Discover millions of ebooks, audiobooks, and so much more with a free trial

Only $11.99/month after trial. Cancel anytime.

Ceramics, Chronology, and Community Patterns: An Archaeological Study at Moundville
Ceramics, Chronology, and Community Patterns: An Archaeological Study at Moundville
Ceramics, Chronology, and Community Patterns: An Archaeological Study at Moundville
Ebook728 pages6 hours

Ceramics, Chronology, and Community Patterns: An Archaeological Study at Moundville

Rating: 0 out of 5 stars

()

Read preview

About this ebook

A clearly written description of the analytical procedures employed on ceramic samples obtained at Moundville and the new chronology discovered

Moundville, located on the Black Warrior River in west-central Alabama, is one of the best known and most intensively studied archaeological sites in North America. Yet, in spite of all these investigations, many aspects of the site's internal chronology remained unknown until the original 1983 publication of this volume. The author embarked on a detailed study of Moundville ceramics housed in museums and collections, and hammered out a new chronology for Moundville.This volume is a clearly written description of the analytical procedures employed on these ceramic samples and the new chronology this study revealed. Using the refined techniques outlined in this volume, it was possible for the author to trace changes in community patterns, which in turn shed light on Moundville's internal development and its place among North America's ancient cultures.
 
LanguageEnglish
Release dateMay 24, 2009
ISBN9780817382506
Ceramics, Chronology, and Community Patterns: An Archaeological Study at Moundville

Read more from Vincas P. Steponaitis

Related to Ceramics, Chronology, and Community Patterns

Related ebooks

Archaeology For You

View More

Related articles

Reviews for Ceramics, Chronology, and Community Patterns

Rating: 0 out of 5 stars
0 ratings

0 ratings0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    Ceramics, Chronology, and Community Patterns - Vincas P. Steponaitis

    CERAMICS, CHRONOLOGY, AND COMMUNITY PATTERNS

    CERAMICS, CHRONOLOGY, AND COMMUNITY PATTERNS

    An Archaeological Study at Moundville

    VINCAS P. STEPONAITIS

    The University of Alabama Press

    Tuscaloosa

    New preface © 2009 by Vincas P. Steponaitis

    Original text © 1983 by Vincas P. Steponaitis

    Publishing history: work originally published by Academic Press

    The University of Alabama Press

    Tuscaloosa, Alabama 35487-0380

    All rights reserved

    Manufactured in the United States of America

    The paper on which this book is printed meets the minimum requirements of American National Standard for Information Sciences-Permanence of Paper for Printed Library Materials, ANSI Z39.48-1984.

    Library of Congress Cataloging-in-Publication Data

    Steponaitis, Vincas P.

         Ceramics, chronology, and community patterns : an archaeological study at Moundville / Vincas P. Steponaitis.

                p. cm.

         Includes bibliographical references and index.

         ISBN 978-0-8173-5576-0 (pbk. : alk. paper) — ISBN 978-0-8173-8250-6 (electronic) 1. Moundville Archaeological Park (Moundville, Ala.) 2. Mississippian culture—Alabama—Black Warrior River Valley. 3. Mississippian pottery—Alabama—Black Warrior River Valley. 4. Land settlement patters, Prehistoric—Alabama—Black Warrior River Valley. 5. Black Warrior River Valley (Ala.)—Antiquities. I. Title.

