a b ­ iased surface collection. Although many collectors pick up everything, some do not (i.e., the Gordon and Wise collections); therefore, tools are overly represented. The Shertzer collection represents a group of over 2,000 artifacts from a single location, and Mr. Shertzer picks up all artifacts. It is assumed that this is a reasonably representative sample of artifacts from this concentration. Tools represent 19.5 percent of the total artifacts. This is still high for an eastern Paleoindian site average. Table 4.4 is a list of artifact categories by weight from a sample of specimens from the State Museum collection (excluding finished projectile points). More than the above numbers, these figures emphasize the high frequency of tools over

Chronology

Lacking radiometric dates, the question of chronology was a major issue in the original report and was later reemphasized by Witthoft (1954) and Carr (1989). Based on projectile point morphology and tool assemblage complexity, 92

Report on the 2008 Field Investigations at the Shoop Site

Figure 4.10. Shoop fluted points from the State Museum collection.

subsequent investigators (Gardner and Verrey 1979; Mason 1962; Wilmsen 1970) suggested that the material was contemporary with the ­Clovis phase. Generally, the fluted points are ­parallel sided, with the widest point above the top of the ground edges (Table 4.3). They are not “fishtail” in form. The flute averages less than half the total length, and the ears are straight or slightly averted. The lateral edges are ground, and the grinding extends slightly past the flute. The sequence at the Thunderbird site in Virginia defined by Gardner (1979) reveals Clovis points (or the eastern equivalent) at the bottom of the stratified soil profile. Gardner argued that these are contemporary with or slightly later than western Clovis assemblages (Gardner assumed a

rapid west-to-east dispersion). Gardner and Verrey (1979:43) argue that a similar, although “extensively re-worked” style is found at Shoop, and, therefore, a date of around 11,000 Bp was suggested. Since Gardner’s work, the Shawnee-­ Minisink site, located 150 km northeast of Shoop in the Delaware Valley, has produced two Clovis points in a well-stratified context, and these are associated with dates averaging 10,950 Bp (Gingerich, Chapter 9, this volume). The most common fluted point type would be classified as Clovis (or the eastern equivalent), but there is a wide variety in shapes. Generally, they are shorter and narrower at the base than Clovis points (Figure 4.10). The length, width, and thickness measurements used to ­characterize the 93

Carr et al.

Early Paleoindian phase in the New England–­ Maritimes region (Bradley et  al. 2008) are all larger than the averages for the Shoop points. The average depth of the basal concavity is 3.7 mm, with a range of 1.2 — ​6.1 mm. Some, with deeply indented bases, are similar to those from the Debert site. Noting these specimens, Cox (1972) suggested that the site dated slightly later than Clovis. However, the Shoop specimens are at the lower end of the Vail-Debert-type basal concavity measurement (Bradley et al. 2008:132). In addition, at least one specimen is similar to the Crowfield type. Finally, there are two that are reminiscent of the Hardaway-Dalton style. These specimens are all in Onondaga chert. Based upon the patination, we do not believe that these were modified by later (Archaic) groups. D. Stanford (personal communication 2010) believes that the “guide flutes” and multiple fluting on some specimens are not a technological trait associated with Clovis and, in fact, represent a later fluted point characteristic. Concurring with Cox (1972), he would date the site closer to 10,700 radiocarbon years. On typological grounds, there seems to be a consensus that the fluted bifaces from Shoop are early in the Paleoindian cultural continuum, certainly within the Early Paleoindian phase as recently defined by Bradley et al. (2008). However, the majority of the points are at the end of their lithic use life, and a typological assessment may be of limited value. The exclusive use of Onondaga chert from western New York strongly suggests that Shoop was used for a limited period of time. Therefore, it does not seem likely that the differences in projectile point forms are temporal or that the site was occupied from Clovis through Hardaway-Dalton times (a period of at least 1,500 calendar years). These variations in shape could represent a combination of functional and social differences such as those discussed by Wilmsen and Roberts (1978) or as implied by Ellis (1987). Assuming the site represents a limited number of visits, the variety of shapes is most likely due to distance from the lithic source, frequent resharpening, and idiosyncratic or band-specific variation. The exclusive use of this single chert source suggests that this may represent an early fluted point–using population in the Susquehanna Valley. Therefore, the site should date to approxi94

mately 11,000 Bp. In addition, the large number of projectile points suggests that this is a hunting site and the focus could be caribou. During the warming Allerød episode, caribou may have been present in the Lower Susquehanna Valley. However, the likelihood of their presence increases during the Younger Dryas episode. Therefore, the occupation most likely dates to 10,900 Bp. Community Patterning and Site Function

Witthoft (1952:493) implied that Shoop was an overlook for hunting and trapping animals, but he did not elaborate on its function within a larger settlement system. Cox (1986:136) suggested that caribou were the most likely game animal. Commenting on the “rugged terrain,” Wilmsen suggests that it would have been more suitable for exploiting deer rather than large herding a­ nimals. He also noted that the generally small size of the individual artifact concentrations suggests small groups of people who are more likely to hunt smaller, solitary animals. However, as stated above, the dimensions of these concentrations are problematic and may not be related to group size. Carr (1986), Cox (1972), Fogelman (1986), and Meltzer (1988) have suggested that Shoop was associated with caribou hunting. Johnson (1982) has synthesized the data on the annual movements of caribou and native hunting strategies, and his analysis partially supports this hypothesis. In the summer, woodland caribou congregate in snowfields on north-facing slopes to avoid mosquitoes. This corresponds to the topographic setting of the Shoop site; however, the site does not exhibit the topography of caribou kill sites. Inuit/ Indian hunting strategy invariably involves river crossings or “funneling” the animals through narrow passages. There is a relatively narrow passage north of the site, but no artifacts have been found in the vicinity or in the saddle immediately adjacent to this “funnel.” Alternatively, using the caribou analogy, it is possible that the humans were the ones avoiding the mosquitoes. Originally, Witthoft (1952) defined 11 artifact concentrations generally corresponding to slightly elevated areas across the site, which were somewhat more than 100 m apart and generally less than 10 m in diameter. Wilmsen stated that they were about 70 m in size. We have found no evidence that Witthoft mapped these areas, and

Report on the 2008 Field Investigations at the Shoop Site

Figure 4.11. Map of Fogelman concentrations superimposed over topographic map (4.0 cm = 2,000 ft; 40-ft contour intervals).

we are uncertain as to where Wilmsen got his data. Fogelman (1986) interviewed the individuals who frequently surface collected the site, and he mapped 15 concentrations. Unfortunately, his map lacks a scale, and the size of individual loci can only be estimated. Superimposing the Fogelman map on a U.S. Geological Survey topographic map produces a rough approximation of the location of the concentrations (Figure 4.11). During the 2008–2009 investigation, we were able to develop a detailed topographic map of most of the site (Figure 4.12). One landowner

could not be reached for permission to access that property, and, therefore, this part of the site could not be mapped at the same level of detail. During the investigation, we also met with landowners and collectors in order to more accurately map artifact concentrations. In many cases, ­Fogelman’s concentrations were combined, although we ­consider this a very speculative map. Some of Witthoft’s and Fogelman’s concentrations may represent relatively high c­ oncentrations of a­ rtifacts, but some are defined by ­agricultural fields, roads, and tree lines. For example, Mr. Lesher reported 95

Carr et al.

Figure 4.12. Revised artifact concentrations based on the 2008 field investigations.

that artifacts could be found throughout his backyard, and that would combine several of Fogelman’s and Witthoft’s concentrations. There seems to be agreement on the location of some concentrations, such as the high point at the far north end of the site, overlooking the stream valley and the cul-de-sac. We were repeatedly told that “bushel baskets of chips and tools” were removed from this spot (although all agree that this was a great exaggeration). There are reports of finding artifacts almost anywhere on the site, although there is some agreement that the flat section in the center of the site contains the least number of artifacts. Based on this more refined level of topographic mapping, it is possible to divide the site into a northern and a southern section. Each of these is composed of several concentrations, although these are difficult to precisely define. The northern section is in the shape of an arc and appears to focus on the streams to the north and east of the site. The southern section focuses on

the spring located adjacent to the Shertzer South locus. According to Mr. Shertzer (former owner of part of the northern concentration), more fluted points were recovered from the northern section, especially the Shertzer North concentration. A preliminary analysis of the tools from the Shertzer South concentration produced a higher frequency of wedges, gravers, and concavities; a lower frequency of endscrapers compared with the northern section; and only two fluted points (Tables 4.5–4.6 and Figure 4.13). This may indicate functional differences between a processing area to the north and a more multifunctional or residential area to the south. These two may represent activities within the same visit or noncontemporary occupations. Obviously, additional research is necessary to support this hypothesis. At present, only the Lesher West concentration has been systematically tested. Considering the size of the artifacts and the lack of tools, we assume that this location has been affected by uncontrolled collecting. However, the distribution 96

Report on the 2008 Field Investigations at the Shoop Site Table 4.5. Tool Categories and Frequencies from the Shoop Site Excluding the Shertzer Collection.

Type

Count

Debitage Core Utilized flakes Retouched flakes General scrapers Sidescrapers Endscrapers Bifaces Wedges Gravers and concavities Fluted points Total

3,037 12 274 244 89 218 452 169 91 19 75 4,680

% of Tools

% of Total Artifacts

16.80 14.96 5.46 13.37 27.71 10.36 5.58 1.16 4.60 100.00

64.89 .26 5.85 5.21 1.90 4.66 9.66 3.61 1.94 .41 1.60 100.00

Table 4.6. Tool Categories and Frequencies from the Shertzer Collection.

of artifacts has enabled the demarcation of the size of the concentration at 12 × 25 m, or 300 m2. This gives us some idea of the size of the group involved, suggesting at best a microband or extended family of 15–25 people. The Shertzer South concentration is similar in size, and although not based on a controlled surface collection, it is the only group of artifacts that approaches a representative sample from a single concentration. Additional microwear analysis is planned for the future to explore functional differences between these two areas. Since only one of the concentrations has been systematically tested (with limited results), it is not possible to determine if the 11 to 15 concentrations represent single occupations, distinct activity areas, or large contemporaneous occupations. However, the large and varied collection of tools suggests that a significant number of ­person-hours were spent on-site. The large number of endscrapers and projectile points suggests that this is a processing site. However, the wide variety of tools suggests that it is a residential site that was occupied for some period of time (weeks). Assuming that Onondaga chert is not the exclusive lithic preference of these people, with each visit to the site there is an increasing probability that other lithic sources (high-­quality local cherts, chalcedonies, or jaspers) were located and utilized. Because there are very few other lithic types and there is a large and varied tool assemblage, it is reasonable to propose that the site was

Type

Count

Debitage Core Utilized flakes Retouched flakes General scrapers Sidescrapers Endscrapers Bifaces Wedges Gravers and concavities Fluted points Total

1,663 8 58 91 17 43 62 45 52 24 2 2,065

% of Tools

% of Total Artifacts

14.72 23.10 4.31 10.91 15.74 11.42 13.20 6.09 .51 100.00

80.53 .39 2.81 4.41 .82 2.08 3.00 2.18 2.52 1.16 .10 100.00

Figure 4.13. Bar graph comparing artifacts from the Shertzer South locus with total artifacts from the site.

visited a limited number of times by a group of band-size social complexity. The almost exclusive use of Onondaga chert strongly suggests that there was very little separation in time between periodic (annual?) visits. Without further fieldwork, this problem cannot be resolved. However, based on present knowledge, we estimate that the site probably represents fewer than 10 seasonal visits over less than one generation, or 20 years. In a very interesting analysis of the Bull Brook site in Massachusetts, Robinson et  al. (2009) convincingly make the case that this large site ­represents a single visit by a macroband (named group of 150–300 individuals) engaged in communal caribou hunting. Bull Brook contains 36 concentrations arranged in a circular pattern 97

Carr et al.

c­ overing 1.8 ha. Although the site is much smaller (compared with 8 ha at Shoop), it contains twice as many tools (n = 5,215) and 36,597 flakes (Robinson et al. 2009:424). Especially emphasizing the nonoverlapping loci in a circular pattern, but also the internal consistency of lithic types within each locus, repetitive tool associations along with special activity areas, they argue that this represents a single occupation. The same argument could be made for the Shoop site. There are numerous ­concentrations in what could be interpreted as a semicircular pattern, with a very large number of artifacts, especially tools. However, the open space in the middle of the arc at Shoop is 150 to 200 m across, and it would preclude the social interaction emphasized at Bull Brook. In addition, due to the general lack of provenience on the artifacts, it is not possible to identify artifact patterning. Finally, the north and south concentrations are topographically different. If Shoop represents a single occupation, it is most likely that there are functional differences between these two areas and that they do not represent individual family groups. The specific concentrations have been difficult to define, and there is some indication that they overlap. The blurring of these concentrations either by poor archaeological controls or by multiple occupations makes it difficult or impossible to resolve this issue.

