Early Mesoamerican Social Transformations: Archaic and Formative Lifeways in the Soconusco Region 9780520950566

Between 3500 and 500 bc, the social landscape of ancient Mesoamerica was completely transformed. At the beginning of thi

156 55 6MB

English Pages 304 [303] Year 2011

Report DMCA / Copyright

DOWNLOAD PDF FILE

Table of contents :
Contents
Contributors
Acknowledgments
ONE: Early Social Transformations in the Soconusco
PART ONE. Archaic to Formative. Transformations in Subsistence
TWO. A Gender- Based Model for Changes in Subsistence and Mobility During the Terminal Late Archaic Period on the Coast of Chiapas, Mexico
THREE. Evidence for the Diversity of Late Archaic and Early Formative Plant Use in the Soconusco Region of Mexico and Guatemala
FOUR. Archaic to Formative in Soconusco: The adaptive and organizational transformation
PART TWO. Emergent Complexity. The Archaeological Records of Early Political Centers
FIVE. Building History in Domestic and Public Space at Paso de la Amada: An examination of mounds 6 and 7
SIX Paso de la Amada as a Ceremonial Center
SEVEN. A History of Disaster and Cultural Change in the Coatán River Drainage of the Soconusco, Chiapas, Mexico
EIGHT. La Blanca and the Soconusco Middle Formative
PART THREE. Beyond the Individual Study Area. Grappling With Issues of Scale
NINE. Early Formative Transitions in Settlement and Subsistence at Chiquiuitan, Guatemala
TEN. Jocotal Settlement Patterns, Salt Production, and Pacifi c Coast Interactions
ELEVEN. An Early Mesoamerican Archipelago of Complexity
TWELVE. Concluding Thoughts: Macroregional synthesis in the archaeology of early mesoamerica
Index
Recommend Papers

Early Mesoamerican Social Transformations: Archaic and Formative Lifeways in the Soconusco Region
 9780520950566

  • 0 0 0
  • Like this paper and download? You can publish your own PDF file online for free in a few minutes! Sign Up
File loading please wait...
Citation preview

EARLY MESOAMERICAN SOCIAL TRANSFORMATIONS Archaic and Formative Lifeways in the Soconusco Region

Edited by

Richard G. Lesure

UNIVER SIT Y OF C A LIFORNIA PRE SS Berkeley Los Angeles London

University of California Press, one of the most distinguished university presses in the United States, enriches lives around the world by advancing scholarship in the humanities, social sciences, and natural sciences. Its activities are supported by the UC Press Foundation and by philanthropic contributions from individuals and institutions. For more information, visit www.ucpress.edu. University of California Press Berkeley and Los Angeles, California University of California Press, Ltd. London, England © 2011 by the Regents of the University of California Library of Congress Cataloging-in-Publication Data Early Mesoamerican social transformations : archaic and formative lifeways in the Soconusco region / edited by Richard G. Lesure. p. cm. Includes bibliographical references and index. ISBN 978-0-520-26899-9 (cloth : alk. paper) 1. Indians of Mexico—Mexico—Soconusco Region—Antiquities. 2. Indian pottery—Mexico—Soconusco Region—Antiquities. 3. Indians of Mexico—Implements—Mexico—Soconusco Region. 4. Excavations (Archaeology)—Mexico—Soconusco Region. 5. Social archaeology—Mexico—Soconusco Region. 6. Soconusco Region (Mexico) I. Lesure, Richard G. F1219.1.S6E24 2011 972'.75—dc22 2011003601 19 18 17 16 15 14 13 12 11 10 9 8 7 6 5 4 3 2 1 The paper used in this publication meets the minimum requirements of ANSI/NISO Z39.48-1992 (R 1997) (Permanence of Paper). Cover photos, top to bottom: Conchas-phase fi gurines from the Cuauhtémoc zone (photo by Robert M. Rosenswig); Mound 6, Structure 4, Paso de la Amada, 1993 (photo by Michael Blake); west profi le of the Vásquez Mound, El Varal, 1992 (photo by Richard G. Lesure).

contents

Contributors / vii Acknowledgments / ix

6 • PASO DE LA AMADA AS A CEREMONIAL CENTER / 119

Richard G. Lesure 1 • EARLY SOCIAL TRANSFORMATIONS IN THE SOCONUSCO: AN INTRODUCTION / 1

Richard G. Lesure

Part I • Archaic to Formative: Transformations in Subsistence 2 • A GENDER-BASED MODEL FOR CHANGES IN SUBSISTENCE AND MOBILITY DURING THE TERMINAL LATE ARCHAIC PERIOD ON THE COAST OF CHIAPAS, MEXICO / 27

Barbara Voorhies and Douglas J. Kennett 3 • EVIDENCE FOR THE DIVERSITY OF LATE ARCHAIC AND EARLY FORMATIVE PLANT USE IN THE SOCONUSCO REGION OF MEXICO AND GUATEMALA / 47

Michael Blake and Hector Neff 4 • ARCHAIC TO FORMATIVE IN SOCONUSCO: THE ADAPTIVE AND ORGANIZATIONAL TRANSFORMATION / 67

7 • A HISTORY OF DISASTER AND CULTURAL CHANGE IN THE COATÁN RIVER DRAINAGE OF THE SOCONUSCO, CHIAPAS, MEXICO / 146

Gerardo Gutiérrez 8 • LA BLANCA AND THE SOCONUSCO MIDDLE FORMATIVE / 170

Michael Love and Julia Guernsey

Part III • Beyond the Individual Study Area: Grappling With Issues of Scale 9 • EARLY FORMATIVE TRANSITIONS IN SETTLEMENT AND SUBSISTENCE AT CHIQUIUITAN, GUATEMALA / 191

Molly Morgan 10 • JOCOTAL SETTLEMENT PATTERNS, SALT PRODUCTION, AND PACIFIC COAST INTERACTIONS / 217

Mary E. Pye, John Hodgson, and John E. Clark

Richard G. Lesure and Thomas A. Wake 11

Part II • Emergent Complexity: The Archaeological Records of Early Political Centers 5 • BUILDING HISTORY IN DOMESTIC AND PUBLIC SPACE AT PASO DE LA AMADA: AN EXAMINATION OF MOUNDS 6 AND 7 / 97