         E99.M6815S73 2009

         976.1’43—dc22

    2009005898

    For Laurie

    Contents

    Illustrations

    Tables

    Preface

    Acknowledgments

    Preface to the New Edition

    1 - Introduction

    The Site and Its Setting

    Investigations at Moundville prior to 1978

    Aspects of the Ceramic Sample

    2 - Ceramic Technology

    Clays

    Tempering Materials

    Vessel Forming, Finishing, and Decoration

    Ceramic Mineralogy and Firing Temperature

    The Effect of Paste Composition on Physical Properties

    Discussion

    3 - Classification of Moundville Ceramics

    Introduction

    Types and Varieties

    Representational Motifs

    Painted Decoration

    Basic Shapes

    Secondary Shape Features

    Effigy Features

    4 - Ceramic Chronology

    Introduction to Chronology

    Gravelot Seriation

    Stratigraphy

    Moundville I Phase

    Moundville II Phase

    Moundville III Phase

    Summary and Discussion

    5 - Community Patterns At Moundville

    Relative Dating of Vessels and Burials

    Spatial Context of Vessels and Burials

    Changes in Community Patterns through Time

    6 - Conclusion: A Regional Perspective

    The Late Prehistory of the Black Warrior Drainage

    Some Speculations on the Causes of Change

    Vessel and Sherd Illustrations

    Appendix A - Individual Vessel Descriptions

    Appendix B - Vessels Indexed by Burial Number

    Appendix C - Stratigraphic Level Descriptions

    Appendix D - Sherd Frequencies by Level

    Appendix E - Methods for Measuring Physical Properties

    Tensile Strength

    Porosity

    Thermal Diffusivity

    Elasticity

    Appendix F - Type—Variety Descriptions

    Local Types

    Nonlocal Types

    Appendix G - Vessels Indexed by Type and Variety

    Local Types

    Nonlocal Types

    Appendix H - Vessels Indexed by Dimensions Other Than by Type and Variety

    Representational Motifs

    Painted Decoration

    Basic Shapes

    Secondary Shape Features

    Effigy Features

    References

    Index

    Illustrations

    FIGURE   1. Location of Moundville and other Moundville phase sites in the Black Warrior River Valley.

    FIGURE   2. The Moundville site as it appeared in 1905.

    FIGURE   3. Ceramic artifacts from Moundville.

    FIGURE   4. Fire-clouding on white- and black-surfaced vessels.

    FIGURE   5. Frequency distributions for the size of the third largest temper particle in Moundville vessels.

    FIGURE   6. Frequency distribution for the abundance of visible shell temper, expressed as a percentage of total volume.

    FIGURE   7. Tensile strength plotted against the percentage of visible shell temper.

    FIGURE   8. Tensile strength of a ceramic material as a function of thermal history.

    FIGURE   9. Apparent porosity (percentage of volume) plotted against the percentage of visible shell temper.

    FIGURE 10. Thermal diffusivity plotted against the percentage of visible shell temper.

    FIGURE 11. Elasticity plotted against the percentage of visible shell temper.

    FIGURE 12. The thermal shock resistance parameter R plotted against the percentage of visible shell temper.

    FIGURE 13. The thermal shock resistance parameter R′ plotted against the percentage of visible shell temper.

    FIGURE 14. Strength as a function of thermal history for three sherds from Moundville.

    FIGURE 15. The fraction of strength retained after thermal shock plotted against initial strength.

    FIGURE 16. Dendritic key for local types.

    FIGURE 17. Carthage Incised designs.

    FIGURE 18. Moundville Engraved designs.

    FIGURE 19. Moundville Incised designs.

    FIGURE 20. Representational motifs.

    FIGURE 21. Critical points in a vessel profile.

    FIGURE 22. Basic shapes.

    FIGURE 23. Late prehistoric chronology in the Black Warrior drainage.

    FIGURE 24. Nonmetric scaling of attributes in two dimensions.

    FIGURE 25. Seriated sequence of gravelots.

    FIGURE 26. Segments within the seriated sequence.

    FIGURE 27. Distribution of varieties by vessel shape for Moundville Engraved and Carthage Incised.

    FIGURE 28. Distribution of SCC motifs by vessel shape on Moundville Engraved, variety Hemphill.

    FIGURE 29. Handle-shape ratios, unburnished jars from various phases.

    FIGURE 30. Handle-shape ratios, Moundville Incised jars.

    FIGURE 31. Approximate location of West Jefferson phase component.

    FIGURE 32. Spatial distribution of Moundville I burials and unassociated vessels.

    FIGURE 33. Moundville I to early Moundville II phase burial concentration west of Mound O.

    FIGURE 34. Moundville I or early Moundville II phase structure with burials.

    FIGURE 35. Spatial distribution of Moundville II phase burials and unassociated vessels.

    FIGURE 36. Spatial distribution of Moundville III phase burials and unassociated vessels.

    FIGURE 37. Portion of the late Moundville II to early Moundville III phase burial concentration south of Mound D.