located at the quarry, whereas the serial pattern produces base camps in settings especially favorable for hunting and gathering. Sites in the cyclical system tend to have a limited variety of lithic types, while those representing the serial system would have increased lithic variation. As summarized by Meltzer (1984), there are two general regional patterns that characterize the distribution of Paleoindian sites in the eastern United States. These could be more appropriately termed adaptive strategies rather than settlement patterns. For the Piedmont and Ridge and Valley sections of the east, Gardner (1979) has predicted hunting territories of 40 to 150 km. Considering the generally high biomass and probably the nonmigratory nature of the animals involved, this size seems reasonable. Most sites are small, which suggests small social groups. The ­quarry-​related base camp represents the largest site in this system. These groups were p ­ racticing a broad-­spectrum foraging subsistence pattern. In contrast, for Ontario, the Great Lakes, and New England, Ellis (2011), Roberts (1984), Robinson et al. (2009), Spiess (1984), and Storck (1984) suggest ­caribou-​ hunting territories that are 200 to 400 km in diameter. Most sites are small, but there are larger sites that are characterized by high ratios of tools to debitage, large numbers of finished fluted projectile points, multiple artifact concentrations, and lithic materials that apparently traveled hundreds of kilometers. Interestingly, a single lithic type may dominate these sites, but there are always several other types of local or more distant origin. This is unlike the lithic utilization pattern at Shoop. These sites are defined as hunting-­ related residential base camps, and they are most frequently in nonriverine settings. The Shoop site is most similar to sites in the glaciated region, and Witthoft identified these traits (not the site patterns) in 1952. Due to the apparently long distance involved in the seasonal round at Shoop, it is possible that some type of migratory game was involved. We suggest that caribou, elk, moose, and deer were common in the region, and any one or all may have been the focus of hunting activities. The contrast between the tool kit, lithic preferences, and topographic setting of Shoop compared with those of other sites in the region and generally with sites south of what Gardner (1979) has characterized as the “biotic Mason-Dixon

Settlement Pattern

Although not a major aspect of his 1952 publication, Witthoft presented one of the first settlement pattern models for the Paleoindian period. Shoop was a ridgetop hunting site associated with a culde-sac and an overlook. Witthoft’s model lasted for 20 years, until Gardner (1974, 1977) presented his settlement system based on the exploitation of riverine resources and high-quality lithic resources. He argued that the quarry-related base camp was the center of this system. Paleoindians moved away from the quarry to exploit a variety of food resources in riverine settings and returned to the same quarry in a cyclical fashion. Custer (1984) has since refined Gardner’s model. Briefly, Custer has proposed a cyclical pattern where movements are focused on a single large quarry and a serial pattern where a number of quarries are used. Base camps for the cyclical pattern are 98

Report on the 2008 Field Investigations at the Shoop Site

line” is striking. The differences could be cultural, or they could represent differences in adaptation. Most likely, they represent both. Shoop almost certainly represents a southern outlier of a northern adaptive strategy. In addition, the near exclusive use of Onondaga chert suggests that the inhabitants were unfamiliar with other lithic resources in the region. Shoop could well be one of the few examples of a colonizing event by fluted point–using peoples moving from the north into the Middle Atlantic region. This suggests that fluted points were introduced into central Pennsylvania through the Great Lakes and not from the south.

both cultural and functional differences between the glaciated and unglaciated regions. The Shoop site is possibly on the boundary of these two adaptations and represents the southern edge of an environment that supported (demanded) the exploitation of large territories. During the cold and dry Younger Dryas, it is assumed that there was an overall reduction in food resources. The Shoop site could represent a colonizing group expanding out of the Great Lakes following game south with the deterioration of the climate. Viewed from the south, the Shoop site is an enigma. The topography, the pattern of lithic utilization, and the tool frequencies are unexpected. Based on Snow’s (1981) description of the archaeological manifestation of a migration, this site could represent the last stop in a movement south. Dropping their old tools, they picked up new ones in order to adapt to a changing environment. However, since less than 1 percent of the lithic material represents local sources, they apparently utilized tools made elsewhere. Less likely, the site could also represent some sort of catastrophic event where a band in unfamiliar territory was unable to find sufficient resources and many individuals died. In this scenario, the high frequency of tools represents grave goods, and the high frequency of pot-lid fractures is a possible indication of cremation burials. The topographic setting may have been chosen for its idiotechnic characteristics. Because archaeologists frequently operate with a fragmentary data base, there is always the possibility that we are dealing with some type of idiosyncratic or nontypical behavior at Shoop. Such behavior may be “particularistic” and only marginally related to the overall adaptation of the population under study. When archaeological sites clearly fit a broader pattern, we are more confident that they represent a systematic set of behaviors. During the Paleoindian period, with people moving into a large unoccupied or lightly occupied environment (both very rare cultural events), there is a greater probability that sites representing atypical forms of behavior will be found. However, it is becoming increasingly apparent that Shoop does fit a larger pattern, albeit imperfectly, and its continued analysis will be instrumental in understanding the evolution of Paleoindian cultural ecology in the east.

Conclusion

In summary, the analysis of the Shoop material suggests that this is not primarily a blade technology, the Enterline Chert Tradition is not distinctive from other fluted point production strategies, and this site is not pre-Clovis in age but contemporary with other eastern Clovis sites or ­probably slightly later. However, Shoop’s temporal relationship to other sites cannot solely be based on typology. Fluted points from the Shoop site illustrate the diversity that occurs at single localities over a short period of time, possibly only a few successive occupations. The ecological setting, the artifact concentrations, the frequency of tools to debitage, the large numbers of endscrapers and finished projectile points, and the lithic utilization pattern of the Shoop site are very different from those in sites to the south. These traits are broadly similar to those of sites in what Meltzer (1984) has termed the “glaciated region.” The 2008 field project improved our understanding of the site’s setting and community patterning. The concentrations can be organized into a southern section associated with a spring, possibly representing a residential area, and a northern section, possibly used as a staging area to exploit the adjacent bogs and low-order streams. Based on the microwear analysis, the p ­ rocessing of wood for the production of tools may have been a common activity. However, additional microwear analysis is planned, and this should produce a more detailed understanding of site ­function. The site represents a diagnostic Paleoindian site type that under current models is distinctive for the glaciated region and is probably related to 99

Carr et al.

Acknowledgments There are several individuals without whose cooperation this project could not have been completed. First, without the permission of the landowner Ronald Lesher, this project could not have been accomplished. He was very gracious in allowing us to dig up his field and store our equipment in his barn. For many weeks he had people intruding on his property, and he was always very cordial. William Cook plowed and backfilled the field for us, and we appreciate his skill with farm machinery and patience in getting reimbursed. We are indebted to the National Park Service, Northeast Region, Challenge Cost Share Program, which supported this project with a grant. Specifically, we appreciate the flexibility of Amanda Casper, who administered the grant. She was always very helpful. Arthur Shertzer and his wife allowed us to work with his collection and map the concentration on his property. We are very thankful for the many hours we spent in his dining room going over his collection. Joel, Elaine, and Eric Steigman allowed us to dig up their pasture again to expose the buried swamp. We very much enjoyed their hospitality. Joe Gingerich deserves a special thanks. We sincerely appreciate being invited into this work. He was very patient and very supportive. We apologize for any frustration we might have caused. Our respective institutions, the State Museum of Pennsylvania, Mercyhurst College, and Clarion University of Pennsylvania, were all very supportive and continue to encourage this research. At the State Museum, we would like to especially thank Jack Leighow, former director of the State Museum of Pennsylvania, and Bill Sisson, former curatorial division chief, for their support and encouragement. The topographic map was produced by Brian Fritz with the help of Mike Ward. This was done under very cold and windy conditions, and we certainly appreciate their contribution to making the first accurate map of the site. The Mercyhurst field school was super­vised by Al Quinn. They began the first controlled testing at the Shoop site and performed much of the mundane initial field testing without complaining. The Section of Archaeology at the State Museum has many important responsibilities, and the senior author knows that his colleagues take these responsibilities very seriously. Janet Johnson, James Herbstritt, Elizabeth Wagner, ­David Burke, and Andrea Johnson pulled themselves away from their daily responsibilities and “volunteered” on a regular basis. We especially appreciate David Burke’s assistance with editing this document. Finally, this project could not have been completed without the dedication of a group of loyal volunteers that included Bradley Miller, Mike Ward, Mark Anderson, Naomi Spotts, Mary Pat Evans, and Will Pratt.

They showed up nearly every day, regardless of the heat or the jobs that needed to be done.

References Cited Bordaz, Jacques 1970 Tools of the Old and New Stone Age. Natural History Press, Garden City, New York. Bouldurian, Anthony 1985 Variability in Flint Working Technology at the Krajacic Site: Possible Relationships to the Pre-Clovis Paleoindian Occupation of the Cross Creek Drainage in Southwestern Pennsylvania. Unpublished Ph.D. dissertation, Department of Anthropology, University of Pittsburgh. Bradley, J. W., A. E. Spiess, R. A. Boisvert, and J. ­Boudreau 2008 What’s the Point? Modal Forms and Attributes of Paleoindian Bifaces in the New ­England–​Maritimes Region. Archaeology of Eastern North America 36:119–172. Callahan, Errett 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. Carbone, Victor 1976 Environment and Prehistory in the Shenandoah Valley. Unpublished Ph.D. dissertation, Department of Anthropology, Catholic University of America, Washington, D.C. Carr, K. W. 1986 Core Reconstruction and Community Patterning at the Fifty Site. Journal of Middle ­Atlantic Archaeology 2:79–92. 1987 Lithic Reduction Sequences and Quarry Utilization at the Shoop and “50” Paleoindian Sites. Paper presented at the 52nd Annual Meeting of the Society for American Archaeology, Toronto. 1989 The Shoop Site: 35 Years After. In New ­Approaches to Other Pasts, edited by W. F. Kinsey and R. Moeller, pp. 5–28. Archaeological Services, Bethlehem, Connecticut. 1992 A Distributional Analysis of Artifacts from the Fifty Site: A Flint Run Paleoindian Processing Station. Ph.D. dissertation, Catholic University of America, Washington, D.C. University Microfilms, Ann Arbor. Carr, K. W., and James M. Adovasio 2002 Paleoindians in Pennsylvania. In Ice Age Peoples of Pennsylvania, edited by Kurt W. Carr and James M. Adovasio, pp. 1–50. Pennsylvania Historical and Museum Commission, Harrisburg.

100

Report on the 2008 Field Investigations at the Shoop Site Salwen. Annals of the New York Academy of Sciences 288:257–263. 1979 Paleoindian Settlement Pattern and Site Distribution in the Middle Atlantic. Paper presented at the Middle Atlantic Archaeological Conference, Rehoboth Beach, Delaware. Gardner, William M., and Robert A. Verrey 1979 Typology and Chronology of Fluted Points from the Flint Run Area. Pennsylvania Archaeologist 49(1–2):13–45. Giddings, James L. The Denbigh Flint Complex. American An1951 tiquity 16(3):193–203. Goebel, T., M. R. Waters, and D. H. O’Rourke 2008 The Late Pleistocene Dispersal of Modern Humans in the Americas. Science 319:1497– 1502. Goodyear, Albert 1993 Tool Kit Entropy and Bipolar Reduction: A Study of Interassemblage Variability Among Paleo-Indian Sites in the Northeastern United States. North American Archaeologist 14(1):​ 1–23. Gramly, Richard M. 1982 The Vail Site: A Palaeo-Indian Encampment in Maine. Bulletin of the Buffalo Society of Natural Sciences, Vol. 30. Buffalo. Grimes, John R., W. Eldredge, B. G. Gremir, and A. Vaccaro 1984 Bull Brook II. Archaeology of Eastern North America 12:159–183. Guilday, J. E. 1984 Pleistocene Extinction and Environmental Change: Case Study of the Appalachians. In Quaternary Extinctions: A Prehistoric Revolution, edited by P. S. Martin and R. G. Klein, pp. 250–258. University of Arizona Press, Tucson. Haag, C. M. 2004 An Analysis of Clovis Blade Reduction Sequences from the Adams Site (15Ch90) and the Joe Priddy Site (15Hd583): Implications for Early Paleoindian Mobility and Lithic Procurement in Kentucky. Unpublished ­Master’s thesis, University of Kentucky, ­Lexington. Haynes, G. 2002 The Early Settlement of North America: The Clovis Era. Cambridge University Press, Cambridge. Holland, J. D., and D. Dincauze 1999 A Shoop Site Postulation — ​Again. Paper ­presented at the 64th Annual Meeting of the Society for American Archaeology, ­Chicago.