Michael Blake

• AN EARLY MESOAMERICAN ARCHIPELAGO OF COMPLEXITY / 242

Robert M. Rosenswig 12 • CONCLUDING THOUGHTS: MACROREGIONAL SYNTHESIS IN THE ARCHAEOLOGY OF EARLY MESOAMERICA / 272

Richard G. Lesure Index / 281

contributors

michael blake

michael love

Department of Anthropology University of British Columbia

Department of Anthropology California State University, Northridge

john e. clark

molly morgan

Department of Anthropology Brigham Young University

Department of Anthropology Vanderbilt University

julia guernsey

hector neff

Department of Art and Art History University of Texas, Austin

Department of Anthropology California State University, Long Beach

gerardo gutiérrez

mary e. pye

Department of Anthropology University of Colorado at Boulder

New World Archaeological Foundation Brigham Young University

john hodgson

robert m. rosenswig

Department of Anthropology University of Wisconsin

Department of Anthropology University at Albany—SUNY

douglas j. kennett

barbara voorhies

Department of Anthropology University of Oregon

Department of Anthropology University of California, Santa Barbara

richard g. lesure

thomas a. wake

Department of Anthropology University of California, Los Angeles

Cotsen Institute of Archaeology University of California, Los Angeles

vii

acknowledgments

Most of the chapters in this book were originally presented as papers at the conference “Sociopolitical Transformation in Early Mesoamerica: Archaic to Formative in the Soconusco Region,” held at the Cotsen Institute of Archaeology at UCLA, February 28–March 1, 2008. The introduction and conclusion (Chapters 1 and 12) were written specifically for this book, and Chapter 6 was added to round out coverage of the all-

important site of Paso de la Amada. Funding for the conference was provided by the Cotsen Institute of Archaeology. The editor would particularly like to thank Charles Stanish, director of the Institute, for his generosity and encouragement. Helle Girey’s dedication and organizational skills made the conference itself a great success. Terry Powis and Jeff rey Blomster provided detailed and very helpful critique of the manuscript.

ix

ONE

Early Social Transformations in the Soconusco an introduction Richard G. Lesure

T

he period 3500 b.c. to 500 b.c.

was one of momentous change in Mesoamerica. At the beginning of that span, the region was sparsely occupied by low-level food producers whose rhythms of existence were dominated by the concerns of hunting and gathering. By 500 b.c., it was populated with settled agriculturalists in a landscape increasingly full of people. Proto-urban communities, laid out according to spatial schemes that would continue through the Spanish Conquest two millennia later, were foci of social and political life. Public rituals included worship of deities that were to persist into the Classic and Postclassic eras. Though there was a complex history still to come, Mesoamerica as a culture area or civilization is by that point recognizable. Major developments between 3500 b.c. and 500 b.c. included the shift from the Archaic to the Formative period after 1900 b.c., a farreaching social transformation that involved the appearance of permanent villages, the introduction of pottery, the refocus of subsistence organization toward agriculture, and a steep

rise in population.* The transition to the Formative was followed— still in the second millennium b.c. and thus rapidly on archaeological time scales—by the first indicators of inequality and sociopolitical complexity, including the emergence of settlement hierarchies, monumental sculpture with themes of rulership, and massive earthen platforms or modifications to the landscape. The period also includes, from 1400 b.c., a series of stylistic horizons involving the virtually pan-Mesoamerican dissemination of strikingly recognizable visual culture— including, most famously, Olmec art. Despite the concentration of those three phenomena in the pivotal second millennium b.c., considerations of them as a package, in a unit of time that straddles the divide between Archaic and Formative, is still relatively rare. One reason is the great disparity in the quantity of available evidence on either side of the point at which pottery enters the archaeological record. But relevant as well is a tendency toward fragmentation * Dates throughout the volume are in calibrated radiocarbon years unless otherwise noted.

1

of inquiry into multiple paths with different theoretical inspiration and little in the way of intersection or cross-communication among them. In this book, the period between 3500 b.c. and 500 b.c. is taken as crucial to understanding the genesis of Mesoamerican civilization, but the focus is narrowed from Mesoamerica as a whole to one par ticular region: the Soconusco, on the Pacific coast of Chiapas, Mexico and adjacent Guatemala. A volume focusing on the early archaeology of the Soconusco but with an eye on the topic of early Mesoamerican social transformations more generally is appropriate for several reasons, but most obviously because recent fieldwork makes the region the bestknown case from the lowlands of Mesoamerica for understanding social change through the entire span of 3500 b.c. through 500 b.c. The emerging picture contrasts in various ways with those for the Southern and Central Highlands, and claims that early developments in the Soconusco were precocious, distinctive, or even unique have entered larger interpretive discussions in Mesoamerican archaeology. Yet much of the evidence on which those claims are based remains only summarily published. The time is ripe for a volume of essays showcasing both the diversity of current work on the early archaeology of the Soconusco and the growing opportunities for synergy as research in multiple subregions yields a richer understanding of the whole. My goal in this introduction is to set the topics and perspectives of the papers that follow into a series of larger contexts. Because an excellent recent review of Southeastern Pacific Coast archaeology is available (Love 2007), I have made no effort here to be comprehensive. Instead, I review the larger importance of work on early Soconusco, introduce the region and its prehistory, and provide a brief overview of the chapters that follow.

THE REGION The Soconusco region is a narrow strip of the Pacific coast of Chiapas, Mexico and adjacent

2

early social transformations

Guatemala (Figures 1.1 and 1.2). Sharply delimited inland by the Sierra Madre Mountains, it extends from around the modern town of Mapastepec, Chiapas southeast to the Tilapa River in Guatemala (Lowe, Lee, and Martínez 1982:55– 62). High rainfall feeds numerous rivers that cross a short coastal plain to feed freshwater marshes and brackish estuaries and lagoons. To the northwest, the Chiapas coast is drier, and indeed there is variation within the Soconusco itself. As one moves to the southeast, the coast widens and rainfall increases. In Guatemala, the southeastern boundary of the Soconusco as a geographical region is marked by an inland extension of the coastal plain, but early settlement related to that of the Soconusco stretched throughout Pacific-coastal Guatemala into El Salvador. One essay included here (Chapter 9) extends coverage of the book into this related area. The Soconusco and adjacent coastal Guatemala are rich in natural resources. Since biotic communities tend to run in narrow strips parallel to the ocean, a range of wild resources would have been readily accessible to ancient inhabitants. Indeed, the sheer richness and diversity of aquatic resources have fi gured prominently in claims that subsistence well into the Formative was focused more on hunting, gathering, and fishing than on agriculture.

THE LARGER IMPORT OF EARLY SOCONUSCO The period 3500 b.c. through 500 b.c. is of interest for studying the genesis of Mesoamerican civilization because it unites a set of deeply transformative developments—but also because scholarly work on those developments ranges across the full theoretical diversity of contemporary anthropological archaeology. Understandings of the transition to the Formative, the emergence of sociopolitical complexity, and the pan-Mesoamerican occurrence of Olmec art are drawn increasingly from evidence found in the Soconusco.