    FIGURE 38. Spatial distribution of Alabama River phase burials and unassociated vessels.

    FIGURE 39. Bell Plain, Moundville Engraved, Carthage Incised, and Moundville Incised vessels, Moundville I phase.

    FIGURE 40. Moundville Engraved and Carthage Incised sherds, Moundville I phase.

    FIGURE 41. Carthage Incised, Bell Plain, and Mississippi Plain sherds, some of which show painted decoration, Moundville I phase.

    FIGURE 42. Moundville Incised sherds, Moundville I phase.

    FIGURE 43. Rim and base profiles, Moundville I phase.

    FIGURE 44. Bell Plain and Moundville Engraved vessels, Moundville I phase.

    FIGURE 45. Bell Plain and Moundville Engraved vessels, Moundville II phase.

    FIGURE 46. Bell Plain, Mississippi Plain, and Moundville Engraved vessels, Moundville II phase.

    FIGURE 47. Moundville Engraved and Carthage Incised sherds, Moundville II phase.

    FIGURE 48. Moundville Incised and Mississippi Plain sherds, Moundville II phase.

    FIGURE 49. Bell Plain sherds, some of which have painted decoration, Moundville II phase.

    FIGURE 50. Rim and base profiles, Moundville II phase.

    FIGURE 51. Bell Plain and Carthage Incised vessels, Moundville III phase.

    FIGURE 52. Bell Plain, Carthage Incised, and Moundville Engraved vessels, Moundville III phase.

    FIGURE 53. Bell Plain, Carthage Incised, and Moundville Engraved vessels, Moundville II phase.

    FIGURE 54. Bell Plain and Mississippi Plain jars, Moundville III phase.

    FIGURE 55. Moundville Engraved sherds from Moundville III levels.

    FIGURE 56. Carthage Incised, Barton Incised, Alabama River Incised, Bell Plain, and Mississippi Plain sherds, some with painted decoration, from Moundville III levels.

    FIGURE 57. Mississippi Plain sherds, Moundville III levels.

    FIGURE 58. Bell Plain, variety Hale sherds from Moundville III levels.

    FIGURE 59. Rim profiles of Moundville III and Alabama River phase sherds.

    FIGURE 60. Mississippi Plain and Moundville Incised sherds, found in Moundville III levels but probably intrusive.

    FIGURE 61. Unclassified sherds from various levels.

    FIGURE 62. Bell Plain, Carthage Incised, Moundville Engraved, and Alabama River Incised vessels, various phases.

    FIGURE 63. Bell Plain, Carthage Incised, Moundville Engraved, and Moundville Incised vessels, various phases.

    FIGURE 64. Nonlocal vessels.

    FIGURE 65. Nonlocal vessels.

    FIGURE 66. Nonlocal sherds from various levels.

    FIGURE 67. Woodland period sherds from various levels.

    Tables

    TABLE     1. Locality Prefixes and Special Symbols, Listed in the Conventional Order of Precedence

    TABLE     2. Clay Samples from Outcrops in the Vicinity of Moundville

    TABLE     3. Mineralogical Composition of Clay Samples

    TABLE     4. Mineralogical Composition of Slip and Vessel Wall in Specimen S-1

    TABLE     5. Sherds from Moundville Used in Mineralogical Studies

    TABLE     6. Mineralogical Composition of Sherds from Moundville

    TABLE     7. Relative Abundances of Coarse Mineral Grains in Sherds, Estimated as a Percentage of Total Volume