Carr, K. W., C. A. Bergman, and C. M. Haag 2010 Some Comments on Blade Technology and Eastern Clovis Lithic Reduction Strategies. Lithic Technology 35(2):91–125. Collins, M. B. 1999 Clovis Blade Technology. University of Texas Press, Austin. Cox, Stephen L. 1972 A Re-Analysis of the Shoop Site. Manuscript on file at the State Museum of Pennsylvania, Harrisburg. 1986 A Re-Analysis of the Shoop Site. Archaeology of Eastern North America 14:101–170. Custer, Jay 1984 Delaware Prehistoric Archaeology: An Ecological Approach. University of Delaware Press, Dover. Davis, Margaret Bryan 1983 Holocene Vegetational History of the Eastern United States. In Late Quaternary Environments of the United States, edited by H. E. Wright, Jr., pp. 166–181. University of Minnesota Press, Minneapolis. Dickens, W. A. 2005 Biface Reduction and Blade Manufacture at the Gault Site (41BL323): A Clovis Occupation in Bell County, Texas. Unpublished Ph.D. dissertation, Texas A&M University. Ellis, Chris J. 1987 The Explanation of Northeastern Paleoindian Lithic Procurement Patterns. Paper presented at the 52nd Annual Meeting of the ­Society for American Archaeology, ­ oronto. T 2011 Measuring Paleoindian Range Mobility and Land Use in the Great Lakes/Northeast. ­Journal of Anthropological Archaeology 30:​ 385–401. Fogelman, Gary L. 1986 Shoop: Pennsylvania’s Famous Paleo Site: A Popular Version. Fogelman Publishing. Gardner, William M. 1974 The Flint Run Complex: Pattern and Process During the Paleo-Indian to Early Archaic. In The Flint Run Paleo-Indian Complex: A Preliminary Report 1971–1973 Seasons, edited by W. M. Gardner, pp. 5–47. Occasional Publication No. 1. Archaeology Laboratory, Department of Anthropology, Catholic University of America, Washington, D.C. 1977 Flint Run Paleoindian Complex and Its Implications for Eastern North American Prehistory. In “Amerinds and Their Paleoenvironments in the Northeastern North America,” edited by W. S. Newman and Bert 101

Carr et al. Howard, Edgar B. 1942 A “Fluted Point” Site in Pennsylvania. Pennsylvania Archaeologist 12(1):4–6. Isarin, R. B., and S. J. P. Bohncke 1999 Mean July Temperatures During the Younger Dryas in Northwestern and Central Europe as Inferred from Climate Indicator Plant S­ pecies. Quaternary Research 51(2):158–173. Jennings, T., C. Pevny, and W. Dickens 2010 A Biface and Blade Core Efficiency Experiment: Implications for Early Paleoindian Technological Organization. Journal of ­Archaeological Science 37:2155–2164. Johnson, Michael 1982 The Eastern Paleo-Indian and Caribou. Manuscript on file at the State Museum of Pennsylvania, Harrisburg. Kent, Barry C., Ira A. Smith, III, and Catherine ­McCann (editors) 1971 Foundations of Pennsylvania Prehistory. ­Anthropological Series of the Pennsylvania Historical and Museum Commission, No. 1. Harrisburg. Krieger, Alex 1954 A Comment on “Fluted Point Relationships” by John Witthoft. American Antiquity 17(3):​ 273–275. Lavin, L., and D. R. Prothero Microscopic Analysis of Cherts with and Ad1981 jacent to the Delaware River Watershed. Man in the Northeast 21:3–17. Lundelius, Ernest L., R. W. Graham, E. Anderson, J. Guilday, J. A. Holman, D. W. Steadman, and S. D. Webb 1983 Terrestrial Vertebrate Faunas. In Late Quaternary Environments of the United States, edited by H. E. Wright, Jr., Vol. 1:311–353. University of Minnesota Press, Minneapolis. MacDonald, G. F. 1968 Debert: A Paleo-Indian Site in Central Nova Scotia. National Museum of Canada, Anthropological Papers, 16. Ottawa. Mason, Ronald J. 1962 The Paleo-Indian Tradition in Eastern North America. Current Anthropology 3:227–283. Mayewski, P. A., L. D. Meeker, S. Whitlow, M. S. Twickler, M. C. Morrison, R. B. Alley, P. Bloomfield, and K. Taylor 1993 The Atmosphere During the Younger Dryas. Science 261:195–197. McCary, Ben C. A Workshop Site of Early Man in Dinwiddie 1951 County, Virginia. American Antiquity 17(1), Pt. 1: 9–17.

Meltzer, David J. 1984 On Stone Procurement and Settlement Mobility in Eastern Fluted Point Groups. North American Archaeologist 6(1):1–24. 1988 Late Pleistocene Human Adaptations in Eastern North America. Journal of World Prehistory 2:1–52. Moeller, Roger W. 1989 The Shoop Conundrum. Pennsylvania ­Archaeologist 59(1):70–77. Morse, Dan F. The Hawkins Cache: A Significant Dalton 1971 Find in Northwest Arkansas. Arkansas Archaeologist. Bulletin of the Arkansas Archaeological Society 12(1):10–20. Peterson, James B., Robert N. Bartone, and Belinda J. Cox 2002 The Late Paleoindian Period in Northeastern North America: A View from Varney Farm. In Ice Age Peoples of Pennsylvania, edited by Kurt Carr and James Adovasio, pp. 123–143. Pennsylvania Historical and Museum Commission, Harrisburg. Pope, Melody 2010 A Microwear Study on a Small Sample of End Scrapers from the Shoop Site. Manuscript on file at the State Museum of Pennsylvania. Prasciunas, Mary M. 2007 Bifacial Cores and Flake Production Efficiency: An Experimental Test of the Technological Assumptions. American Antiquity 72:​ 334–338. Ridge, John C. 2003 The Last Deglaciation of the Northeastern United States: A Combined Varve, Paleomagnetic and Calibrated 14C Chronology. In Geoarchaeology of Landscapes in the Glaciated Northeast, edited by David L. Cremeens and John Hart, pp. 15–48. New York State Museum Bulletin 497. Albany. Roberts, Arthur 1984 Paleo-Indians on the North Shore of Lake Ontario. Archaeology of Eastern North America 12:248–265. Robinson, Brian S., Jennifer C. Ort, William E. Eldridge, Adrian L. Burke, and Bertrand G. Pelletier 2009 Paleoindian Aggregation and Social Context at Bull Brook. American Antiquity 74(3):423– 447. Sirkin, Les 1977 Late Pleistocene Vegetation and Environments in the Middle Atlantic Region. In Amerinds and Their Paleoenvironments in Northeastern North America, edited by W. S.

102

Report on the 2008 Field Investigations at the Shoop Site Newman and Bert Salwen, pp. 206–217. New York Academy of Sciences, New York. Snow, Dean 1981 Foundations of Northeast Archaeology. Academic Press, New York. Spiess, Arthur 1984 Arctic Garbage and New England Paleo-­ Indians: The Single Occupational Option. ­Archaeology of Eastern North America 12:​ 280–285. Spiess, Arthur E., Deborah B. Wilson, and James W. Bradley 1998 Paleoindian Occupation in the New ­England–​Maritimes Region: Beyond Cultural Ecology. Archaeology of Eastern North America 26:201–264. Steffy, Kenn, and Albert Goodyear 2006 Clovis Macro Blades from the Topper Site, 38AL23, Allendale County, South Carolina. Current Research in the Pleistocene 23:147– 149. Storck, P. L. 1984 Glacial Lake Algonquin and Early Paleo-­ Indian Settlement Patterns in South-Central Ontario. Archaeology of Eastern North America 12:286–298. Topping, W 1995 A Revisionist View of Shoop. Ohio Archaeologist 45(2):23–24. U.S. Department of Agriculture 1972 Soil Survey of Dauphin County. U.S. Department of Agriculture, Pennsylvania Department of Agriculture, Pennsylvania State University, State College. Vento, F. J., Harold B. Rollins, Anthony J. Vega, James M. Adovasio, Patricia A. Stahlman, David B. Madsen, and Jeffry S. Illingworth 2008 Development of a Late Pleistocene–­Holocene Genetic Stratigraphy Framework for the Mid-Atlantic Region: Implications in Archaeology. Paper presented at the 73rd Annual Meeting of the Society for American

Archaeology, Vancouver, British Columbia, March 26–30. Copies on file at Clarion University and Mercyhurst Archaeological Institute, Mercyhurst College. Vento, F. J., T. Vega, R. S. Toomey, L. P. Fay, P. A. Delcourt, H. R. Delcourt, G. S. Brush, and F. B. King 1994 Paleoenvironmental and Paleoclimatic Reconstruction of Pennsylvania Over the Last 15,00 Years. Manuscript on file at the State Museum of Pennsylvania, Harrisburg. Verrey, R. 1986 Paleoindian Stone Tool Manufacture at the Thunderbird Site (44WR11). Unpublished Ph.D. dissertation, Catholic University of America, Washington, D.C. Watts, W. A. 1979 Late Quaternary Vegetation History of Central Appalachia and the New Jersey Coastal Plain. Ecological Monographs 49:427–469. Webb, William S. The Parrish Village Site. University of Ken1951 tucky, Reports in Anthropology and Archaeology 7(6):407–451. Wilmsen, Edwin N. 1970 Lithic Analysis and Cultural Inference: A ­Paleo-​Indian Case. University of Arizona Press, Tucson. Wilmsen, Edwin N., and F. H. H. Roberts, Jr. 1978 Lindenmeier, 1934–1974: Concluding Report on Investigations. Smithsonian Contributions to Anthropology No. 24. Washington, D.C. Witthoft, J. 1952 A Paleo-Indian Site in Eastern Pennsylvania: An Early Hunting Culture. Proceedings of the American Philosophical Society 96(4). Philadelphia. 1954 A Note on Fluted Point Relationships. American Antiquity 19(3):271–273. Wray, C. 1948 Varieties and Sources of Flint Found in New York State. Pennsylvania Archaeologist 18(1– 2):​25–45.

103

5

Spatial Organization at Bull Brook Brian S. Robinson and Jennifer C. Ort

Ongoing research on the Bull Brook site in Ipswich, Massachusetts, is directed toward an understanding of the spatial organization of different activities and what this reveals about Paleoindian social relationships, hunting practices, and ritual domains. In accord with the theme of reinvestigating classic Paleoindian sites, Bull Brook is informative both for the remarkable organization of the largest Paleoindian settlement yet discovered in North America and for the history of changing ideas that allowed a site, salvaged by a dedicated group of avocational archaeologists in the 1950s, to persevere through the wildly changing discipline of archaeology over the past 60 years (Byers 1954, 1955; Curran and Grimes 1989; Eldridge and Vaccaro 1952; Grimes 1979; Grimes et al. 1984; Jordan 1960; Robinson et al. 2009; Spiess et al. 1998). In this chapter we review changing views of Bull Brook spatial organization, recent analyses of activity patterning associated with the large aggregation site at Bull Brook (with 36 house sites or activity clusters), and the critical role the excavators played in the recent analysis. The transformation of archaeological interpretation is one of the standard tasks of archaeology. In 1927, Folsom became an (archaeological) household word when stone tools were convincingly associated with an extinct species of bison in New Mexico. This fact is now so commonplace and without controversy that Folsom’s legacy might have been limited to one of the premier case studies of scientific proof and revolutionary

change in perspective. Recent analyses and excavations have greatly expanded the questions that can be addressed at Folsom, with the claim that “Folsom is no longer just one of the best-known sites in American Archaeology; it is also now reasonably well understood” (Meltzer 2006:21). Folsom and subsequent discoveries at Clovis (Boldurian and Cotter 1999; Cotter 1937) were fresh topics when Joe Vaccaro discovered the first fluted point at Bull Brook on November 17, 1950. The mere presence of Bull Brook in Massa­chusetts and the Late Paleoindian Reagan site in northern Vermont (Ritchie 1953) posed questions about the existence of Pleistocene hunters in the glaciated Northeast. By early 1953, Bull Brook began to break ground it would take decades to appreciate, when the excavators recognized an arc-like pattern of artifact concentrations that resembled part of a large camp circle. By 1958 the pattern grew to an oval or circular arrangement of 40 concentrations or loci. Settlement pattern archaeology was in its infancy in the 1950s (Trigger 2006:21), and the idea of a large spatially organized Pleistocene settlement was so challenging that it was more expedient for the professional archaeologists involved to relegate the pattern to coincidence than to find a means to test it (Byers 1954; Jordan 1960:​ 192–193; Robinson et al. 2009:431). This observation is not intended to criticize the professional archaeologists, who made crucial contributions to the Bull Brook record, but to acknowledge the profound changes in interests that accompanied

104

Spatial Organization at Bull Brook

the development of archaeology over the years. Spatial organization has become a fundamental part of Paleoindian research, which in the case of Bull Book means turning still photographs, notes, and memories into geographic information system (GIS) coordinates. Any effort to further refine or scrutinize spatial relationships at Bull Brook must incorporate spatial clues derived from the timing of archaeological work and the rapidly changing landscape during sandpit operations at the site in the 1950s. The passing of a single day might dictate whether a locus was open for excavation or irretrievably lost, and critical clues may include clothing worn on a particular day or the dates that one of the excavators was away on a job or vacation. Frank, Joe, Nick, and Tony Vaccaro and Bill Eldridge, Billy DiPaolo, and Tony Orsini were the principal excavators, while Doug Byers, director of what is now the Robert S. Peabody Museum, and Doug Jordan, then a Harvard graduate student, contributed guidance and records of their own. The Bull Brook site was excavated just ahead of bulldozers that mined it for sand and gravel. Most of the excavators and the sandpit owners were recent Italian immigrants who had known each other from their youth in Beverly, Massachusetts, a crucial cooperation in the history of the site. Before Bull Brook was discovered, Nick Vaccaro and Bill Eldridge worked with Ripley Bullen at the Great Neck Shell Heap (Bull Brook Records [BBR] 417) in Ipswich and Johnson’s Spring (Bullen 1950; BBR 216) through the Robert S. Peabody Museum.1 Bullen is frequently cited in Eldridge’s notes as the inspiration for keeping “copious notes” on Bull Brook (letter, June 19, 1952, BBR 258). Bill Eldridge was the principal record keeper from the day the first fluted point was discovered. He kept his copious notes long after the excavations were completed, tracking the collections and the cast of characters that became involved in the research. Bill Eldridge passed away on April 22, 2009. Joe, the last surviving of the Vaccaro brothers, died on September 8, 2009. Now we have the notes, the ­movies, the photographs, and recorded ­memories. We have a lot of them. For 30 years the Bull Brook Club met twice a week in a room at the side of Nick Vaccaro’s workshop, a haven for those interested in Paleoindian studies. In the late 1970s and early 1980s the club

attracted John Grimes, Peter Fetchko, and Frederick West of the Peabody Essex Museum. All of the excavators eventually donated their collections to this museum, and Bill Eldridge was employed there for over two decades, maintaining continuity and contact with artifacts and records. Mary­ lou Curran’s (1984) excavations at the Whipple Paleoindian site in New Hampshire brought her into the fold, and she and Brian Robinson were employed at the museum working with curator Frederick West in 1985. Eldridge also maintained continuity with the Robert S. Peabody Museum after Doug Byers died in 1978 (MacNeish 1979), through directors Richard MacNeish, James Bradley, and Malinda Bluistain. For Robinson, the learning curve working with Eldridge continued to grow with every meeting right up to Eldridge’s ninetieth birthday party, as finer details that were unimportant at first became critical clues when fit into the fabric of records and activities. It may seem unlikely that inquiries about a particular locus on a particular day could be aided by memories a half century later, but such was the case on numerous occasions. Memories are not perfect, however, and as the richness of memories and stories grows, what is to stop the analyst from picking the memories that fit a particular model? What if memories are correct only 90 percent of the time, or 60 percent, or 30 percent? In this case it became Robinson’s explicit task to scrutinize modern memories for contradictions between photographs, original field records, and earlier memories. Digitized field r­ ecords were searched for key words and dates to confirm memories or sometimes to confront Eldridge with his own words. Eldridge worked tirelessly to increase the content of each event, to find creative means to cross-check data, and to debunk faulty associations made by younger archaeologists in search of solutions to key problems. Dated documents are the cornerstones of chronology, but as Doug Jordan observed, dates recorded in the field are not always complete or accurate and need to be cross-checked. In one case an important set of field notes was dated March 22, 1953 (BBR 229), including a report of discoveries made in the vicinity of Locus 9. The same event was described in a letter to Nick Vaccaro (BBR 370), who was then in Florida. But postcards from Nick and Anna Vaccaro (one