Changing Subsistence From Archaic to Formative The subsistence systems of ancient urban Mesoamerica were the products of a complex history. Of the three key crops—maize, beans, and squash—the staple maize, in par ticular, was deeply interwoven into the economic, social, and symbolic fabric of life (Taube 2000). There were, though, other domesticated plants, including avocado, maguey, and amaranth (McClung de Tapia and Zurita Noguera 2000). The Soconusco, in the fifteenth century a.d., was important as a source of cacao for the distant Aztec capital, and cacao residue has been identified on a pottery sherd from the period of interest here (Powis et al. 2008). There has been ongoing speculation that root crops such as manioc were important Formative-period crops in the Soconusco and on the Gulf Coast (Clark, Pye, and Gosser 2007:28–29; Pool 2007:75). Wild animals everywhere provided sources of protein, but in certain contexts domesticated animals formed an important component of the diet. Dogs were eaten at Early Formative San Lorenzo on the Gulf Coast (Wing 1981) and at Middle Formative La Blanca in the Soconusco (Wake and Harrington 2002). The full story of agricultural origins in Mesoamerica is thus long and complicated. Still, despite some recent efforts to displace it in favor of a subsistence transformation (to maize as a staple) at 1000 b.c., the transition to the Formative stands out as pivotal in terms of demography, subsistence, and lifeways. Debate over the causes of the transition has centered around climatic fluctuations, population growth, and the systems perspective developed by Kent Flannery. In Flannery’s model for the Valley of Oaxaca, the growing productivity of maize in relation to that of wild mesquite eventually triggered removal of mesquite trees to clear the  way for farming on river banks, the most desirable lands of the valley for agriculture (Flannery 1986a:504– 506, 1986b:26–27). Such a shift in priorities among subsistence options could have generated a cascade of other changes,

including greater sedentism and population growth. Recent work has tended to add to, rather than winnow down, the list of potentially relevant variables. Now—to a set of possibilities that included climatic fluctuations, the changing productivity of maize, and localized population dynamics—we might add status competition (Blake et al. 1992a; Clark and Gosser 1995), differences between highland and lowland environments (Piperno and Pearsall 1998), population histories at vast scales (Bellwood 2005), and specific technological innovations (Neff et al. 2006c). More than one of these factors may be relevant for explaining the Formative transition, but it is unlikely that all are equally relevant. One way to weigh their respective importance would be to trace their articulation from Archaic to Formative in different contexts and compare the results. Such an effort is hampered by the chronic low visibility of preceramic sites in Mesoamerica and the consequent rarity of sustained work on Archaic adaptations. Pathbreaking efforts in the semi-arid highland settings of Tehuacán (MacNeish 1964, 1981) and Oaxaca (Flannery 1986c) sparked relatively few comparable efforts in other regions. Voorhies’s (1976, 2004) work over three decades therefore places the Soconusco alongside Tehuacán and Oaxaca as one of a handful of Mesoamerican regions in which a social understanding of Archaic adaptations is possible. Voorhies and colleagues suggest that Archaic inhabitants of the Soconusco were mobile low-level food producers and hunter-gatherer-fishers who visited the estuary to process marsh clams for transport to inland base camps (Kennett, Voorhies, and Martorana 2006; Michaels and Voorhies 1999; Voorhies 2004). A recent addition to this work is a better understanding of how Archaic adaptations changed in the centuries leading up to the transition to the Formative (Chapter 2 in this book). Despite the greater archaeological visibility of sites from the ceramic era, the continued evolution of subsistence systems during the Formative period also needs further work. An

early social transformations

3

BASIN OF MEXICO TEHUACAN VALLEY

Tlatilco

SOUTHERN GULF COAST

Tlapacoya Chalcatzingo MORELOS

La Joya Teopantecuanitlán

La Venta

San Lorenzo San José ISTHMUS OF Mogote TEHUANTEPEC

VALLEY OF OAXACA

Chiapa de Corzo GRIJALVA VALLEY

N 0

100 km SOUTHEASTERN PACIFIC COAST

Cantón Corralito, Ojo de Agua

án at Co

ch Su

t ia

e

Ujuxte ran Na

jo

a al m Sa

Izapa

Paso de la Amada

N

ua ah

te ta

M

Takalik Abaj

Cuauhtémoc La Blanca

SO

El Ujuxte

re ad

a ej Vi te la yo o C

ía ar M

a nd Li cl Es

os av

El Mesak

CO N

US

CO Vidal, Leonidas, Salinas Sinaloa SUC HITE

CEN

TRA

PÉQ UEZ

Ujuxte Nueve Cerros

z Pa

La Nueva Revolorio, Peta, Landa, Medina

María Linda Chiquiuitan

El Carmen

TECOJA TE

L GU ATEM

ALA

COA ST

EASTERN GUATEMALA COAST

EL SALVADOR

100 km

FIGURE 1.1 Map of Mesoamerica, with inset of Southeastern Pacific Coast, showing archaeological sites and regions

important issue is again the degree of variability between regions, particularly during the Early Formative (the second millennium b.c.). The case of the Soconusco has figured prominently in arguments supporting a high degree of variability, with discussion centering on three topics: manioc, aquatic estuary fauna, and maize.

4

early social transformations

Lowe (1975:10– 14) suggested that root crops, specifically manioc, might have been a significant food source in Early Formative Soconusco. The idea was based on a perceived lack (at Altamira, Chiapas) of manos and metates suitable for the grinding of maize and an abundance, instead, of small obsidian chips seemingly suitable for the grater boards used in the

o ler vil No

Pajón a lut ca Vuelta Limón Ca

PAJÓN Chantuto

NW FRINGE

Tlacuachero El Chorro Zapotillo Los Cerritos

ACAPETAHUA

Cerro de las Conchas CANTILEÑA SWAMP

Aquiles Serdán a Co

Cantón Corralito



n

n

e iat ch Su

Ojo de Agua Paso de al Amada Los Álvarez

CHIAPAN SOCONUSCO

Sandoval El Varal

MAZATÁN



ua

h Ca

Izapa Niño

San Martín

njo ra Na

La Blanca

Cuauhtémoc El Ujuxte

CUAUHTÉMOC

GUAMUCHAL SWAMP Salinas la Blanca El Mesak

La Victoria

NARANJO

N

0

10

Mangrove

MANCHÓN

20 km

Cacaluta

River

Los Alvarez

Site

MAZATÁN

Region

GUATEMALAN SOCONUSCO

Swamp

FIGURE 1.2 Map of the Soconusco, showing subregions and archaeological sites

processing of manioc. Direct evidence of manioc continues to be rare or nonex istent at lowland Formative sites (Pool 2007:75), and Clark’s (1981, 1994) use-wear analysis of Early Formative obsidian indicates that the chips were used for a variety of domestic tasks that did not include grating. Clark, Pye, and Gosser (2007:28–29, 31) nevertheless favor keeping open the possibility of manioc as a food source in the earlier Formative. A second suggestion— championed particularly by Blake et al. (1992a, 1992b) as part of a larger argument on maize that I will come to a  moment—is that wild aquatic fauna of the Soconusco estuaries was an important focus of Early Formative subsistence. Neff et al. (2006c)

pick up on that suggestion in their efforts to explain the Formative transition in coastal Guatemala, but there have been relatively few empirical contributions to this topic in the last fifteen years (however, see Lesure, Wake, and Steadman 2009; Wake 2004a, 2004b; Wake and Steadman 2009). Wake and I review the available data from the Mazatán zone in Chapter 4. The third topic is the role of maize in the diets of Formative villagers during the second millennium b.c. Isotopic studies of human bone provide an important source of evidence. Certain aspects of diet can be investigated by measuring ratios of stable isotopes of carbon and nitrogen. Maize in par ticular has an unusually high 13C/12C ratio in comparison with other