    TABLE     8. Physical Properties Measurements

    TABLE     9. Cross-tabulation of Head Form versus Orientation on Effigy Bowls

    TABLE   10. West Jefferson Phase Radiocarbon Dates

    TABLE   11. Chronologically Sensitive Attributes Used in Seriation

    TABLE   12. Seriated Distribution of Types and Varieties

    TABLE   13. Seriated Distribution of Representational Motifs

    TABLE   14. Seriated Distribution of Painted Decoration

    TABLE   15. Seriated Distribution of Basic Shapes

    TABLE   16. Seriated Distribution of Secondary Shape Features

    TABLE   17. Seriated Distribution of Effigy Features

    TABLE   18. Stratigraphic Distribution of Types and Varieties

    TABLE   19. Stratigraphic Distribution of Painted Decoration

    TABLE   20. Stratigraphic Distribution of Basic Shapes

    TABLE   21. Stratigraphic Distribution of Secondary Shape Features

    TABLE   22. Handle Measurements of Moundville I Unburnished Jars

    TABLE   23. Moundville I Phase Radiocarbon Dates

    TABLE   24. Handle Measurements of Late Moundville II Unburnished Jars

    TABLE   25. Handle Measurements of Early Moundville III Unburnished Jars

    TABLE   26. Handle Measurements of Late Moundville III Unburnished Jars

    TABLE   27. Handle Measurements of Moundville Incised Jars

    TABLE   28. Handle Measurements of Moundville III Burnished Jars

    TABLE   29. Summary Chronology of Types and Varieties

    TABLE   30. Summary Chronology of Representational Motifs

    TABLE   31. Summary Chronology of Painted Decoration

    TABLE   32. Summary Chronology of Basic Shapes

    TABLE   33. Summary Chronology of Secondary Shape Features

    TABLE   34. Summary Chronology of Effigy Features

    TABLE   35. Temporal Placement of Burials

    TABLE   36. Temporal Placement of Unassociated Vessels (Local)

    TABLE A.1. Vessels Arranged by Burial Number

    TABLE A.2. Excavated Levels in Unit 6N2W

    TABLE A.3. Excavated Levels in Unit 8N2E

    TABLE A.4. Depositional Correlation between Levels in Units 6N2W and 8N2E

    TABLE A.5. Sherd Frequencies by Level, Unit 6N2W

    TABLE A.6. Sherd Frequencies by Level, Unit 8N2E

    Preface

    Moundville, located on the Black Warrior River in west-central Alabama, is one of the best known and most intensively studied prehistoric sites in North America. It first gained wide recognition just after the turn of the century, when it was visited, excavated, and reported on by C. B. Moore. Later, during the Great Depression, large-scale excavations at Moundville produced a wealth of information on the lifeways and material culture of the Mississippian people who lived there. Later still, in the 1960s and 1970s, the site gained further importance as the locus of investigations into the social and political structure of one of the most complex indigenous societies north of Mexico (e.g., Peebles 1971, 1978a, 1978b).

    Yet despite all this work, even as late as 1978, many aspects of the site’s internal chronology remained obscure. Prior to this time, all artifacts and features at the site were assigned to a single Moundville phase, which, in effect, compressed five centuries of cultural development into a single undifferentiated mass. The excessive length of this phase had many undesirable consequences, not the least of which was that it prevented anyone from clearly delineating the pathways and mechanisms by which this complex site and its regional system evolved.

    The research discussed in the following chapters was mainly designed to rectify this problem. It was evident from the start that the most practical way to improve the chronology was through a detailed study of Moundville ceramics, large samples of which were available in museum collections. We visited the museums, photographed the ceramics, and with these and other data in hand, gradually hammered out the new chronology. Using this refined sequence, it was possible to trace changes in community patterns, which in turn shed light on Moundville’s development.

    This book should prove useful to scholars in a variety of ways, for it deals with culture-historical, methodological, and evolutionary issues. In addressing matters of culture history, the study presents a detailed and well-documented Mississippian chronology—one of the very few in existence. It systematizes and describes the late prehistoric ceramics from an important site, providing a benchmark that can be used in future studies of trade and stylistic interaction. And, perhaps even more importantly, it synthesizes the available information on Moundville’s development, taking into account settlement, subsistence, and mortuary evidence.

    The methodological contributions of this book are twofold: First, the chapter on ceramic technology introduces a variety of approaches, borrowed from materials science, that heretofore have not been applied to archaeological materials. Second, the methods used to define the ceramic chronology are in certain ways novel and provide one of the few existing illustrations of how to seriate grave assemblages using modern numerical techniques.

    Finally, with regard to the subject of cultural evolution, the changes in complexity that occurred in the Moundville region during late prehistoric times are reconstructed, and some possible causes are suggested. The cultural evolutionary and methodological aspects of this book should prove interesting to archaeologists regardless of their geographical specialty.