105

Robinson and Ort

postmarked shortly after they arrived in Florida on November 23, 1953) demonstrate that the page of field notes is more confidently dated to March 22, 1954, a year later than the recorded date. It is sometimes the year that is written in error, when the new year rolls around and the more habitually written date is repeated. Notes in Doug Jordan’s handwriting on this same page of Bill’s field notes represent an effort to correlate dates, a frequent practice of Jordan’s that we repeated 40 years later. Cross-checking and correcting errors from 50 years ago require considerable documentation. The determination of Bill Eldridge and the V ­ accaros to get the record straight, weathering constant scrutiny by younger archaeologists, permitted refinements beyond the written record. Increasing interest in spatial context made this scrutiny necessary. The original Bull Brook site plan remained almost unknown for 20 years. It was first published by Grimes (1979) in a major article that initiated new interest in Bull Brook. The plan itself remained largely undescribed, however. Some accepted the plan with little question, for example, in a Time-Life volume, The First Americans, Bull Brook is cited as a gathering at a joint campsite that “could accommodate as many as 225 people” (1992:24). With growing evidence of large northeastern Paleoindian sites, the potential for a structured settlement pattern at Bull Brook was reinforced. It remained a premier example of a possible large aggregation site (Curran 1999:6; Dincauze 1993; Ellis and Deller 2000:​ 242; Spiess et al. 1998; Storck 2004) but was insufficiently published to further evaluate. One of the major goals of our recent grant proposal to the National Science Foundation was to evaluate the structure of the ring-shaped pattern at Bull Brook: Is the Ring-Shaped Settlement Pattern Real? This question may be particularly irritating to the excavators who spent ten years uncovering the pattern, considerable effort convincing skeptical professionals, and then at last (with the blessing of new theory) to have it raised again as a central problem. But the major thrust of the proposal is to test whether the ring-shaped pattern was culturally real to the Paleoindians, and this is a multifaceted problem. There is no question that a large oval

pattern of concentrations was excavated, but the problem requires demonstration that the empty place in the middle was well tested, along with the flat expanses outside. The position of each of the 42 loci must be documented to evaluate internal relationships. The plan must be placed accurately on the landscape, searching for possible topographic or hydrologic controls that could have influenced site selection [Robinson 2003]. Major aspects of that investigation are ongoing, especially the analysis of lithic material sources and distributions by Adrian Burke, along with the smaller-scale variability of artifact distributions. Here we focus on the original problems of mapping and artifact distributions that bear on largescale activity patterning. The original scaled site plan of Bull Brook was produced by Doug Jordan in 1959 with the assistance of all of the excavators. After being remapped from photographs and original site ­records (Figure 5.1; Robinson et  al. 2009), the ring-shaped pattern survived intact, justifying the decades of emphasis placed on the original published plan. A major scale error was identified, however, with the revised plan being 40 percent smaller than the 1959 plan. The revised plan has 36 house-sized activity areas covering an area of 170 × 135 m. Twenty-five of the loci are now located with GIS coordinates, while 11 have good relative positions, among specific trees and landmarks on aerial photographs from the 1950s. Relative positions remain largely the same as in the 1959 plan, but subtle shifts based on well-defined landmarks provide added confidence in the configuration of segments around the ring. It was clearly recognized by the excavators that different loci had strongly contrasting proportions of different artifact classes. In the more normative cultural models of the time, Byers considered the strong differences in artifact frequency most likely to represent “change in ways of living brought about by the passage of time” (letter, March 6, 1972 [BBR 784], in Robinson and Eldridge 2005:69). If this explanation for different artifact frequencies was correct, there should be no discernible structure to activities within the ring-shaped settlement pattern. On the other hand, if the well-defined settlement plan repre­

106

Spatial Organization at Bull Brook

Figure 5.1. Location map for the Bull Brook site showing areas and dates stripped for sand removal.

sents an organized event, we may find ­activities that are structured relative to the circular pattern as a whole. With Jennifer Ort’s reanalysis and computer coding of the full Bull Brook assemblage, combined with the newly mapped configuration of loci, we conducted the first full analysis of artifact distributions. Years of mapping, checking provenience, and cataloging culminated in a series of computer sorts that with relative suddenness had the potential to reveal either hidden structure within the settlement plan or more arbitrary and ambiguous patterns. The results showed more pronounced evidence of internal organization than we had anticipated. Differences in activities were tested by contrasting the set of loci situated on the inside of the circle with the larger sample from the outer circumference. Eight loci on the interior were strongly represented by fluted point fragments, channel flakes made in the process of fluting, drills, and flakeshavers (limaces), while the outer ring was most strongly represented by endscrapers (Robinson et al. 2009).

The concentric activity patterning, with ­inner and outer rings of specialized activities, provides strong support for the excavator’s claim that Bull Brook represents a single organized event, and a significant focus for investigating social relations. Possible explanations for the concentric patterning include the separation of h ­ unting preparations and processing activities, the r­ itual treatment of animals, and associated gender roles (Robinson et al. 2009). The extreme size of Bull Brook may further represent the gathering of a maximum band, as proposed for Paleolithic and Paleoindian social systems (Anderson 1995; Wobst 1974). Spatial organization at Bull Brook provides a rare window on multiple scales of social organization, all of which rest, in turn, on the accuracy of the mapping and artifact proveniences. In this chapter we continue the documentation of that process, presenting additional data on how the concentric patterning was defined. There are certainly details of Bull Brook that will never be known, but with documentation of the assemblages, ­records, and memories of each

107

Robinson and Ort

locus at Bull Brook, there is much potential for refining and testing patterns so that such early excavations can remain “reasonably well understood” (Meltzer 2006:​21) and contribute to ­ever-​developing archaeological questions. Bull Brook: Evaluating Memories and Records

The basic effort to remap Bull Brook from the original records using GIS-registered aerial photography, visible landmarks, and intersecting lines of sight derived from still photographs and home movies has been described (Robinson et al. 2009). Here we focus on one of the variables that made the present study possible, the remarkable memories of the excavators. Eldridge and the Vaccaros remembered almost every locus by its field name, number, and excavator (e.g., Bill’s Deep Fire Pit [#11], Nick’s Red Point [#15], Gus’s Bones [#32]) along with associated stories. Since the final numbering system was not assigned until 1959, one of the necessary activities was to annotate the field notes to properly identify field records and correlate them with photographs. Some loci were well recorded and easily identifiable, but others required a concerted search of the records, remembering that while Bill Eldridge was the principal record keeper, most of the loci were excavated by other excavators. Memories and context clues were sometimes needed to interpret the records, recognizing the subtlety with which firsthand memories can resolve problems as well as the potential to make modern errors. Locus 22, excavated by Tony Orsini, is one of the more copiously recorded loci, both with references in Bill’s field records and because Orsini made a practice of dating each day’s finds. The locus was discovered on September 28, 1957 (BBR 723), after which Orsini variously referred to the location as “the same spot as September 28” or as “the Boundary Mound” (Jordan 1960:123) or “boundary line” because it was situated on what was interpreted as a boundary ditch between two pieces of property. Locus 22 was excavated on at least 23 days over a period of two years and is thus one of the more easily identified loci in the field records. High visibility has its problems, however. Bill Eldridge recorded in his notes attributed to October 6, 1957, that “on arrival at the site we went right to work, Tony in his spot near the b ­ oundary

line behind Frank’s old hole, where he had 2 small trees to work around” (BBR 537). In the recent effort to correlate field notes with locus numbers, Robinson matter-of-factly attributed this reference to Tony Orsini’s excavation of Locus 22. When discussed with Bill Eldridge, however, he immediately responded with equal assurance that this record was not of Tony Orsini but, rather, of Tony Vaccaro because the field notes recorded that Tony came to the site with Frank and Joe and that Tony Orsini never came with the Vaccaros because they lived in different towns. This became clear with a closer reading, but it was Bill’s familiarity with multiple contexts that exposed Robin­ son’s error. The record was in fact one of very few references to Tony Vaccaro’s Locus 23, which led to the identification of the only known photograph of Locus 23, taken by Tony Vaccaro (BBR 2582) on the same date as Eldridge’s field notes. Tony Vaccaro’s locus was near the boundary line, not on the boundary line. In this and many other cases, it was the excavator’s intimate knowledge of friends and circumstances that enabled correct identification of a locus. In other cases the memories were direct accounts of unique events. Locus 5 became universally known as the “Tent Site” for the fact that it was literally discovered under a tent. However, Bill’s earliest notes occasionally refer to Locus 5 as “the #2 tent section” (BBR 181) or “the second tent section” (BBR 176). One artifact now cataloged in Locus 2 was originally attributed to “tent section #1, May 1952” (BBR 180), but there were conflicting memories of where the “first tent” might have been, and further evidence was sought to confirm the association with Locus 2. A black-and-white print (BBR 2425) showed excavations taking place with a large tarpaulin and birch stakes lying in the background. The dimensions and character of the photograph suggested that it may have been taken with Joe Vaccaro’s Brownie 127 camera. Photographs of an important tool group found by Nick Vaccaro in Locus 2 were also taken with this camera (BBR 2422; Figure 5.2). With these photographs in hand Eldridge and Robinson discussed the problem with Joe Vaccaro on August 31, 2006. Joe agreed that the photograph was probably taken with his camera, even though he was in one of the photographs. He identified the shadow of his brother Tony as the photographer. But Joe did

108

Spatial Organization at Bull Brook

Figure 5.2. Nick Vaccaro’s tool group from Locus 2. Photograph taken on May 17, 1952, with Joe Vaccaro’s Brownie camera and 127 film. The three complete artifacts from left to right include a scraper (PEM 6843), a large flakeshaver (PEM 6842), and a fluted point (PEM 6841). (Photograph [BBR 2422] courtesy of the Peabody Essex Museum.)

not think that the photograph with the tarpaulin was taken on the same day that Nick’s tool group was found because he said it was not raining the day of Nick’s discovery. Bill concurred: BILL: No I think it was a pleasant day. JOE: It was a very pleasant day, and I’ll tell you why. It was Armed Forces Day. BILL: Yes there were... JOE: ...planes flying over head. BILL: C-47s flying overhead.... JOE: The Armed Forces Day is in May, and that’s one way to remember it anyway.... BILL: That may be the day those pictures were taken from the airplanes. I remember there were people leaning out the doors on the big planes that went overhead. Saturday, May 17, 1952, was, in fact, Armed Forces Day, and while there was disagreement as to the day the first tent was put up, some memories of the specific events could be confirmed, adding to the context of the event. Later in the same interview, Bill remembered an episode from his notes that could represent the tent before Locus 5. When dated and undated notes and pho-

tographs were assembled, many of the discrepancies in the interview were accounted for. In a retrospective account written in about June 1952, a month after the excavation of Locus 2, Bill Eldridge wrote: The next weekend continued to be foul weather and in view of that Nick and I brought a large tarp and several lengths of old line and arriving earlier than the rest we set up the tent in the rain over our holes of the previous week [Locus 2]. By 9:00 the weather cleared and others arrived to add their trowel fulls to the already mountainous piles of back dirt. As our holes broadened many articles came to light, a knife, several nice scrapers and tools with the little points. But by far the most important find in that area was a group of artifacts uncovered by Nick [BBR 299–300]. The discovery of Nick’s tool group in Locus 2 was a momentous occasion. The close association of tools so impressed the excavators that the tools were left in situ while Joe drove 26 mi to get his Brownie camera (Figure 5.2). The discovery

109

Robinson and Ort

i­ nfluenced the decision to keep artifacts separate by locus. As Nick Vaccaro explained in an interview on March 13, 2004, to see “three beautiful things lay on that day. . . . [w]e all thought it was fabulous, and we like to keep it that way.” Joe’s memory was correct about the clear day, but the combined records indicate that he arrived after the weather had cleared, and the dismantled tent caught in the background of his photograph had been built earlier the same day. This association was confirmed later when eight negatives (127 film) of this event were found and pieced together into their original order. The negatives were still in the developer’s envelope, dated May 18, 1952. It is now known with considerable confidence that the “first tent site” was a brief episode related to Locus 2 on the morning of May 17, while the “second tent site” was discovered the following weekend, on May 24. The memory of the first tent site faded, and Locus 5 became the tent site, after which the name was never replaced by the tents that followed. While correlations of memories and records were often straightforward and confidence building, the errors are instructive for evaluating the record and for the depth of caution required to avoid compounding errors during later analysis. Following is an account of the Little Pine and the Juniper Bush. Most of the large loci are referred to multiple times in the field notes, which is one of the criteria used for evaluating them. Some smaller loci received numbers but had few artifacts or ambiguous records. Among these, Locus 42 has not been clearly identified in the written field notes from the time of the excavations. It is included on the 1959 site plan and is described by Jordan in his dissertation as follows:

2451). Fowler was then secretary of the Bronson Museum and a noted author and artist for the Bulletin of the Massachusetts Archaeological Society. In an interview with Joe Vaccaro and Bill Eldridge on November 14, 2002, Joe identified a photograph of himself (BBR 911; Figure 5.3) as having been taken the day Fowler was there, near Locus 5. The photograph included an artifact that Fowler called a “meat pick.” Doug Jordan’s chalkboard is in the photograph, dated November 22, 1953. The meat pick was illustrated in situ by Fowler (1972:Figure 2, see also Figure 3, artifact 24) in the Bulletin, clearly representing the same event. Later (about April 23, 2003), Joe Vaccaro and Bill Eldridge identified a color slide as “Joe Vaccaro digging at the location where he and Bill Fowler dug.” This color slide (BBR 1103, frame 15 of the first roll of slides taken by Bill on his new Balda Baldinette camera) had a “little pine” in the foreground that resembles a Scotch pine. This little pine was confidently identified in a 1952 aerial photograph, 15 m southeast of the center of Locus 5, providing a close and apparently confident location for Locus 42, consistent with that on the 1959 plan. All of this seemed to provide ­multiple lines of supporting evidence that the artifacts in question were from Locus 42, excavated in part on November 22, 1953. However, the “meat pick” was not among the artifacts attributed to Locus 42 but was, rather, cataloged in the PEM catalog and by Douglas Jordan as from Locus 41 (Jordan’s #41JV 6084; PEM 8356). Locus 41 is located about 27 m southwest of Locus 42. Two other artifacts in the same artifact lot attributed to Locus 41 were also illustrated in Fowler’s 1972 article. Important as a cautionary note, Robinson considered the association of the meat pick with Locus 42 to be quite secure and suspected that Jordan had made an error cataloging the artifacts from November 22, 1953, as from Locus 41. Research showed that Nick Vaccaro’s original label from Locus 41 recorded the location as “near little pine tree,” and a photograph of Nick at Locus 41 shows him beside a small jack pine or pitch pine. The jack pine of Locus 41 is also shown across the road behind Locus 11 in another photograph taken by Jordan on November 22, 1953 (Figure 5.4). It appeared that there may have been confusion between two “little pines.” This problem was raised at an interview with

Locality 42 was a small concentration in the northwest sector, a few yards to the south of the “tent site” (number 5). A small amount of material was recovered here by Frank and Nick Vaccaro in the fall of 1954 [1960:130]. The only artifacts cataloged for Locus 42 (Peabody Essex Museum [PEM] 8016 to 8052) were found by Joe Vaccaro, rather than Frank or Nick. Joe’s artifacts were in envelopes labeled “near small pine tree Nov. 22, 1953,” with a later addition in black marker, “Bill Fowler’s Dig” (BBR 110

Spatial Organization at Bull Brook

Figure 5.3. Joe Vaccaro pointing to what William Fowler called the “meat pick” attributed to Locus 41. The tool is shown at right (PEM 8356, 41JV 6084). Photograph taken by Douglas Jordan on November 22, 1953. (Photograph [BBR 911] courtesy of the Peabody Essex Museum.)

Bill Eldridge and Joe Vacarro (August 11, 2006) specifically addressing the problem of two loci labeled as “the little pine.” The visit by William Fowler was clearly remembered by Bill and Joe, and early in the interview there was some agreement that this occurred near the “Scotch pine” (Locus 42) to the south of Locus 5. However, when the problem of the two little pines was discussed relative to other landmarks on the site, Joe Vaccaro became quite adamant that the episode with Fowler occurred at the little jack pine at Locus 41. Robinson suspected that the interview would show that Jordan had erred in his catalog, but after a long discussion (filmed by Melanie Tossell), it was concluded that Jordan was correct and that Joe Vaccaro and Fowler had actually excavated Locus 41, a locus previously discovered by Nick Vaccaro. This was confirmed on March 7, 2007, when fire-spalled biface fragments found by Nick Vaccaro at Locus 41 (41NV2459; PEM 5704) were refit to a fire-spalled biface fragment found by Joe Vaccaro (41JV6085; PEM 8354) from the same series of artifacts that included those illustrated by William Fowler from November 22, 1953. The fact that all the refit artifacts are fire broken suggests that they came from the same locus. Joe’s original field packets dated “Nov. 22, 1953”

still exist (BBR 2451 and 2451), and it is apparent that Doug Jordan marked them “41” in 1959 but that the “41” was altered to “42” when the artifacts were cataloged in about 1980. Among materials previously attributed to Locus 42 is an important sample of calcined bone (PEM 8051). With recognition of this error, there are no artifacts attributed to Locus 42, and without further documentation Locus 42 was not included in the final list of 36 loci. The “little pine” episode contributed to resolving other inconsistencies, including the location of “the juniper bush” at the edge of the bulldozer cut. There are two prominent bulldozer cuts (among many others) that occurred parallel to each other in 1952. The more prominent one, also called the grassy edge, is here labeled bulldozer cut #1 (Figure 5.5). This cut is prominent in many early field records and sketch plans. Bulldozer cut #2 is visible in the 1952 aerial photograph as a U-shaped slash north of the truck road (Figure 5.5). It was remembered during interviews in the 1990s, but much less clearly than the first one. In recent interviews Bill referred to a number of locations near Locus 42 and Locus 6 as being near a juniper bush on the edge of a bulldozer cut. After extensive search of records and p ­ hotographs 111

Robinson and Ort

Figure 5.4. Bill Eldridge (front center) digging in Locus 11 (A), looking across the truck road at Locus 41 (B to C), with “the little pine” directly above Howard Jones’s left shoulder (D). Joe Vaccaro’s location in Figure 5.3 is thought to be near the letter C. Photograph taken by Douglas Jordan on November 22, 1953. (Photograph [BBR 905.1] courtesy of the Peabody Essex Museum.)

the only juniper bush identified directly on the Bull Brook site was that associated with Locus 7 (excavated by Frank Vaccaro) adjacent to Locus 41. Although subject to continued scrutiny, Robinson’s present interpretation is that two little pines occurred near two parallel bulldozer cuts and after 25 years the two bulldozer cuts became merged in memory, along with the little pines and the juniper bush, at the more prominent bulldozer cut #1. Records suggest that the memories may have become merged in about 1980, in which case they were already “old memories” by the time they were recorded in the last decade. While it may seem that this “thick description” of reconstructed excavation contexts is ephemeral to the activities of Paleoindian hunters and gatherers, we cannot overstate the degree to which the ability to scrutinize aspects of spatial relations at the Bull Brook site is dependent on it. We mostly lack field grid coordinates, and we are therefore dependent on more fleeting proxies of space and time that allow reconstruction of GIS coordinates, which are in turn only as good as the as-

sumptions they are built upon. With sufficient records these assumptions are replicable, and there are many avenues of refinement and evaluation that can still be pursued. It is equally important to understand potential errors of memory, especially when reconciling many hours of recent interviews with field records and photographs. These vignettes also serve as cautionary notes for those of us who were not there, as it is tempting to collapse fragmentary records into a coherent solution, much like similar events may become compressed in memory. The extensive body of field notes and photographs provides a means to cross-check memories over successive decades, increasing confidence in most of the loci but also identifying inconsistencies that require greater scrutiny. Although other provenience problems were found, Locus 42 is the only one identified in which conflated memories may have contributed to the creation of a separate locus. It should be obvious from the preceding examples that while it was possible to improve upon mapping data and provenience records, 112

Spatial Organization at Bull Brook

Figure 5.5. Mapping locations superimposed on aerial photograph taken on August 6, 1952. Triangles = loci mapped directly by Eldridge and others about January 1953; pentagons = loci mapped by triangulation and landmarks; circle = hypothetical location of Locus 42; white outlines = trees, including (A) the little pine near Locus 42, (B) a small deciduous tree near Locus 9, (C) the little pine near Locus 41, and (D) the juniper bush near Locus 7. Note that the trees are located just south of their black shadows as determined with a mirror stereoscope.

there are almost certainly provenience errors left that may be impossible to resolve. This is added to the fact that approximately 30 percent of the collection remains unprovenienced. Despite these problems, the provenienced sample of artifacts from Bull Brook is very large, and the research problem remains the same. If the well-defined settlement plan represents an organized event, we may find activities that are structured relative to the circular pattern as a whole, with the

assumption that provenience errors would likely blur, rather than create, a structured plan. Artifact Distributions and the Ring-Shaped Structure

The key to the analysis of artifact distributions is that the artifacts were stored and recorded as artifact lots that could be evaluated separately based on a number of lines of evidence. Although single artifacts sometimes presented clues 113

Robinson and Ort

terior, prior to quantification of the artifacts (Figure 5.6), with the exception of one locus (#14) that was considered ambiguous but on regular spatial grounds was included with the set of eight interior loci. Ort completed the cataloging of all artifacts that could be provenienced before quantifying frequencies of artifacts for each locus. The process was quite literally one in which years of research culminated in a final master sort, with partial results recorded in Figure 5.7 and Tables 5.1–5.2. Statistical analysis showed a significant relationship with flakeshavers, drills, and bifaces concentrated on the interior, while ­endscrapers are more concentrated on the exterior (Table 5.1). Weaker correlations are also shown with ­gravers and wedges, which are part of an ongoing analysis, while sidescrapers are evenly distributed between the two areas. The concentrations of flakeshavers (also called limaces [Grimes and Grimes 1985]) and drills are particularly strong, given that 54 percent and 70 percent of these artifacts, respectively, occurred in 22 percent of the loci. Calculated in terms of the average number of tools in each locus, this means that there are, on average, 6.3 and 8 times more flakeshavers and drills on interior loci (Table 5.2, top). Even more extreme, Ort’s analysis was the first to systematically isolate channel flakes from the 36,000 provenienced flakes, with the result that 84 percent of channel flakes occurred on the interior. On average there were 19 times more channel flakes in interior loci. These figures are based entirely on the spatial analysis, contrasting interior and exterior loci. Although interior and exterior assemblages were statistically different to a very high degree, this does not necessarily mean that interior and exterior samples are significant categories as a whole. If one locus produced a large majority of a given tool type, for example, then the larger group to which that locus belonged may also have a statistical dominance of that tool type, even if there was no particular relationship with the spatial groups. Given the statistical association of the respective tool types, the loci were then sorted to identify the spatial relationship of the respective tool groups, in which the “biface-dominated group” (defined as the total number of bifaces, flake­shavers, and drills) was contrasted with the number of

Figure 5.6. Bull Brook activity distributions distinguishing loci dominated by endscrapers from those defined as the biface group, including biface fragments, drills, and flakeshavers. (Modified from ­Robinson et al. 2009.)

to p ­ rovenience evaluation, the lots were largely fixed, and the evaluation of provenience occurred largely independent of the artifact content. The potential for differences between interior and exterior loci was suggested because Eldridge recorded three loci (15, 21, and 34) that contained an unusual number of drills, and it was observed that these loci all occurred toward the interior of the circle in the original site plan. However, this distinction was never made by the excavators, and we had no basis for comparison for any other artifact forms. Robinson evaluated the provenience of artifact lots, most of which did not change from those of either the Peabody Essex Museum or Doug Jordan’s catalogs, while Jennifer Ort compiled the catalog of artifact classes in lots that were fixed. While our prediction was that some structure may be discernible in artifact distributions if an organized event was represented, we had few preconceptions of what that structure might be beyond the observations of drill concentrations in three loci. With final mapping, Robinson designated the set of loci that were most situated toward the in114

Table 5.1. Bull Brook Artifact Frequencies Separated by Interior and Exterior Loci.

Interior Loci

Exterior Loci % of Site Total

Count

22

28

Artifacts Included in Statistical Analysis Flakeshaver 81 8.46 Drill 48 7.12 Biface 127 6.31 Endscraper 203 –5.03 Wedge 75 –3.12 Graver 32 –2.42 Sidescraper 93 .24 Total 659

64 70 45 18 18 17 27 26

45 21 155 910 339 157 257 1,884

Flakes and Channel Flakes Flakes 17,169 Channel flakes 103

47 84

19,306 19

Bull Brook Category

Loci Included

Count

Z-score

8

Z-score

–5.00 –4.21 –3.73 2.97 1.84 1.43 –.14

Six Biface Dominated Loci

Site Total Count

Count

% of Site

36

6

17

126 69 282 1,113 414 189 350 2,543

79 44 97 61 40 15 59 395

63 64 34 5 10 8 17 16

36,475 122

16,807 108

46 89

Note: Z-scores show negative and positive correlations for each set, with absolute values of Z > 2.58 significant at .01. Source: Reproduced from Robinson et al. 2009.

Table 5.2. Relative Artifact Frequencies Comparing Interior to Exterior Loci and Between the Biface Group and the Endscraper-Dominated Loci.

Interior (8 Loci) Artifact Class

Flakeshaver Drill Biface Channel flake Endscraper Wedge Graver Sidescraper

No. of Tools

81 48 127 103 203 75 32 93

Tools/Loci

10.1 6 15.9 12.9 25.4 9.4 4 11.6

Exterior (28 Loci) No. of Tools

45 21 155 19 910 339 157 257

Tools/Loci

1.6 .8 5.5 .7 32.5 12.1 5.6 9.2

Total Tools

126 69 282 122 1,113 414 189 350

Ratio of Interior to Exterior

6.3 8 2.9 19 .8 .8 .7 1.3

Note: Artifact proportions between interior and exterior loci.

Biface Dominated 6 Loci Artifact Class

Flakeshaver Drill Biface Channel flake Endscraper Wedge Graver Sidescraper

No. of Tools

79 44 97 108 61 40 15 59

Tools/Loci

13.2 7.3 16.2 18 10.2 6.7 2.5 9.8

Endscraper Dominated 30 Loci No. of Tools

47 25 185 14 1,052 374 174 291

Note: Artifact proportions between biface group and endscraper-dominated loci.