early social transformations

5

common plant foods in the Americas, and it is the only cultivated C4 plant. As a result, if maize was a dietary staple, it should have left an isotopic signature in human skeletons. Blake et al. (1992a) found no such signature in the Soconusco during the second millennium b.c. and suggested that agriculture was not particularly important in the Formative transition there (see also Ambrose and Norr 1992; Blake et al. 1992b; Chisholm and Blake 2006; Chisholm, Blake, and Love 1993; Clark and Gosser 1995). Blake and his colleagues, inspired by Hayden’s (1990) ideas on the role of plentiful environments in the emergence of social inequality and food production, have suggested that greater sedentism, demographic expansion, and the adoption of pottery in the Soconusco were the result of fundamentally social processes, such as spiraling status competition. Maize might have been consumed as beer at feasts rather than as a dietary staple. This work in the Soconusco helped bring diversity and variation to the foreground in studies of Formative-period subsistence. Temporal variation is one factor. If maize was originally cultivated for its stalk sugar (Smalley and Blake 2003), then it becomes important to determine the point at which the focus of selection changed from stalk to grain (Webster 2011). There is some evidence to suggest that the shift to maize as a staple may have occurred around 1000 b.c. in multiple parts of Mesoamerica (Clark, Pye, and Gosser 2007; Rosenswig 2006; Smalley and Blake 2003; Webster 2011). Still, regional diversity is an important topic. Was the initial Formative subsistence system in the Soconusco unique, or did it fit into some larger pattern of variation? For instance, was it typical of early adaptations in lowland tropical forests but different from those of the semi-arid highlands (Piperno and Pearsall 1998)? If there were such differences between highlands and lowlands, then why was the transition from Archaic to Formative essentially simultaneous in the Gulf Coast, the Valley of Oaxaca, Morelos, and the southern Basin of Mexico (Lesure 2008)? Evidence on interregional diversity should prove

6

early social transformations

important for evaluating a more sophisticated generation of explanations for the ArchaicFormative transition.

Sociopolitical Complexity A second major social transformation in the period of interest here is the emergence of sociopolitical complexity. With the initial Early Formative period (1900– 1400 b.c.) poorly known in many areas, the sequences of Tehuacán and Oaxaca have again proved influential as rough models of the trajectory that might have been followed in other areas. The sequences in those cases begin early, are relatively complete, and suggest a developmental sequence of gradually increasing complexity that satisfies common sense. Agricultural villages of the initial Early Formative were egalitarian, though in the Valley of Oaxaca San José Mogote was already by far the largest center (Marcus and Flannery 1996:78). In the later Early Formative, a continuum of status differences, increasing investment in public architecture, and a two-tiered settlement hierarchy suggest the emergence of chiefdoms (Marcus and Flannery 1996:96– 110). The subsequent Middle Formative (1000–400 b.c.) witnessed consolidation of the hierarchical system, with population expansion, greater disparities in the distribution of high-status goods, truly monumental public architecture and sculpture, and the emergence of three-tiered settlement hierarchies (Marcus and Flannery 1996:111– 134). In Morelos, the development of settlement patterns around Chalcatzingo from the Amate phase suggests a basically similar trajectory of increasing complexity (Hirth 1987), but the initial Early Formative (before 1400 b.c.) remains poorly understood both there and in the southern Basin of Mexico. By the later Early Formative, there were significant concentrations of people at a scattering of regional centers—some with platform architecture—in both areas, including Tlatilco, Tlapacoya, Chalcatzingo, and San Pablo (Grove 2007; Niederberger 2000). The Tlatilco burials reflect village social relations

of considerable complexity (García Moll el al. 1991; Joyce 1999; Tolstoy 1989). The Middle Formative is characterized by rapid population expansion, the colonization of new areas, and the replication of hierarchical political units everywhere (Grove 2007; Lesure et al. 2006; Sanders, Parsons, and Santley 1979:95– 97; Serra Puche, Lazcano Arce, and de la Torre Mendoza 2004). On the Southern Gulf Coast, the period from 1400 b.c. to 1000 b.c. witnessed the spectacular florescence of San Lorenzo (Cyphers 1999; Cyphers and Di Castro 2009; Symonds, Cyphers, and Lunagómez 2002). The site extended across 700 hectares (ha), and its sustaining area appears to have been integrated into a well- developed settlement hierarchy. Dozens of sculptures incorporate a rich and varied iconography (Coe and Diehl 1980; Cyphers 2004). Arguably, the later Early Formative sociopolitical system here was organized at a qualitatively different scale from coeval developments in the highlands (for example, Clark 2007), particularly if these patterns coalesced earlier (toward 1400 b.c.) rather than later (toward 1000 b.c.). Given all of this, the trajectory of social and political developments on the Gulf Coast during the initial Early Formative is of par ticular importance, but on that issue the data are not yet in. Coe and Diehl (1980) defined three phases earlier than the San Lorenzo phase (Ojochi, Bajío, and Chicharras), but Symonds, Cyphers, and Lunagómez (2002) were able to distinguish only between “Ojochi-Bajío” and “San Lorenzo” in their regional survey. There are earthen constructions pre- dating the San Lorenzo phase (Coe and Diehl 1980:106– 109; Cyphers and Di Castro 2009:23), but they are deeply buried. It is hoped that eventually it will be possible to trace the trajectory of change in sociopolitical institutions and practices on the Gulf Coast throughout the Early Formative—we already have a glimpse of a developing ritual tradition at El Manatí (Ortiz and Rodríguez 2000)—but we are not yet at that point. In this context, the Soconusco Formative sequence has significance simply because its

completeness rivals that of the Valley of Oaxaca. In particular, despite a variety of unresolved issues, it is possible to trace a sequence of changing social and political institutions over the entire Early and Middle Formative beginning from 1900 b.c. Other observations underscore the importance of the Soconusco case. First are the unexpected finds. A two-tiered hierarchy of site sizes appears surprisingly early—in the initial Early Formative in the Mazatán area—and there is supporting evidence for the sort of hierarchy of functions that would indicate a network of chiefdoms by 1650 b.c. (Clark 1991, 1994, 2004). The site of Paso de la Amada has yielded two dramatic discoveries, both also dating to the initial Early Formative. There is an earthen ballcourt of a size that is impressive by later Mesoamerican standards (Hill 1999) and, in Mound 6, the 22-meter-long apsidal Structure 4, with its inset porches and low clay benches or walls (Chapter 5; also Blake 1991; Blake et al. 2006). Such finds support claims of rapid— even precocious— emergence of small-scale hierarchical societies in the Soconusco (Blake and Clark 1999; Clark 2004; Hill and Clark 2001), but they also raise additional questions. Could the large buildings indicate residential patterns and social organization fundamentally different from the nuclear-family focus suggested by the residential architecture of Early and Middle Formative sites in the highlands? This new, relatively complete Early-Middle Formative sequence therefore again raises the issue of variation across Mesoamerica, here in trajectories of emerging sociopolitical complexity. Were patterns of emerging complexity in the Soconusco unique— or perhaps characteristic of lowland or coastal areas? The larger import of the archaeology of early complexity in the Soconusco actually extends beyond the evidence itself. Theoretical understandings of the emergence of social inequality and political complexity in Mesoamerica have been directly influenced by the Soconusco case over the last twenty years, particularly through the work of John Clark and Michael Blake. In a seminal 1994 paper on the