    Chapter 1 provides the general background for the study, describing the Moundville site and its setting, the history of archaeological research there, and describing the nature of the ceramic sample on which the various analyses were carried out. Chapter 2 deals with the materials and techniques of pottery manufacture at Moundville and would profit any reader with a general interest in ceramic technology. Chapter 3 presents a new classification of Moundville ceramics, intended primarily for the regional specialist. Chapter 4 sets forth the refined ceramic chronology, along with the evidence that led to its formulation; this chapter should be of use not only to the regional specialist, but also to anyone interested in the application of numerical seriation techniques. Chapter 5 examines changes in community patterns at Moundville as revealed by this chronology, and Chapter 6 discusses the significance of these and related changes in the context of the region as a whole. For anyone concerned with the evolution of chiefdom-level societies, these two concluding chapters are especially relevant.

    Acknowledgments

    The present endeavor is actually part of a larger project that was organized by Christopher Peebles in 1977. The overall aim of the project was to attain a better understanding of the Moundville phase, particularly with regard to questions concerning the development and decline of the complex Mississippian society that the phase appeared to represent. At its inception, the project was planned as a cooperative venture among four researchers: Margaret Scarry was to reconstruct subsistence using excavated food remains; Margaret Schoeninger was to be concerned with the biocultural aspects of nutrition, using osteological data from human burials; Peebles was to conduct a surface reconnaissance in order to gather detailed information on settlement patterns; and I was to construct a ceramic chronology so that fine-grained temporal control in all areas of investigation could be achieved. Although each of these lines of inquiry was to be pursued somewhat independently, the hope was that ultimately the various lines would converge to attain the project’s overall aims. The project was funded by a National Science Foundation grant to the University of Michigan (BNS78–07133). Fieldwork began in June 1978, and was carried on intermittently until August 1979. A tremendous amount of information was gathered during this interval by all the investigators, and much of this information is, at this writing, still in the process of being analyzed.

    For my own part of the project, fieldwork was mostly carried out indoors, recording data on whole vessels from Moundville in extant museum collections. I was assisted in this task by Laurie Cameron Steponaitis, who photographed all the vessels, developed and printed the film, and helped the fieldwork along in innumerable other ways. It is safe to say that without her talents, we would never have accomplished as much on this project as we did. Other people who, at various times, helped us sift through these collections are John Blitz, Gail Cameron, Mary Meyer, Masao Nishimura, Jeffrey Parsons, John Scarry, Margaret Scarry, Letitia Shapiro, Deborah Walker, and Paul Welch. Their willingness to put up with the tedium and dust while looking through countless boxes is gratefully acknowledged.

    Our obsessive search for Moundville collections eventually took us to five museums in four different states. Although not all of the collections were found to contain whole vessels, the work was invariably made more comfortable and productive by the staffs of the institutions we visited. Among those to be thanked are Joseph Vogel, John Hall, and Dorothy Beckham of the Alabama Museum of Natural History; Richard Krause, Kenneth Turner, and Amelia Mitchell of the Department of Anthropology, University of Alabama; Carey Oakley, Eugene Futato, and Tim Mistovic of the Office of Archaeological Research, University of Alabama; David Fawcett, James Smith, and Anna Roosevelt of the Museum of the American Indian, Heye Foundation; Vincent Wilcox, Joseph Brown, and Marguerite Brigida of the Department of Anthropology, National Museum of Natural History, Smithsonian Institution; Barbara Conklin of the American Museum of Natural History; and Richard MacNeish of the R. S. Peabody Foundation, Andover.

    Once the data had been collected, the bulk of the analysis was carried out at the Smithsonian Institution, where I was appointed a predoctoral fellow. Bruce Smith, my advisor while on fellowship, was truly a stimulating colleague with whom to work. He not only shared freely his ideas on Mississippian culture, but also provided logistic, bureaucratic, and moral support in more ways than I can possibly enumerate. A number of other people at the Smithsonian contributed substantially to the effort as well. David Bridge was instrumental in helping me grasp the complexities of SELGEM, the data-banking program with which I managed to keep track of all the vessels. Jane Norman helped by reconstructing beautifully a vessel which seemed to be fragmented beyond hope. Also to be acknowledged is Florence Jones, who, as a Smithsonian Institution volunteer, ably drew most of the rim profiles that appear in this report.