Tools/Loci

1.6 .8 6.2 .5 35.1 12.5 5.8 9.7

Total Tools

126 69 282 122 1,113 414 189 350

Ratio of Biface to Endscraper Dominated Loci

8.4 8.8 2.6 38.6 .3 .5 .4 1

Robinson and Ort

Figure 5.7. Selected artifact frequencies that most distinguished interior from exterior loci at the Bull Brook site. The graph shows the gradual decrease in the percentage of endscrapers, followed by the relatively sudden increase and dominance of artifacts assigned to the biface group. (Adapted from Robinson et al. 2009.)

e­ ndscrapers. Sorted by the relative proportions of artifacts in these groups, 30 loci were dominated by endscrapers in gradually decreasing proportions, until a relatively abrupt change in which six loci were dominated by the biface group of artifacts (Figure 5.7). Although the respective tool groups were originally identified on the basis of interior and exterior spatial groups, this alternate means of sorting more accurately identifies the spatial relationship of loci that most strongly represent the respective tool groups and their activities. Five out of six loci dominated by the biface group of artifacts are among those identified as the interior group (Figure 5.6). Sorted in this manner, the relative contrasts between the artifact types is increased, with the most extreme contrast in channel flakes, with 89 percent of channel flakes occurring in the six biface-dominated loci. Channel flakes were on av-

erage 38 times more abundant in these loci (Table 5.2, bottom), even though this artifact class was not among those used in identifying the groups. It may be interpreted from this, in part, that the significant contrast between the two groups of activities is not the presence or absence of bifaces but, rather, the activity of fluting and the finishing aspects of fluted point production. In this scenario the channel flakes are left behind with the remains of other specialized production activities, including drills and flakeshavers, while the finished fluted points are more widely dispersed among the occupation loci. For example, Locus 25 (one of the exterior loci) yielded the largest number of complete fluted points (seven) from any locus but only one drill and no flakeshavers or channel flakes, and the assemblage as a whole is dominated by endscrapers. Thus it is proposed that the interior and exterior loci are character116

Spatial Organization at Bull Brook

ized by different kinds of production (such as hunting weapons and skin working) but not necessarily the distribution of end products.

atively hunting migrating caribou. At the same time, there “is such a diversity of ethnographically known peoples with caribou-dependent seasonal adaptations in their seasonal rounds that we canInterpretation and Implications not realistically say that caribou hunting tells us The concentric activity patterning at Bull Brook anything about the rest of the year’s adaptations” provides a platform from which to investigate (Spiess 1979:131). multiple lines of research. The veracity of the conLithic source analysis has been a primary focus centric activity patterning is supported by simi- of Northeast Paleoindian studies (Burke 2006; lar patterns at two other large Paleoindian sites. Curran and Grimes 1989; Ellis and Lothrop 1989; At Parkhill in Ontario, a partly excavated line of Pollock et al. 1999; Spiess et al. 1998). Ongoing four biface-dominated loci was situated in front studies of lithics at Bull Brook by Adrian Burke (on the terrace edge) of a larger number of more (2006; Robinson et al. 2009:426) and others (Poldiverse assemblages situated behind them (land- lock et al. 1999; Pollock et al. 2008) have provided ward [Ellis and Deller 2000:226, 246]). At Debert confident identifications of materials including in Nova Scotia the incidence of projectile points, Normanskill/Mount Merino Formation chert drills, and chipping debris increased in a smaller from New York, Munsungun Formation chert number of loci downslope, while skin-working from Maine, Hardyston Formation jasper from tools increased upslope in a higher number of loci Pennsylvania, and spherulitic rhyolite from New (MacDonald 1968:133). Production and ­gender-​ Hampshire, with a minority presence of local rhyspecific activities were suggested among possible olites. Ninety-five percent of Bull Brook materials correlations at both sites. The spatial separation of come from greater than 250 km away (Robinson different social contexts may be amplified at large et al. 2009:427). social gatherings (Slobodin 1962:61–62; Whitelaw Chronology remains a problem for northeast1991:158), and it is possible that some activities ern Paleoindian studies. While the Bull Brook and tool forms may be more prominent at large phase (Curran 1999; Grimes et al. 1984) is most gatherings, beyond the direct effect of increased often associated with the Gainey phase (Curran sample size. Large numbers of drills, for example, 1999:​7; Ellis and Deller 1997:10; Spiess et al. 1998) are found concentrated in a few loci at Bull Brook of the Great Lakes region, firm dates are scarce, and at the Vail site in Maine (Gramly 1982) and let alone the time ranges for these periods. The are otherwise comparatively rare at smaller-scale Shawnee-Minisink site in Pennsylvania provides settlements. the earliest reliable dates in the Northeast, with an Communal hunting for caribou is among the average radiocarbon age of 10,935 ± ​15 Bp (­Bradley traditional explanations for large Paleoindian et al. 2008:124; Gingerich 2007:120, and Chapter sites in the Northeast (Ellis and Deller 2000:241; 9, this volume). With recent accelerator mass MacDonald 1968:116–120; Spiess 1984:282), al- spectrometry dates on charcoal from Bull Brook though a variety of more generalized or alterna- it was proposed that virtually all prior attempts tive subsistence and settlement patterns have been to date the site with charcoal represent Holoproposed (Dincauze 1993; Dincauze and Jacob­ cene forest fires (Robinson et al. 2009:425). Two son 2001). Given the occurrence of caribou bone new dates on calcined bone from samples associin multiple loci at Bull Brook (Spiess et al. 1998) ated with caribou bone yielded dates of 10,410 ± ​ and the proximity of a possible summering loca- 60 Bp (Beta-240629, 10,700–10,100 cal Bc, 2σ) tion for caribou on a now submerged maritime and 10,380 ± ​60 Bp (Beta-240630, 10,670–10,040 island (Pelletier and Robinson 2005; Robinson cal Bc, 2σ). The dates were run by Beta Analytic et al. 2009),we find communal caribou hunting using the newly developed pretreatments for cala likely subsistence model for Bull Brook, s­ ubject cined bone (Lanting et al. 2001). The closeness to further testing. The site is proposed to be an of the dates gives confidence to their precision, unusually large and well-organized aggregation, but given that they employ a relatively new dating consistent with the demanding subsistence, so- method and that previous studies have warned cial, and ritual practices associated with cooper- of the possibility of contamination by absorbed 117

Robinson and Ort

carbonate (Ambrose and Krigbaum 2003:195), it remains uncertain how closely they would agree with charcoal dates. The dates may be too young based on prior expectations, or they may be ­accurate, ­suggesting that Bull Brook is as much as 500 years younger than Shawnee-Minisink. Bradley et al. (2008:​136) accept the later attribution, defining the Bull Brook–West Athens Hill point style. The two dates on calcined bone are at present the only potentially reliable radiocarbon dates from Bull Brook (Robinson et al. 2009:426). These and other important issues are among the many avenues of research that need continued scrutiny and which may allow Bull Brook to inform us on the next round of issues bearing on fundamental interpretations of culture and social organization in the Pleistocene. How widespread are Paleoindian cultural patterns beyond the mobile technologies that are most visible in the archaeological record? In the search for origins and understanding of a broad array of social and ecological transformations, underlying cultural principles and organization are as important as material culture itself (Storck 1988:249). With the extraordinary persistence of the excavators and associated museums at keeping the Bull Brook loci intact, we consider that the basic pattern of concentric artifact patterning is strongly expressed, supporting the ­interpretation of Bull Brook as an organized event and setting a foundation for added layers of inquiry. We strongly suspect that if internal spatial ­distributions within

loci were available and artifact proportions within each locus were complete and e­ rror-​free, the concentric activity patterning would be even more strongly expressed and added layers of significantly patterned variability would also be more confidently identified. We emphasize that with the copious notes, the photographs, and the insightful observations of the original excavators and investigators, much of the spatial information gleaned from Bull Brook is replicable and testable. There is the potential for further refinement as needed. Here we have elaborated on the methods employed to critically evaluate the spatial pattern of loci and artifacts. Having had the opportunity to work intensively with the o ­ riginal excavators, we emphasize their integrity and dedication toward the goal of making Bull Brook “reasonably well understood.” There are clear limitations to how much the data can be resolved, but given the large sample sizes and unequaled spatial organization of Bull Brook, this early excavation of a large northeastern Paleoindian site has much more to offer. In a retrospective account dated to about 1960 Bill Eldridge succinctly summarized their effort: Fortunately early in the work, the material was found to be concentrated in small isolated areas or house sites. These we called hotspots, mapped them carefully as they were discovered and kept the material from each one separated. This forethought will be the key to learning what happened at Bull Brook [BBR 1956].

Acknowledgments Research on Bull Brook spans 60 years and generations of researchers. The Bull Brook excavators and their families made it happen with their dedication, persistence, and creativity. The present effort was greatly dependent on their continued efforts. Bill Eldridge dedicated much of his life to “learning what happened at Bull Brook,” and he would be the first to acknowledge the critical roles played by so many people and institutions. The Robert S. Peabody Museum and the Peabody Essex Museum have been major supporters of the research, and more recently the University of Maine, among others. Long support from and collaboration with Frederick H. West and funding by the National Science Foundation (Grant No. BCS 0352918) made the most recent research possible, with research collaborators on that project including Adrian Burke, Marylou

Curran, William Eldridge, Douglas Jordan, and Joseph Kelley, in addition to students and staff of the respective institutions. Since this chapter focuses on interviews, we are particularly grateful to John Grimes, William Eldridge, and Melanie Tossell, who arranged interviews spanning decades. We appreciate Joe Gingerich’s invitation to participate in this volume. The manuscript was read by Joe Gingerich, Ann Surprenant, and two reviewers. Opinions and errors are our responsibility.

Note 1. Primary source materials are from the original Bull Brook records, stamped with BBR catalog numbers by B. Robinson, archived at the Peabody Essex Museum, Salem, Massachusetts. 118

Spatial Organization at Bull Brook

References Cited

The Bull Brook Phase, Antecedents, and Descendants. In The Archaeological Northeast, edited by Mary Ann Levine, Kenneth E. Sassaman, and Michael S. Nassaney, pp. 3–24. Bergin and Garvey, Westport. Curran, Mary Lou, and John R. Grimes 1989 Ecological Implications for Paleoindian Lithic Procurement Economy in New England. In Eastern Paleoindian Lithic Resource Use, edited by Christopher J. Ellis and Jonathan C. Lothrop, pp. 41–74. Westview Press, Boulder. Dincauze, Dena F. 1993 Pioneering in the Pleistocene: Large Paleoindian Sites in the Northeast. In Archaeology of Eastern North America: Papers in Honor of Stephen Williams, edited by James B. Stoltman, pp. 43–60. Archaeological Report No. 25. Mississippi Department of Archives and History, Jackson. Dincauze, Dena F., and Victoria Jacobson 2001 The Birds of Summer: Lakeside Routes into Late Pleistocene New England. Canadian Journal of Archaeology 25:121–126. Eldridge, William, and Joseph Vaccaro 1952 The Bull Brook Site, Ipswich, Mass. ­Bulletin of the Massachusetts Archaeology Society 13(4):​39–43. Ellis, Christopher J., and D. Brian Deller 1997 Variability in the Archaeological Record of Northeastern Early Paleoindians: A View from Southern Ontario. Archaeology of Eastern North America 25:1–30. 2000 An Early Paleoindian Site near Parkhill, Ontario. Mercury Series. Archaeological Survey of Canada Paper 159. Canadian Museum of Civilization, Hull. Ellis, Christopher J., and Jonathan C. Lothrop (­editors) 1989 Eastern Paleoindian Lithic Resource Use. Westview Press, Boulder. Fowler, William S. 1972 Bull Brook: A Paleo Complex Site. Bulletin of the Massachusetts Archaeological Society 34(1–2):1–6. Gingerich, Joseph A. 2007 Picking Up the Pieces: New Paleoindian Research in the Upper Delaware Valley. Archaeology of Eastern North America 35:​­117–124. Gramly, Richard M. 1982 The Vail Site: A Palaeo-Indian Encampment in Maine. Bulletin of the Buffalo Society of Natural Sciences, Vol. 30. Buffalo. Grimes, John R. 1979 A New Look at Bull Brook. Anthropology 3:109–130.