early social transformations

7

“power of prestige,” they proposed a general model for the emergence of hereditary social inequality as the result of competition for prestige among political actors (“aggrandizers”) in zones of resource plenty. They illustrated the model with data from their work at Paso de la Amada and other sites in the Mazatán area of the Soconusco. The impact of that model is evident in recent synthetic treatments of emerging complexity in Mesoamerica (for example, Joyce 2000: chapter 2) and by inclusion of the paper in two collections intended for classroom use (Preucel and Hodder 1996; Smith and Masson 2000). The empirical status of the aggrandizer model in the Soconusco itself is thus of wider interest. Most of us working on Formative topics in the area have shared, absorbed, or responded to Clark and Blake’s emphasis on the unintended consequences of internal political competition. Clark and several colleagues have elaborated and refined the model, adding, among other things, consideration of collectivities to an original version that overemphasized the individual (Blake and Clark 1999; Clark 2004; Clark and Gosser 1995; Clark and Pye 2000; Hill and Clark 2001). Still, puzzles remain. At Paso de la Amada itself, construction of platform-top residences has a complicated history (Chapter 6 in this book; also Lesure 1997). Further, despite dramatic differences in residential architecture, there is no convincing artifactual evidence for economic differences among households in such categories as decoration or size of serving vessels, access to imported obsidian and exotics, and the manufacture of greenstone ornaments (Lesure and Blake 2002). There are hints of differential mortuary treatment, but the assemblage of excavated burials is quite limited (Clark 1991; Love 2007:285). While acknowledging the settlement hierarchy and earthen constructions that would have required coordination of labor, Love (2007:284–286) emphasizes the need for debate concerning how well a model of competitive aggrandizers actually characterizes Soconusco data.

8

early social transformations

The Historical Specificity of Olmec A third major phenomenon of the period of interest—wide dissemination of stylistic horizons from around 1400 b.c.—invites inquiry into the specific historicity of early Mesoamerican civilization. Perhaps it is not so surprising that, despite recurring theoretical tensions between, for instance, agency and evolutionary perspectives, the deepest passions generated by investigations of the period between 3500 b.c. and 500 b.c. surround culture-historical issues raised by the pan-Mesoamerican sharing of Olmec art (for example, Clark 2007; Flannery and Marcus 2000; Neff et al. 2006a, 2006b; Sharer et al. 2006). The term Olmec can by itself trigger intense reactions. Grove (1989) noted problems in the common dual usage of this term as the name for an art style with characteristic subject matter and to denote the culture of the Early- and Middle-Formative inhabitants of  the Southern Gulf Coast (or the inhabitants themselves—”the Olmecs”). Ongoing defi nitional discussion has yielded a variety of more or less nuanced positions, with no version fully laying to rest the tenacious problems identified by Grove (see Clark and Pye 2000; de la Fuente 2008; Flannery and Marcus 1994, 2000; Love and Guernsey 2008; Pohorilenko 2008; Stark 2007). In this book, Gutiérrez (Chapter 7) uses the term Olmec sparingly to refer to Early Formative people of the Gulf Coast, while Morgan (Chapter 9) uses it equally sparingly to refer to a style of material expressions. Rosenswig (Chapter 11), like Morgan, uses it to refer to the style, while Love and Guernsey (Chapter 8), following Clark and Pye (2000:217–218), emphasize instead the network of cultural interaction that they envision having generated the stylistic expressions. Indeed, they would identify multiple “Olmec” styles (Love and Guernsey 2008). Pye, Hodgson, and Clark (Chapter 10) seem to again blur style, culture, and people, but they embrace the notion of multiplicity, so that it becomes possible to identify distinctive “Gulf Olmec” and “Highland Olmec” expressions.

A larger theoretical issue here is how far an  explanation of the rise of Mesoamerican civilization (through 500 b.c.) should venture from generalizing analysis to historical particularism. Should we attribute key innovations to the peoples of specific regions or even  par tic u lar sites? What degree of impact should we accord to specific events or actions— and at what spatiotemporal scales? Recent work in the Soconusco again makes that area an important point of reference in larger discussions. A major issue is the nature of interrelations between the Soconusco and the Southern Gulf Coast during the later Early Formative. Clark and Blake (1989) envisioned two stages of interaction, beginning with local emulation of Gulf Coast practices in the Cherla phase (1400– 1300 b.c.) and culminating in political domination in the subsequent Cuadros phase (1300– 1200 b.c.). Clark has since elaborated on that suggestion and worked to expand the evidential base from the crucial Soconusco sites of Cantón Corralito and Ojo de Agua. He argues for early (c. 1400 b.c.) state or even empire status for the site of San Lorenzo and profound “Olmec” (Gulf Coast) influence throughout Mesoamerica (Clark 1997, 2007; Clark and Pye 2000). Cheetham’s (2009, 2010) careful comparisons of material culture between Cantón Corralito and San Lorenzo help make what originally seemed like inflated claims from Clark— a Gulf Coast colony at the Soconusco site—increasingly convincing. A host of fascinating questions follow concerning the nature of the immigrant group, its relations with local populations in the Soconusco, and the degree of continuing contact with its homeland. Any such discussion would revolve around par ticular events and actions with impact at large spatial scales.

Toward Integration In sum, major topics in the archaeology of the period 3500 b.c. through 500 b.c. include adap-

tive transformations, the rise of social inequality and political complexity, and the widespread sharing of artistic expressions from 1400 b.c. All three topics are important for any larger investigation of the genesis of Mesoamerican civilization, and in each case evidence from the Soconusco is part of larger debates. In each topic, as well, larger synthetic efforts point toward the importance of variation among different parts of Mesoamerica. The completeness of the Soconusco sequence is, by itself, enough to establish the importance of this area for larger efforts to characterize variation and assess its implications. Yet those points do not touch on interrelations among the three topics— or, rather, the chronic problem of the lack of such interrelations. Indeed, research in each tends to proceed largely independently of research in the others. Studies of adaptive change ignore agency and politics. Research on sociopolitical complexity becomes snobbish about politics and ignores subsistence, with its whiff of vulgar materialism. Finally, generalizing work on emergent complexity is impatient with or even embarrassed by the complications of historical specificity, while the latter becomes so self-absorbed that it ignores larger anthropological relevance or so devoted to squabbles over an Olmec “mother culture” that interpretation chases itself around in circles. Essays in this book chip away at those divides: Voorhies and Kennett (Chapter 2) bring gender into the study of Archaic adaptations; Lesure and Wake (Chapter 4) review the first systematic comparison of Formative Soconusco faunal assemblages; Gutiérrez (Chapter 7) explores a new dimension of historicity in the disastrous tropical storms to which the region is subject; and Chapters 10 and 11 grapple with shared stylistic horizons at multiple scales. Still, a more extended consideration of the relations between different theoretical perspectives on early Mesoamerica will help further situate this book within a larger, integrated research effort.