    The technological studies of ceramics were all done at the National Bureau of Standards in Gaithersburg, Maryland. I arrived at the bureau as a complete novice in materials science, mindful of issues that needed to be studied, but with no inkling of how to actually go about doing it. I was most fortunate, therefore, to fall in with a group of experienced colleagues who never seemed to tire of my endless questions. Carl Robbins, for one, spent countless hours showing me how to use a petrographic microscope, and helping me with mineralogical identifications. He also produced all of the x-ray diffraction patterns on which many of my conclusions are based. When it came to matters concerning physical properties (or, as the division title quaintly put it, Fracture and Deformation), Alan Franklin, C. K. Chiang, Ed Fuller, and Steve Freiman were the experts. Together, they introduced me to some rather unaccustomed definitions of stress and strain, and showed me how to make the measurements that were critical to the successful outcome of my research. All of this work was made possible by my appointment as a guest worker at the bureau, under the sponsorship of Carl Robbins. I am also particularly grateful to Alan Franklin and Jacqueline Olin (Conservation Analytical Laboratory, Smithsonian Institution), who were both instrumental in bringing about this arrangement. Ceramic thin sections were obtained through the courtesy of Daniel Appleman, chairman of the Department of Geology, Smithsonian Institution.

    Most of the actual writing was done while I was on the faculty at the State University of New York at Binghamton. Several people at this institution helped a great deal in edging the (sometimes reluctant) manuscript toward its completion. Sherd photographs were taken by David Tuttle, who also did some supplementary darkroom work. The vast majority of figures were capably prepared for publication by Laurie Cameron Steponaitis, and the rest by Stan Kauffman. Robert Stuckart of the Computer Center helped by showing me how to get SUNY—Binghamton’s overgrown (and accident-prone) calculator—to do what it was supposed to do.

    As noted earlier, the project was, from the very start, a multifaceted affair, designed to make the most out of a collaboration between researchers working on different aspects of a common goal. To my colleagues on this project—Christopher Peebles, Margaret Scarry, Paul Welch, Tandy Bozeman, Margaret Hardin, and Margaret Schoeninger—I owe a great deal for their stimulating thoughts, for their willingness to share information, and, in general, for making our collaboration such a pleasant one.

    Other people who offered valuable help and suggestions are William Autry, Jeffrey Brain, Ian Brown, Ben Coblentz, David DeJarnette, Richard Ford, James B. Griffin, Ned Jenkins, David Kelley, Keith Kintigh, William Macdonald, Dan Morse, J. Mills Thornton, Sander van der Leeuw, Stephen Williams, and Henry Wright. It was van der Leeuw who first opened my eyes to the need for a detailed understanding of ceramic technology. Brain, Brown, Ford, Griffin, Kintigh, and Williams were particularly generous with their comments on the manuscript, all of which were appreciated, although not all were accepted. Autry, Coblentz, Kelley, and especially Jenkins shared their knowledge of southeastern pottery, and helped identify nonlocal vessels in the collections. Finally, it was Morse who spared me untold embarrassment by gently pointing out that the weirdest Mississippian jar I had ever seen was actually a Formative period Mexican type that had been mistakenly catalogued as coming from Moundville.

    With all the help that I have received from these people and others, it should be clear that any and all faults that remain in the book are entirely my own responsibility.

    Preface to the New Edition

    This book grew out of my involvement with the University of Michigan Museum of Anthropology’s Moundville Project led by Christopher Peebles in the late 1970s. At the outset, Moundville’s chronology was poorly understood and it was my role to fix this problem. I was acutely aware that a finer chronology was essential not only for my own purposes, but also for all the other research being conducted under the project’s umbrella (e.g., Bozeman 1982; Scarry 1986; Welch 1986). Hence, I felt it was especially important to produce a chronological scheme that others could easily use, which in turn required devoting considerable attention to basic description and systematics—a style of research that, during the heyday of the New Archaeology, was not wildly popular (to put it mildly). Be that as it may, I was very fortunate to have colleagues and mentors who saw the value of this approach and encouraged me bring it to completion. The results were first published as a monograph in 1983.