Ambrose, Stanley H., and John Krigbaum 2003 Bone Chemistry and Bioarchaeology. Journal of Anthropological Archaeology 22(3):​­193–199. Anderson, David G. 1995 Paleoindian Interaction Networks in the Eastern Woodlands. In Native American Interactions: Multiscalar Analysis and Interpretations in the Eastern Woodlands, edited by Michael S. Nassaney and Kenneth E. Sassaman, pp. 3–26. University of Tennessee Press, Knoxville. Boldurian, Anthony T., and John L. Cotter 1999 Clovis Revisited: New Perspectives on Paleoindian Adaptations from Blackwater Draw, New Mexico. University Museum, University of Pennsylvania, Philadelphia. Bradley, James W., Arthur E. Spiess, Richard A. Boisvert, and Jeff Boudreau 2008 What’s the Point? Modal Forms and Attributes of Paleoindian Bifaces in the New E ­ ngland–​Maritimes Region. Archaeology of Eastern North America 36:119–172. Bullen, Ripley P. 1950 The Johnson’s Spring Site. Bulletin of the Massachusetts Archaeological Society 2(2):​­37–44. Burke, Adrian L. 2006 Paleoindian Ranges in Northeastern North America Based on Lithic Raw Materials Sourcing. In Notions de territoire et de mobilité: Exemples de l’Europe et des premières nations en Amérique du Nord avant le contact européen, ERAUL 116, edited by C. Bressy, A. Burke, P. Chalard, and H. Martin, pp. 77– 89. Études et Recherches Archéologiques de l’Université de Liège, Liège. Byers, Douglas A. 1954 Bull Brook — ​A Fluted Point Site in Ipswich, Massachusetts. American Antiquity 19(4):​ 343–351. 1955 Additional Information on the Bull Brook Site, Massachusetts. American Antiquity 20(3):274–276. Cotter, John L. 1937 The Occurrence of Flints and Extinct Animals in Pluvial Deposits near Clovis, New Mexica, Pt. IV: Report on the Excavations at the Gravel Pit in 1936. Proceedings of the Philadelphia Academy of Natural Sciences 89:1–16. Curran, Mary Lou 1984 The Whipple Site and Paleoindian Tool Assemblage Variation: A Comparison of Intra­ site Structuring. Archaeology of Eastern North America 12:5–40. 1999 Exploration, Colonization, and Settling In: 119

Robinson and Ort Grimes, John R., William Eldridge, Beth G. Grimes, Antonio Vaccaro, Frank Vaccaro, Joseph Vaccaro, Nicolas Vaccaro, and Antonio Orsini 1984 Bull Brook II. Archaeology of Eastern North America 12:159–183. Grimes, John R., and Beth G. Grimes 1985 Flakeshavers: Morphometric, Functional and Life-Cycle Analysis of a Paleoindian Unifacial Tool Class. Archaeology of Eastern North America 13:35–57. Jordan, Douglas F. 1960 The Bull Brook Site in Relation to “Fluted Point” Manifestations in Eastern North America. Unpublished Ph.D. dissertation, Harvard University, Cambridge. Lanting, J. N., A. T. Aerts-Bijma, and J. van der Plicht 2001 Dating Cremated Bone. Radiocarbon 43(2):​ 249–254. MacDonald, George F. 1968 Debert: A Paleo-Indian Site in Central Nova Scotia. National Museum of Canada Anthropological Papers 16. Ottawa. MacNeish, Richard S. 1979 Douglas Swain Byers 1903–1978. American Antiquity 44(4):708–710. Meltzer, David J. 2006 Folsom: New Archaeological Investigations of a Classic Paleoindian Bison Kill. University of California Press, Berkeley. Pelletier, Betrand G., and Brian S. Robinson 2005 Tundra, Ice and a Pleistocene Cape on the Gulf of Maine: A Case of Paleoindian Transhumance. Archaeology of Eastern North America 33:163–176. Pollock, Stephen G., Nathan D. Hamilton, and Robert A. Boisvert 2008 Archaeological Geology of Two FlowBanded Spherulitic Rhyolites in New England, USA: Their History, Exploitation and Criteria for Recognition. Journal of Archaeological Science 35(3):688–703. Pollock, Stephen G., Nathan D. Hamilton, and ­Robson Bonnichsen 1999 Chert from the Munsungun Lake Formation (Maine) in Palaeoamerican Archaeological Sites in Northeastern North America: Recognition of Its Occurrence and Distribution. Journal of Archaeological Science 26:269–293. Ritchie, William A Probable Paleo-Indian Site in Vermont. 1953 American Antiquity 3:249–258. Robinson, Brian S. 2003 Testing for Paleoindian Aggregations: Internal Site Structure at Bull Brook. Grant

proposal submitted to the National Science Foundation, No. BCS 0352918. Robinson, Brian S., and William E. Eldridge 2005 Debating Bull Brook, 1965–1972. Bulletin of the Massachusetts Archaeological Society 66(2):​67–75. Robinson, Brian S., Jennifer C. Ort, William E. Eldridge, Adrian L. Burke, and Bertrand G. Pelletier 2009 Paleoindian Aggregation and Social Context at Bull Brook. American Antiquity 74(3):423– 447. Slobodin, Richard 1962 Band Organization of the Peel River Kutchin. National Museum of Canada Bulletin No. 179. Ottawa. Spiess, Arthur E. 1979 Reindeer and Caribou Hunters: An Archaeological Study. Academic Press, New York. 1984 Arctic Garbage and New England Paleo-­ Indians: The Single Occupation Option. ­Archaeology of Eastern North America 12:​ 280–285. Spiess, Arthur E., Deborah B. Wilson, and James W. Bradley 1998 Paleoindian Occupation in the New ­England–​Maritimes Region: Beyond Cultural Ecology. Archaeology of Eastern North America 26:201–264. Storck, Peter L 1988 The Early Palaeo-Indian Occupation of Ontario: Colonization or Diffusion? Bulletin of the Buffalo Society of Natural Sciences 33:243– 250. 2004 Journey to the Ice Age: Discovering an Ancient World. Royal Ontario Museum, Ontario. Time-Life 1992 The First Americans. American Indians. Time-Life Inc., Alexandria, Virginia. Trigger, Bruce G. 2006 A History of Archaeological Thought. 2nd ed. Cambridge University Press, Cambridge. Whitelaw, Todd 1991 Some Dimensions of Variability in the Social Organisation of Community Space Among Foragers. In Ethnoarchaeological Approaches to Mobile Campsites: Hunter-Gatherer and Pastoralist Case Studies, edited by Clive ­Gamble and William A. Boismer, pp. 139– 188. International Monographs in Prehistory, Ann Arbor. Wobst, H. Martin 1974 Boundary Conditions for Paleolithic Social Systems: A Simulation Approach. American Antiquity 39(2):147–178. 120

6

Fifty Years of Discovery at Plenge: Rethinking the Importance of New Jersey’s Largest Paleoindian Site Joseph A. M. Gingerich

The Plenge site (28WA636) is the largest known Paleoindian site in New Jersey. Located along the Musconetcong River in northwestern New Jersey, the site covers over 22 ac. Nearly every known Paleoindian point type in eastern North America has been recorded at this site. The last extensive study of Plenge was in the early 1970s, when Herbert Kraft (1973) conducted limited excavations and a preliminary analysis of the lithic assemblage. This chapter reports on over three decades of newly collected material and provides a discussion of how Plenge may enhance our understanding of regional patterns of Paleoindian land use.

plow zone between 9 and 11 inches in thickness that overlies sterile subsoil. These interpretations suggest that there are no intact archaeological deposits at the site (Kraft 1973:61). Despite the disturbed contexts from which artifacts were found, Kraft (1973:61) felt confident in assigning many of the artifacts of “superior quality” or made from high-quality jaspers and other exotic raw materials to the Paleoindian occupation. Based on these criteria, Kraft reported that there had been 1,017 diagnostic Paleoindian artifacts collected from the site, including 117 fluted points, knives, or preforms; 7 Plano-style points; 218 endscrapers; 5 limace or slug-like scrapers; 10 spokeshaves; 28 gravers; and 23 channel flakes. If all of these artifacts are indeed from the Paleoindian occupation as Kraft asserts, these counts rival the largest Paleoindian sites known in eastern North America. The occurrence of multiple fluted point types, which are believed to span the entire Paleoindian period, is rare in the east. No other site in eastern North America equals this diversity or temporal span in Paleoindian point types. The Plenge site, as noted by Kraft (1973:112), shows the reuse of this single locality throughout prehistory. According to Kraft (1973:112), prehistoric people may have been attracted to the site for a number of factors: raw material availability, strategic point of caribou migration, ­proximity

Summary of Original Investigations

In the 1950s surface collection of the site began as a number of avocational archaeologists surveyed the plowed fields for artifacts, which spanned the Archaic and Woodland time periods. In the 1960s, the first fluted points were found by the Ziegler, Staats, and Stanley families. Realizing the importance of these finds, Leonard Ziegler and F. Dayton Staats contacted Herbert Kraft (1973:57) of Seton Hall University, who began investigations of New Jersey’s first known extensive Paleoindian occupation site. Kraft provided a brief description of the site’s geologic and geomorphic context. Based on limited field investigations, he (1973) reported that all artifacts were in a shallow 121

Gingerich

to plant or water resources, and warmer seasonal temperatures. Subsequent Analyses

Prior to my study of Plenge, only three other ­studies were conducted with the Paleoindian artifacts after Kraft’s work. These analyses are briefly summarized below. In 1975, Leonard Eisenberg (1978) included Plenge in his study of Paleoindian settlement patterns along the Hudson and Delaware river drainages. He (1978:60) performed a cluster analysis with 100 artifacts from Plenge, r­ esulting in 23 clusters and three major groups. From the analysis of these groups, Eisenberg (1978:64, 132–​ 133) concluded that (1) while Plenge contains many tool forms, analyses of unifaces suggest a more homogeneous artifact inventory than reported by Kraft (1973); (2) many tools exhibit similar edge angles and wear, which suggests a signifi­ cant amount of contact with hard materials; and (3) accessory edge modifications frequently occur on the stone tools at Plenge. Based on these interpretations and a review of the other artifacts reported by Kraft (1973), Eisenberg (1978:133) believes that a fair amount of manufacturing took place at Plenge, with the variety of tools and hunting and butchery implements (i.e., points and knives) being indicative of base camp activities. In the late 1990s, Pollock et al. (1999) ­examined a selection of exotic or unknown raw materials recovered at Plenge. This included fluted point fragments, flake tools, and debitage considered to be a part of the Paleoindian assemblage. In all, 68 artifacts were examined from Plenge. Of these, 46 were “wholly consistent” with two reddish varieties of Munsungun Lake chert, 11 grayish cherts were consistent with three different varieties of Munsungun Lake chert, and three additional black chert artifacts were inconclusive (Pollock et al. 1999:290). This study, which employed optical microscopic analysis, suggests that Paleoindians at Plenge utilized materials from the Munsungun chert quarries in Maine, nearly 800 km distant. Continuing with the investigation of raw material use, Davis (2003) presented a short paper on Paleoindian mobility strategies and raw material use at Plenge. Based on the location of five crypto­ crystalline sources heavily used by Paleoindians

within the Middle Atlantic region (Hardyston, Helderberg, Normanskill, Onondaga, and Shriver formations), a system of ranking was used to plot expected raw material frequencies at Plenge by distance to source (Davis 2003:3). From visible inspection of the fluted points within the Plenge collection, Davis reports that more than half of the raw material utilized in the manufacture of points is derived from nonlocal sources (i.e., not brown jasper). While it is well known that fluted projectile points are often made on nonlocal raw materials, over 40 percent of the points or preforms at Plenge are made of jasper, suggesting a more frequent use of local toolstone and less reliance on nonlocal materials (Davis 2003). Resharpening patterns on fluted points made from nonlocal raw materials also support this theory, as they are often heavily resharpened, showing reduction in width and retouched edge angle (Davis 2003:7). While Davis attempts to develop ways in which to link patterns of material use to strategies of mobility (i.e., logistical vs. residential mobility), perhaps his most important contribution is the recognition that some high-quality but local sources of toolstone are missing from the assemblage at Plenge. Based on macroscopic analyses, Helderberg and Shriver cherts located within 45 km of the site are absent from the assemblage. These patterns may tell us more about landscape and raw material use than speculating about the mobility strategies of people based on the percentages of nonlocal raw materials. Physical Setting

The Plenge site is best viewed as a complex of sites that are spread across 22 ac of floodplain, terrace, and low upland settings along the Musconetcong River in northwestern New Jersey (Gingerich and Stewart 2010). The Musconetcong River flows southwest and eventually empties into the Delaware River, 19 km away (Figure 6.1). At the Plenge site, two spring-fed drainages flow into the Musconetcong (Figure 6.2). The main concentration of Paleoindian artifacts recorded at Plenge occurs on a high west-facing terrace, approximately 8 m (26 ft) above the Musconetcong River (­Figure 6.2). Geologically the site is located on part of the Kittatinny Formation, which is Cambro-­Ordovician in age and contains chert-bearing members (Kraft 1973). Mapping of the H ­ ardyston quartzite for-

122

Fifty Years of Discovery at Plenge

Figure 6.1. Location of Plenge with nearby rivers and Paleoindian sites (see Figure 6.15 for inset).

mation in Pennsylvania, which produces jasper outcroppings, extends into New Jersey within a few miles of Plenge (Davis 2003). In addition to being nearby jasper sources, the site is situated between the terminal moraines of the Wisconsin and Kansan ice sheets, 3 mi north and 6 mi south, respectively (Kraft 1973:61–62, citing Dr. Raymond Murray). These moraines likely provided access to some raw materials suitable for toolmaking. 123

New Research

Although Kraft did little with the Plenge site after his seminal report in 1973, the site continued to be surface-collected. In the summers of 2009 and 2010, I examined four separate collections with Paleoindian material from the Plenge site. One of these collections was unknown or else mostly assembled after Kraft’s study, and the three other collections had distinguished the new finds from those examined by Kraft with different catalog

Figure 6.2. Map of Plenge with artifact concentrations (all measurements in meters).

Fifty Years of Discovery at Plenge Table 6.1. Paleoindian Artifacts Collected from the Plenge Site.