early social transformations

9

SCALE, RHETORIC, AND THE ANALYSIS OF MACROREGIONS In the eighties and nineties, anthropological archaeologists tended to treat the relations between diverse theoretical perspectives as at best competitive and at worst antagonistic. These days there is instead considerable enthusiasm for the idea that multiple perspectives might yield insight into topics as broad as social transformations in early Mesoamerica or the genesis of Mesoamerican civilization. The idea would not be to banish theoretical competition completely but to shift it to another level. In its place, front and center, we would place an effort to elicit synergistic contributions from seemingly disparate theoretical programs. This is not the place for an extended justification of that idea, which admittedly involves numerous challenges. I confine my attention to two important arenas for integrative work: rhetoric and scale. In each case, it becomes clear that theoretical integration will require a suite of complementary research endeavors— endeavors with respect to which this book may be situated as an important category of empirical work. Different perspectives on early Mesoamerica may be theoretically and philosophically antagonistic, but one area of common ground is the shared goal of tracing and explaining  ancient social change. Rhetoric— in this case, the shared prospect of narrating and explaining—therefore emerges as an arena in which interconnections between theories may be explored. To narrate the rise of Mesoamerican civilization over the long term, archaeologists draw on a variety of rhetorical forms. The  developmentalist narratives of processual archaeology, in which the story told is one of ever-increasing complexity from the origins of agriculture to the rise of the state, are readily at hand. Given three decades of critiques of developmentalism from positions ranging from postprocessualist to Darwinian, it is noteworthy how often we still end up telling such stories. Developmentalism proves particularly

10

early social transformations

attractive (seductive?) in Mesoamerica, where the transition from earliest village to urban state unfolded over just 1,500 or 2,000 years (depending on how one defines terms)—rapidly, in comparison with other world areas. Alternatives to developmentalism include the historicist forms long applied to study of early Mesoamerica: long-term continuity, divergence from initial homogeneity, and convergence from initial heterogeneity. These last are hardly mutually exclusive, and they may be built into very different narratives. Divergence from an initial state of homogeneity is posited in mother-culture arguments that envision an “Olmec” heartland on the Gulf Coast as a crucible of Mesoamerican civilization. Flannery and Marcus, to whom such suggestions are anathema, nevertheless chose the same rhetorical form—manifested in the divergence of Otomanguean languages—as a central motif for The Cloud People (1983). The third form, in which heterogeneity is early and interaction leads to convergence, is of obvious interest given the linguistic diversity that characterized Mesoamerica at the time of the Spanish Conquest. In The Cloud People, application of the model of early homogeneity specifically to Otomanguean actually implies initial diversity at a larger scale, between the major language families. Accounts positing early heterogeneity in the period between 3500 b.c. and 500 b.c. include Grove’s (1974, 2007) proposals concerning a Tlatilco culture, Clark and Blake’s (1994) Mokaya for the Soconusco, and the synthetic notion of “sister cultures” as an alternative to an Olmec “mother culture” (Grove 1989; Hammond 1988). A holistic synthesis of social transformations in early Mesoamerica will probably require a variety of rhetorical forms, but the forms prompt different sorts of research endeavor since they face distinct challenges for empirical justification. In par ticular, convergence from initial heterogeneity—if it is to be placed on an equal footing with the others—requires that we resist any urge to assume patterns in one area based on knowledge of another or received wis-

dom concerning the inevitability of complexity. Sustained empirical scrutiny of individual regions—with par ticular attention to the earlier epochs, where evidence tends to be scarce—is essential for the successful identification of early heterogeneity. The Soconusco, with its robust early record that diverges in many ways from well-known highland patterns, is a prime candidate for such work. The empirical requirements that would allow scholars to weigh the importance of convergence from initial heterogeneity to the genesis of Mesoamerican civilization thus provide one of the larger endeavors to which this book contributes. As we consider the relevance of multiple rhetorical forms, a key issue will be scale. A satisfying narrative of early Mesoamerican civilization will probably involve a complex interplay of the forms at different scales. For instance, we might trace divergence from initial homogeneity within macroregions—the Central Highlands, the Southern Highlands, the Southeastern Pacific Coast—but if those units differ considerably early on, a larger-scale comparison of those as units might reveal instances of convergence from initial heterogeneity. Given the pace of societal change between village and state in Mesoamerica, any effort to banish developmentalism completely—attractive as the prospect might be in abstract— seems doomed to failure. A more promising approach would be to test the power of developmentalism with shifts of focus between scales. For example, there is  an unfortunate tendency, in Formative- era regional studies across Mesoamerica, to envision chiefdoms arising, de novo, in any and every area studied. Despite the fact that these occurrences are strung out over 1,500 years, emphasis is on local causation rather than historical accretions or the snowballing effects of interlinkages between regions. Historical analysis at large scales is thus one antidote to facile developmentalism. Another strategy would be to scrutinize trajectories that appear developmental at smaller scales of time. For example, change in sociopolitical organization in the

Soconusco looks developmental when it is resolved in increments of 300 to 500 years; though no single site was dominant for long, complexity increased steadily from initial Early Formative through Late Formative. Resolved at a scale of 100 to 200 years, however, the narrative becomes more complicated and challenging (Chapters 5, 6, and 7). The upshot of all this is that we approach social transformations in early Mesoamerica with uncertainty about relevant scales of space and time. Choice of narrative scale, in other words, requires empirical input. An important area for research is thus exploratory work on the relations between causal processes and the multiple scales on which they might conceivably have played out. Macroregions like the Southeastern Pacific Coast need to be examined for internal structure. Was the transition to the Formative or the emergence of hierarchical societies a widespread shift across this area or were there centers and peripheries? Work in distinct study areas along the Chiapas and Guatemala coast has begun to reach sufficient density that such questions can be answered, and several of the chapters in this book make progress in that area (see Chapters 3, 8, 9, 10, 11, and 12). In sum, a volume of collected essays on the social transformations in the Soconusco between 3500 b.c. and 500 b.c. has relevance well beyond the narrow culture-historical concerns of archaeological work in that region. The period in question encompasses transformative events of major importance in the rise of Mesoamerican civilization—adaptive transformations that culminated in sedentary, agricultural societies; the rise of sociopolitical complexity; and the pan-Mesoamerican dissemination of stylistic horizons from 1400 b.c. Soconusco evidence has proved relevant to larger efforts in each of these cases, and patterns in the Soconusco in each case seem surprising or distinctive, points of great interest to theoretical work increasingly inclined to emphasize the importance of diversity across Mesoamerica. That larger interest in diversity is not idle enthusiasm.

early social transformations

11

Identification of variation— or the lack thereof—is important in the selection of rhetorical form and relevant scale or scales for framing our narratives of early Mesoamerica.

ARCHAEOLOGY OF THE SOUTHEASTERN PACIFIC COAST In the remainder of this chapter, I first review the early settlement history of the Southeastern Pacific Coast macroregion, focusing on the preeminent constituent region for that era, the Soconusco. I then briefly introduce the remaining chapters in this book.