    Now that the book is being republished, it seems appropriate to review briefly some of the things we have learned over the past 25 years, particularly with reference to how well my original arguments and conclusions have held up. I present this review in four broad areas that parallel the book’s sequence of chapters: technology, classification, chronology, and community patterns.

    It is no accident that my book began with an extensive treatment of how Moundville pots were made (see chapter 2). In the late 1970s, studying Mississippian ceramic technology seemed like the cutting edge. Yet surprisingly little work in this area has been done since that time. One series of studies looked at the firing temperature of Mississippian pottery from other sites and concluded that some pots may have been fired to 800–900°C, considerably higher than my estimates for Moundville (Stimmell et al. 1982; Klemptner and Johnson 1985, 1986). Such higher temperatures might have been made possible by the addition of salt, which inhibits the destructive process of lime spalling that otherwise occurs when a shell-tempered pot is fired above 800°C. However, the archaeological evidence presented thus far is still open to question. The best indicator of high temperatures is the apparent presence of the mineral gehlenite, which only forms above 800°C and was detected by x-ray diffraction (Stimmell et al. 1982: 221–223). The main x-ray peak that Stimmell et al. identified as gehlenite occurred at 2.85 Å, yet calcite (the major constituent of shell temper) can also produce a peak, albeit usually smaller, at virtually the same location. This peak occurred in all my Moundville samples and was associated with shell, not gehlenite, as evidenced by the fact that treating a sample with dilute hydrochloric acid caused the peak to disappear. Clearly, more work needs to be done before the matter is settled, but for now I see no compelling reason to alter my original conclusion that Moundville pots were generally fired below 800°C (p. 33).

    Another study examined the physical properties of fine and coarse shell-tempered pottery, and compared these properties to those of sand- and grog-tempered wares (Bronitsky and Hamer 1986; Bronitsky 1986: 248–259). This study relied on a large sample of replicated briquettes and used impact testing to measure the strength and thermal shock resistance of various paste compositions. Without going into full detail, suffice it to say that their results, although not identical, were broadly similar to mine. Among briquettes tempered with burned shell (the kind found in Moundville pottery), finely tempered wares were found to be more resistant to impact damage than coarsely tempered ones, although the amount of temper seemed to have no effect on strength. Also, wares tempered with larger amounts of shell were found to be more resistant to thermal shock than ones tempered with lesser amounts. The main problem in comparing our results is that Bronitsky and Hamer used briquettes in which the shell-temper particles were randomly oriented, rather than preferentially aligned with the surface as is typically the case with Mississippian pots (1986: 91). Such preferential orientation, mainly caused by paddle-and-anvil finishing, has a major effect on both strength and thermal-shock resistance in that it impedes the propagation of fatal cracks through the vessel wall (Steponaitis 1984: 112). Given my small sample size and reliance on archaeological sherds (whose physical properties may have been affected by use and postdepositional processes), the work I presented in chapter 2 was at best a pilot study, with results that were more suggestive than conclusive. Yet a quarter century later, a conclusive body of research on this subject is still lacking—perhaps a testament to the obstacles inherent in pursuing interdisciplinary work of this kind.

    Turning now to matters of classification, the type-variety system I introduced in chapter 3 has, for better or worse, been widely adopted and survives largely intact. A few new varieties have been added, as one would expect, but the overall framework has proved to be resilient (e.g., Wilson 2008; Knight, 2010). Advances have come in three areas. First, vessel forms have been described in more detail, showing the range of sizes and forms present in Moundville II–III middens (Taft 1996). Second, stylistic and iconographic studies have greatly refined our understanding of the representational art on Moundville pottery, in terms of both form and meaning (Lacefield 1995; Schatte 1997, 1998; Gillies 1998; Lankford 2007; Steponaitis and Knight 2004). Third, geochemical and stylistic studies have in some cases changed my original interpretations of provenance, i.e., whether the vessels were locally made or imported (Steponaitis et al. 1996; Neff et al. 1991). Specifically, I now believe that certain vessels I originally classified as Caddoan were locally made (e.g., fig. 65d), and the same can be said of some vessels that I originally considered to be from the Gulf Coast (e.g., fig. 65p, 66d–e) or from the Tennessee-Cumberland region (e.g., fig. 66j; Knight and Franke 2007: 148–150). But again, the vast majority of provenance assignments have remained intact.