Artifact Type

Kraft 1973

New Artifacts

Total

Fluted points and preforms Channel flakes Plano points and preforms Endscrapers Sidescrapersa Limaces (slug-like scrapers)b Gravers Total

117 23 7 218 330 5 28 728

72 29 4c 50 3 0 0 158

189 52 11 268 333 5 28 886

Includes Kraft’s broad category of sidescrapers and utilized flake scrapers. b Referred to as Hump-back scrapers in the original report. c There may be more present in the collection; some bifaces were too fragmented to be placed in this category. a

numbers. This allowed me to generate new estimates on the number of artifacts that had been collected since the early 1970s. The ­numbers were astonishing — ​72 fluted pieces (points, fragments, and preforms), 50 endscrapers, three side­scrapers, and 29 channel flakes (Table 6.1). In addition, thousands of pieces of debitage and hundreds of utilized and retouched flakes were collected. In revisiting Plenge, I had three goals: (1) to create a better understanding of the Paleoindian material recovered from the site, (2) to further assess the raw material use at Plenge, and (3) to determine whether there was any spatial discreteness to particular Paleoindian surface finds and whether these areas had the potential to contain buried intact deposits. Analysis of the Assemblage

To better characterize the Paleoindian assemblage it was important to try to compare the types that Kraft (1973) defined with our current knowledge of Paleoindian projectile point styles and their hypothesized temporal order. Although Plenge contains a large number of fluted points and preforms, the point bases, early-stage preforms, and fluted point fragments were difficult to place into distinct or known types. But from a technological standpoint, some basic designations could be made by examining fluting preparation, preform or point shape, flute length, presence or absence of co-lateral flaking or midline formation, and depth of basal concavity (Bradley et al. 2008; Bradley et al. 2010; Callahan 1979; Deller and E ­ llis 125

1992; Ellis 2004; Ellis and Deller 1997; Morrow 1996). These designations, however, could only be organized into broad categories of Early, Middle, and Late Paleoindian. As used in this chapter, the basic criteria for these categories are as follows: 1. Early Paleoindian: low to moderate basal concavities, slightly divergent to parallel sides, fluting less than half the length of the total biface length, and flaking that shows less emphasis on the creation of a defined medial ridge. One variation in this category is the inclusion of Vail-Debert points, which adhere to many of the aforementioned attributes but exhibit deep basal concavities. 2. Middle Paleoindian: moderate to deep basal concavities, tendency of more divergent sides, occurrence of bases that are more waisted, creation of well-defined medial ridges, and flutes extending over half the length of the biface. Variation within this category would include Crowfield, which lacks medial ridges, exhibits shorter flutes, and is often extremely thin. 3. Late Paleoindian: includes Plano forms and Dalton-style points, which obviously ­differ greatly but are distinctive from Early and Middle Paleoindian phases in that they lack fluting or fluting preparation. Late Paleoindian forms are also well flaked ­compared with many later forms, with Plano forms often exhibiting co-lateral flaking and ­Dalton-​like forms displaying extensive ­pressure flaking. The fluted points at Plenge show tremendous variation, representing forms that span the entire Paleoindian period (Figure 6.3). Based on my classification scheme, outlined above, the majority of fluted points, preforms, and fragments at Plenge fall within the Middle Paleoindian phase (Figure 6.4). Early forms that could be recognized are considered to be Clovis variants, which might be characterized as Gainey and Kings Road– Whipple (Bradley et al. 2008). Vail-Debert forms, which have been dated at their respective sites to ca. 10,700 14C  bp (Gramly 2009; MacDonald 1968; Figure 6.5), are also present. Judging the fluted points at Plenge on both metrics and technology, few (if any) fit the general Clovis form (see Bradley et al. 2008; Bradley et al. 2010).

Gingerich

Figure 6.3. Examples of the different Paleoindian point styles recovered from the Plenge site.

Middle Paleoindian forms at Plenge include Barnes-style points, which exhibit medial ridges and long flutes (Figure 6.6a–b); Crowfield, which are thin and fluted on one side (Figure 6.6c); and a smaller fluted form referred to by Kraft as a “stubby” fluted point (Figure 6.7). This stubby form is included in the Middle Paleoindian phase as it was defined and described by G ­ ardner (1974) in levels above Clovis and below Dalton phase points at the Thunderbird site (see also Carr et al., Chapter 8, this volume). These small fluted points do not appear to be resharpened larger points or recycled bases/tips but, rather, a distinct type that has been manufactured in this way (see also

Gardner and Verrey 1979). At least three specimens consistent with Crowfield-type points have also been recovered at Plenge and are considered to be Middle Paleoindian (Bradley et al. 2008; Deller and Ellis 1984, 1992). Late Paleoindian forms at Plenge are marked by the presence of Plano-like points, which are relatively rare in the region. No complete Plano forms have been found at Plenge, but based on the bases and fragments identified, these specimens are consistent with definitions for the type (cf. Bradley et al. 2008; Figure 6.8). A second probable Late Paleoindian form identified at Plenge is a pentagonal form that is well flaked and exhibits

126

Fifty Years of Discovery at Plenge

some evidence of basal thinning and grinding but no clear fluting (Figure 6.9). Kraft (1973:83) identified the pentagonal form and also considered this technology to be Late Paleoindian. Raw Material Use and Patterns of Lithic Reduction at Plenge

Another research question when revisiting Plenge involved evaluating Kraft’s (1973:62) hypothesis that the majority of raw material utilized at the site, that is, jasper, was transported from large quarries located 30 to 50 mi distant in Pennsylvania. Davis (2003) has also suggested that a more local source of jasper may have been present within 10 mi of Plenge. Even at this distance, I expect that toolstone would have arrived at Plenge in a reduced blank form, in which case the presence of primary flakes and large cores would be rare at the site.

Figure 6.4. Breakdown of Paleoindian points by phase.

Figure 6.5. Specimens resembling Vail-Debert-style points.

127

Gingerich

Figure 6.6. (a) Middle Paleoindian points, most resembling Barnes-style points. A probable Crowfield base is shown in the upper right corner.

As previously mentioned, the Paleoindian component of the site contains numerous preforms (Figure 6.10). The preforms range in all sizes from Stage 5, at which you can easily recognize Paleoindian forms, to the final fluting stages, which are marked by at least 52 channel or ­later-​stage end thinning flakes at Plenge (Figure 6.11 and Table 6.2).1 Furthermore, of the fluted

pieces that were fully analyzed from the Ziegler and Stanley collection (n = 118), 30 percent were classified as earlier-stage broken preforms, and 32 percent did not exhibit ground bases, suggesting that many of these points may have been abandoned before completion. This indicates that a significant amount of fluted point production took place at Plenge. Because Plenge is a surface

128

Figure 6.6. (b) Middle Paleoindian preform, with nicely developed midline (­obverse and reverse).

Figure 6.6. (c) Complete Crowfield point, flute present on the obverse side.

Figure 6.7. Small or “stubby” Middle Paleoindian points.

Figure 6.8. Plano-like Late Paleoindian forms.

Fifty Years of Discovery at Plenge

site, containing no known intact deposits, fluted point preforms and obvious debris of fluted point manufacture are the only items that can be confidently assigned to the Paleoindian period. Cores, primary flakes, hammerstones, and other materials indicative of lithic reduction cannot be assigned to a specific time period. However, to explore access to raw material and the presence of early stages of tool manufacture at the Plenge site, temporal assignment of cores and debitage is unnecessary. Assuming that only sites near toolstone sources (≤3 km) should contain large cores and primary reduction debris, the presence of these items at Plenge from any time period would suggest a nearby source of stone. The Plenge collection provides an opportunity to evaluate access to lithic raw material. The Plenge collection contains tens of thousands of pieces of debitage. Following Kraft’s investigation of the site in the late 1960s and early 1970s, collectors began picking up all artifacts they encountered in the fields. They often joke that they “­vacuum-​cleaned the site,” as today it can be difficult to locate any artifacts on portions of the site. In surveying the collection of debitage that

Figure 6.9. Pentagonal form assumed to be Late Paleoindian.

Figure 6.10. Examples of fluted preforms from Plenge.

131

Gingerich

Figure 6.11. Channel flakes from Plenge.

had been assembled by the Ziegler family, over 1,000 pieces of jasper that weigh in excess of 50 g have been collected. These pieces mainly consist of angular debris and primary flakes interpreted to be related to early reduction stages. The largest pieces analyzed during this survey included primary flakes with cortex weighing over 200 g. These large flakes exhibited remnant weathering on the dorsal aspect consistent with both cobble and bedrock cortex. As exhausted cores cannot be used as a good measure of the original blank size carried to Plenge, primary flakes in excess of 200 g suggest that cores or blanks greater than this size were present on the site. The presence of these primary flakes within the collection supports the hypothesis that early-stage reduction took place at Plenge and access to jasper was in close proximity. Additionally, in the original analysis of the assemblage, Kraft (1973:111) reports that 11.4 percent of 6,557 jasper artifacts retained cortex. Items containing cortex within the Plenge collection consist of brown jasper, red jasper, 132

and black cherts. Other materials including gray cherts, white cherts, and chalcedonies exhibit only cobble cortex and occur in low frequencies. This suggests that some cobble sources, containing ­exotic-looking cherts, may have been located near Plenge. This may also support Kraft’s (1973:61, 112) hypothesis that cobbles were collected from the glacial moraines near Plenge. However, the Musconetcong River must also be considered as a possible source of white, gray, and chalcedony-like cryptocrystalline material. Inspection of cobbles within the Musconetcong showed the occurrence of medium-sized cobbles of jasper and black chert that were of knappable quality; these pieces, however, did not exceed 1,070 g. In addition to these sources, several low-quality jasper boulders were present along the periphery of the upper ­terrace (Gingerich and Stewart 2010). Although these boulders are not of knappable quality, they point to a source closer than the jasper quarries of Pennsylvania. Evidence of lower-quality jasper within a few kilometers of high-quality sources

Table 6.2. Metrics and Descriptions of Channel Flakes.

Catalog #

Artifact Portion

Mass (g)

Length (mm)

Width (mm)

P441P P264P P274P P429P P418P P262P P263P Snn-5 Z555 P500P P2295P P2321P P387 P677P P454P Z295 P875 P514 & 513 P569P Z388 P669P P2294 P668P P510P P273 Z547 P606 P676 Z546 Z543 P453P P830P P180P Z678 Z677 P275P Z679 Z545 Z549 Z550 P848P P533P P272P P705P Z680

PR ME PR PR ME ME ME ME ME ME ME PR PR DI ME DI ME PR DI ME ME ME ME ME ME PR PR ME LS DI ME DI ME DI PR ME PR DI PR ME ME ME ME PR PR

.7 .9 .5 .4 .9

20.06 22.93 17.85 17.65 22.27 18.43 18.05 34.46 14.83 21.3 17.4 23.49 25.88 18.38 17.4 29.49 18.74 49.63 22.55 23.06 13.64 14.95 20.56 22.33 20.79 20.47 22.1 13.2 25.88 18.57 21.13 20.18 19.13 8.81 19.5 14.08 17.04 24.61 18.82 15.27 11.59 12.89 17.66 18.71 14.46

12.71 14.71 14.07 10.65 15.11 13.86 12.58 14.62 12.53 13.48 14.21 13.41 21.5 12.8 17.77 19.29 12.31 16.31 14.64 13.72 16.82 13.19 13.61 13.26 9.18 13.53 13.59 10.61 14.33 14.56 14.56 17.22 13.63 11.7 11.02 11.57 13.93 20.77 14.06 14.79 11.38 15.58 14.78 11.21 12.32

1 .5 1.6 .3 .8 .7 .8 2.3 .6 1 2.7 .8 0 1.2 1 .9 .4 .6 .6 .3 .7 1 .2 1.6 .8 .9 .7 .5 .2 .4 .5 .7 2.1 .6 .8 .2 .5 .6 .5 .3

Thickness Raw (mm) Material

2.83 1.96 2.29 1.75 2.22 3.15 1.87 2.39 1.85 2.73 2.31 2.49 4.16 2.17 2.27 4.3 2.71 3.07 3.08 2.9 3.24 1.94 2.18 1.97 1.68 2.82 3.03 1.54 2.96 2.89 2.54 1.76 1.86 1.77 2.19 2.91 2.88 3.75 2.12 2.8 1.81 1.99 2.49 2.35 1.48

RJSP RJSP RJSP JSPR RJSP RJSP RJSP UNC RJSP RJSP JSPR RJSP JSP RJSP RJSP JSP JSP JSP RJSP RJSP JSP RJSP JSPR RJSP RJSP JSPR RJSP RJSP JSP JSP JSPR RJSP RJSP RJSP JSP JSPR RJSP RJSP JSP RJSP JSPR JSPR JSPR JSP JSPR

Multiple Flutes

Yes

Yes? Yes

Yes

Yes

Yes Yes? Yes

Medial Ridge

Yes — — — — Yes — — — Yes — — — — — Yes Yes — — Yes — — yes — — Yes Yes — Yes? — — — — — — — Yes Yes — Yes — — — — —

Notes

Possible channel

Early thinning Probable channel

Spokeshave use?

Probable channel

Possible channel Possible channel? Probable channel

Note: DI = distal; JSP = jasper; JSPR = brown jasper with red mottles or red on the surface; LS = longitudinal split; ME= medial; PR = proximal; RJSP = red jasper; UNC = unidentified chert. Some artifacts listed may not be channel flakes or may have other modifications.

Gingerich

is well documented throughout the region. Bedrock mapping of the jasper-bearing Hardyston Formation within the Reading Prong also runs from Pennsylvania into New Jersey near the site (Davis 2003). All of these data support the idea that toolstone sources were present at or around Plenge. The presence of these sources likely encouraged the continual prehistoric occupation of Plenge, which began during the Late Pleistocene and continued into the Holocene. Presence of Exotic Toolstone at Plenge

As stated above, the majority of raw material used at Plenge appears to be local toolstone. Bedrock cortex and cobble cortex are both represented at the site, but the presence of large primary flakes and higher frequencies of bedrock cortex suggest the more common use of a major outcrop. Flakes made of gray cherts, white cherts, and chalcedonies within the assemblage occasionally have ­cobble cortex and also suggest a local but secondary source. Other cherts, however, are recognized as being transported from primary sources in the Northeast. These cherts include Devonian chert (probably Onondaga), Normanskill, and possibly Munsungun Lake cherts (Davis 2003; Kraft 1973; Pollock et al. 1999).2 Most of these raw material types appear as finished tools or projectile points. Other artifacts made from these exotic materials are almost exclusively small debitage (