Regional Subdivisions How the Soconusco—and, more generally, the Southeastern Pacific Coast of Mesoamerica— should be divided will be an important concern for any effort to empirically assess the scales at which explanations of social transformations should be framed. In Figure 1.2, the Soconusco “region” is broken down by “zones” more or less the size of river drainages—in other words, the sort of area that is typically the subject of investigation by a single archaeological project. Only zones that have actually seen significant archaeological work are included. In the chapters that follow, the term zone is used when subregions of the Soconusco are referred to by name; however, region is such a useful word that it has not seemed advisable to try to restrict its application to the scale of the Soconusco itself. The northwestern stretch of the Soconusco is known primarily from the work of Pailles (1980) at late Early Formative and Middle Formative mounds in the Pajón estuary (Figure 1.2). The Acapetahua estuary is best known for its Archaic occupation (Voorhies 2004 and Chapter 2 in this book), but Kennett, Voorhies, and McClure (2002) report on the Early Formative site of Los Cerritos. The Cantileña (or Hueyate) Swamp is best known for the Archaic site of Cerro de las Conchas (Voorhies et al. 2002), but it has also been the subject of ongoing investigations by John Clark and John Hodgson. The Mazatán

12

early social transformations

zone has undergone the greatest density of investigations of any subregion, with multiple generations of projects under the general sponsorship of the New World Archaeological Foundation, first under the directorship of Gareth Lowe and subsequently of John Clark. A host of publications too numerous to mention is available, with a strong emphasis on excavations of  Early Formative sites (see Chapters 4, 5, 6, and 7). A survey by Clark is nearing publication, but Clark (2004) and Pye, Hodgson, and Clark (Chapter 10) provide preliminary glimpses. Rosenswig (2007, 2008, 2010) has reported on both survey and excavation in the Cuauhtémoc zone, with occupations from initial Early Formative through Middle Formative and beyond. The Naranjo river drainage has been surveyed by Love (2002) and excavations at Early and Middle Formative sites have been reported by Coe (1961), Coe and Flannery (1967), Love (2002, 2007; also, Chapter 8 in this book), and Shook and Hatch (1979). Survey in the Manchón estuary at the southeastern edge of the Soconusco is reported by Pye (1995), and excavations of Early Formative sites by Pye and Demarest (1991; see also Chapter 10 in this book). In Guatemala, the Soconusco extends only to the western part of the Pacific coast (Figure 1.1). The early archaeology of the Central Guatemalan Coast is known particularly from the work of Arroyo and colleagues (Arroyo 1994, 2004; Arroyo, Neff, and Feathers 2002) in the Suchitepéquez and Tecojate zones. The Eastern Guatemalan Coast has been studied by EstradaBelli (1999) and Morgan (Chapter 9 herein). Patterns of early estuary-focused settlements extended into El Salvador (Arroyo 1995). If the coastal strip from western Chiapas to El Salvador is designated a macroregion—the Southeastern Pacific Coast of Mesoamerica— then an important topic is its internal structure (Chapter 12). Of the contributors to this book, Rosenswig (Chapter 11; see also Rosenswig 2010) develops the most elaborate model of macroregional structure. He regards the southeastern two-thirds or so of the Soconusco as an early “island of complexity” exhibiting greater

cal. B.C.

General Periods

Northwestern Fringe

500

Middle Formative 1000

Encanto

Chiapan Soconusco

Duende/ Escalon

Guatemalan Soconusco

Central Guatemala Coast

Camarelo

Guatalon

Conchas

Sis

Eastern Guatemala Coast

Tamarindo

Conchas Late Dunas

Late Early Formative

1400

Initial Early Formative

Early Dunas

Jocotal

Jocotal

Cuadros Cherla Ocós

Cuadros Cherla? Ocós

Locona

Locona

Barra

Tecojate Coyolate II

Cangrejo Huiscoyol

Coyolate I Madre Vieja

1900

Late Archaic

Late Archaic Chantuto B

Late Archaic

3500

Middle Archaic

Chantuto A

5000

FIGURE 1.3 Chronology of the Soconusco and neighboring coastal Guatemala, c. 5000– 500 b.c. Sources: Arroyo 1994; Arroyo, Neff, and Feathers 2002; Blake et al. 1995; Clark and Cheetham 2005; Love 2007; Lowe 2007:66; Morgan, Chapter 9, this book; Rosenswig, Chapter 11, this book; Voorhies 2004.

sedentism, higher populations, and more elaborate sociopolitical systems during the second millennium b.c. than were present in adjacent areas of the Pacific coast to the northwest (in Chiapas) and the southeast (in Guatemala). Rosenswig emphasizes long-term cultural coherence in the area “between” the massive wetlands of Acapetahua- Cantileña and Guamuchal-Manchón—including, in Figure 1.2, the Mazatán, Cuauhtémoc, and

Naranjo zones. Still, his discussion effectively includes settlement within the wetland areas themselves. Pye, Hodgson, and Clark (Chapter 10) more explicitly include the Cantileña Swamp and Guamuchal-Manchón wetlands in their “greater Mokaya settlement region,” but they exclude the Acapetahua and Pajón zones. Early pottery styles change only gradually as one moves from Manchón to the Central and then to the Eastern Guatemalan Coast (Arroyo

early social transformations

13

1994, 1995; Arroyo, Neff, and Feathers 2002), and over the course of the first millennium b.c. any “island of complexity” status for Soconusco in comparison to adjacent Guatemala gradually erodes. In Figures 1.1 and 1.2, the main “zones” in which early settlement has been studied are indicated. Also, in Figure 1.2, the Soconusco itself is subdivided. I make no special claims for the groupings of zones that I propose— Acapetahua through Cuauhtémoc in a “Chiapan Soconusco,” Naranjo through Manchón in a “Guatemalan Soconusco,” with Pajón in the rather drier “northwestern fringe.” I introduce these designations mainly for the purpose of composing a manageable chronological chart for the macroregion (Figure 1.3). For that purpose, the groupings do seem sensitive to patterns of variation within the Soconusco. The earliest Formative phase, Barra, is documented in Cuauhtémoc, Mazatán, and the Acapetahua estuary. It has not, however, been identified in Guatemalan Soconusco. Likewise, the Cherla phase is attested from Acapetahua through Cuauhtémoc. In Guatemalan Soconusco, its presence is debatable, complicated by the apparently coeval Navarijo complex (Shook and Hatch 1979; see Arroyo, Neff, and Feathers 2002 for discussion). Figure 1.3 provides phase sequences for the Soconusco and Pacific coastal Guatemala. The sequences are complicated enough that it has proved useful to designate archaeological cultures that transcend several phases. Voorhies (1976, 2004) introduced the term Chantuto to refer to the Archaic inhabitants of the Soconusco; Clark and Blake (1989, 1994) proposed the term Mokaya to refer to the subsequent initial Early Formative complexes (Barra through Cherla phases). Clark, in par ticular, would then use the term Olmec for the following Cuadros complex (Clark 1997, 2007; Clark and Blake 1989; Clark and Pye 2000). It is the more general temporal subdivisions in the left-most column of Figure 1.3 (Middle Archaic through Middle Formative) that provide the most convenient structure for a brief sketch of the early culture history of the area.