    The Moundville chronology presented in chapter 4 has also held up reasonably well, at least in terms of the sequence and content of ceramic phases. Curren (1984) renamed the final phase Moundville IV to distinguish it from Alabama River phase assemblages farther south, a change that has now been widely adopted and confirmed by more detailed analysis (Little and Curren 1995; Regnier 2006). Knight’s extensive program of mound excavations in the 1990s essentially corroborated my ceramic sequence with only minor adjustments (Knight, 2010). Perhaps the most important refinement has been in absolute dates. Knight’s excavations yielded a large suite of radiocarbon determinations which, when calibrated and statistically combined (Knight et al. 1999), yielded new phase boundaries as follows: West Jefferson, AD 1020–1120; Moundville I, AD 1120–1260; Moundville II, AD 1260–1400; Moundville III, AD 1400–1520; and Moundville IV, AD 1520–1650. Essentially the first two phases have been compressed and pushed a bit later in time, while the last three phases have stayed approximately the same.

    Subsequent work has also brought the pottery of the early Moundville I phase into clearer focus (Steponaitis 1992; Welch 1994; Michals 1998; Scarry 1998; Johnson 2005; Wilson 2008). These assemblages are dominated by Mississippi Plain, var. Warrior, with a small but significant presence of Baytown Plain, var. Roper. Decorated varieties include Moundville Incised, var. Moundville, Barton Incised, var. Oliver, very little Carthage Incised, and little or no Moundville Engraved. Perhaps the best marker is a large percentage of neckless jars with folded-flattened rims (which can occur in both shell- and grog-tempered forms).

    A key question regarding the early end of the sequence has been the nature of the shift from grog- to shell-tempered pottery. Originally, I saw the transition as fairly abrupt and resulting from local development (pp. 164–167), whereas Jenkins (1978) saw it as more gradual and resulting from intrusions of foreigners into the region. The problem then, as now, lay in interpreting contexts that contained both kinds of pottery together. Is such co-occurrence the result of stratigraphic mixing or concurrent use? It is very difficult to be sure, especially when looking only at published sherd counts. Even a perfect battleship seriation showing what appears to be a long, gradual transition could just as easily be an ordering of contexts showing varying degrees of mixture (cf. Mistovich 1988). While acknowledging this ambiguity, I still believe that the ceramic transition in the Black Warrior Valley was relatively quick, largely based on the rarity of transitional forms and the complete absence of a grog-tempered version of Moundville Incised (Welch 1994). That said, I am also much more willing than I was in 1983 to accept the possibility that Moundville’s early history may have involved population movements and the coalescence of groups with distinct ceramic traditions (Jenkins 2003). It now appears that grog-tempered pottery assemblages in the West Jefferson style continued to be used near the headwaters of the Black Warrior River as late as AD 1200, a date fully contemporary with the Moundville I phase (Jones 2004). Although the social and political meaning of this persistence has yet to be worked out, it presents some very interesting avenues for future research.

    The biggest change over the past quarter century has not been in matters of systematics or chronology, but in our understanding the history of Moundville itself. In this book I presented a model, based exclusively on burial evidence, of a gradual increase in the site’s size and complexity, followed by an abrupt decline (see chapter 5). Today we know this scenario was wrong. More comprehensive analyses of pottery from middens and the constructional evidence gleaned from Knight’s mound excavations have led us to exactly the opposite conclusion, that population and mound construction at Moundville peaked early in the sequence, after which there was a gradual decline (Knight and Steponaitis 1998a). According to the new model, the site began as a minor civic-ceremonial center during the early Moundville I phase. In late Moundville I and early Moundville II times, the site’s population grew and major construction of the palisade and mounds took place. By the late Moundvile II phase, most of Moundville’s resident population had dispersed and the site had become a necropolis, a center of ritual and a place of burial for the region’s dead. This state of affairs continued until late Moundville III times, when the site was largely abandoned. Only an ephemeral occupation took place in Moundville IV. A complete account of this history has been presented elsewhere (Knight and Steponaitis

    Enjoying the preview?
    Page 1 of 1