14

early social transformations

Culture History The Archaic in the Soconusco is known primarily from shell mounds of the Acapetahua estuary and the Cantileña Swamp (Voorhies 2004). There are no shell mounds in Guatemalan Soconusco and the Central Guatemalan Coast, but geomorphological data indicate human presence in the later Archaic (Neff et al. 2006c). By the Late Archaic (from 3500 b.c.), the inhabitants of the Pacific coast were low-level food producers who relied also on a variety of wild foods (Chapters 2 and 3). In the Acapetahua zone, the shell mounds probably derived from bulk production of dried marsh clam meat for transfer to home bases located inland, on the coastal plain (Kennett, Voorhies, and Martorana 2006; Voorhies 2004 ). Little is currently known about social and political organization among Archaic visitors to the estuaries, but the situation could soon change: Clark and Hodgson report finding a preceramic site in the Cantileña Swamp with large, constructed platforms (John Clark and John Hodgson, personal communication, 2009). The very earliest Formative is known as the Barra phase in the Chiapan Soconusco (Lowe 1975). Madre Vieja represents a coeval development on the Central Guatemalan Coast (Arroyo 1994). Exuberant plastic decoration of wellmade pots—particularly flat-based, restrictedrim vessels—is something of a surprise for the initial stages of pottery production. The most likely explanation is that pottery at this time served a relatively narrow range of functions centered on feasting and the pursuit of prestige (Clark and Blake 1994; Clark and Gosser 1995). In Mazatán, there is evidence of Barra occupation at most large sites of the initial Early Formative. It appears that the social contexts of status competition were small, permanent, essentially egalitarian villages (Clark and Pye 2000). The rest of the initial Early Formative, from 1700 b.c. through 1400 b.c., witnessed a population surge in both Chiapan and Guatemalan Soconusco. The best- documented settlements

reached dozens of hectares in extent (Chapters 5 and 6). They were located close to (without actually being in) the estuaries (Chapters 4 and 9). The extent of settlement further inland on the coastal plain may have varied. Love (2002, 2007) identified a balanced distribution of settlements between estuary and inland zones in the Naranjo river drainage. Nothing similar has thus far been reported for Chiapas, though Locona- Ocós sherds did appear in mound fi ll at Izapa (Ekholm 1969). There is much work still to be done on subsistence, but it is known that food production, fishing, and more generalized hunting and gathering were all important. Stylistic similarities along the coast point to intensive interaction and broadly similar developments, probably facilitated by canoe traffic through the wetlands. Still, there was probably structural variation in sociopolitical developments within the Soconusco, with the Mazatán region one of the centers (or perhaps the center) for the initial appearance of hereditary inequality and multicommunity political formations (Clark 2004; Clark and Blake 1994). Emerging sociopolitical entities in the Cuauhtémoc region were not different in kind from those of Mazatán, though somewhat smaller in scale (Chapter 11 and Rosenswig 2008, 2010). The later Early Formative (1400– 1000 b.c.) was marked by significant changes in material culture throughout the Soconusco and to the southeast along the Guatemalan Coast. The ultimate source of inspiration for stylistic and iconographic changes in the area appears to have been the Gulf Coast, and probably particularly the vast site of San Lorenzo (Cheetham 2009, 2010; Clark 2007; Clark and Pye 2000; Lesure 2000). The period 1400 b.c. through 1300 b.c. was transitional between initial- and late-Early Formative complexes and seems to have been a time of localized stylistic diversification (Arroyo, Neff, and Feathers 2002). The subsequent Cuadros phase, though, was an era of renewed homogeneity of ceramic complexes along the Pacific coast and perhaps expansion of population into new areas such as Pajón.

Details on sociopolitical developments are available only for the Mazatán region, and they point to a complex history in which specific events loom large. This is the era in which Gulf Coast immigrants established a colony at Cantón Corralito (Cheetham 2009, 2010; Clark 1997), an event that likely capped decades of political turmoil punctuated by the abandonment of the massive initial Early Formative site of Paso de la Amada. Cantón Corralito was itself rather short-lived, and Gutiérrez (Chapter 7) argues that the site was destroyed by another singular event— a massive flooding of the Coatán River triggered by a hurricane, or perhaps volcanic activity in the Sierra Madres. More generally, Clark’s (1997; Clark and Pye 2000) reconstructions indicate a significant increase in the scale of political units at this time, perhaps a shift from simple to complex chiefdom. The trajectory of change in subsistence over the course of the Early Formative is poorly understood, though changes in grinding-tool technology and quantities of fire- cracked rock indicate a gradual shift from Archaic practices. There is some consensus that a subsistence focus on maize as a staple dates to the beginning of the Middle Formative around 1000 b.c. (Clark, Pye, and Gosser 2007; Rosenswig 2006; Smalley and Blake 2003). In the political realm, the Middle Formative witnessed another increase in scale. It is at this time, with the rise of La Blanca in the Naranjo area (Chapter 8 and Love 2002; Love and Guernsey 2007, 2008), that the scale of sociopolitical integration first definitively extended beyond the boundaries of any single study region. The phenomenal rise of population in the Naranjo zone correlates with the abandonment of Mazatán (to the northwest) and Manchón (to the southeast) to suggest significant population movements. Whether these were coercive or voluntary, and whether they are more reflective of strong centralized power or systemic weaknesses that limited the area that could be effectively administered and controlled, are topics open to debate. The lack of disruption of

early social transformations

15

settlement at Pajón suggests that the northwestern fringe of the Soconusco was outside the La Blanca realm. Still, during the initial Middle Formative— through 800 b.c. or so—the Soconusco was dominated by a single, large political entity. The later Middle Formative was characterized by growing population throughout the region (the Mazatán zone, for instance, was reoccupied) and the rise of multiple large centers. At some point, then—whether before, around, or after 500 b.c. depending on how one defines terms—the Pacific coastal region generally came to be characterized by well- established urban societies (see Love 2007, 2010). Love and Guernsey (Chapter 8) suggest that political transformations in the area involved five successive cycles from Early Formative Paso de la Amada to the Late Formative archaic states at El Ujuxte, Izapa, and Takalik Abaj.

THE CHAPTERS IN THIS BOOK As contributions to the larger agenda of understanding early social transformations in the Soconusco, the chapters in this book divide readily into three broad topics. The three chapters of Part 1 bring a variety of  perspectives to bear on the transition from Archaic to Formative. Voorhies and Kennett have previously noted changes in the subsistence record at the Acapetahua shell mounds toward the end of the Late Archaic: increasing involvement in agriculture accompanied by a shift from dry season to wet season harvesting of marsh clams. Here, in Chapter 2, they explain those changes by proposing that transformations in subsistence systems at the end of the Archaic affected the gendered division of labor, leading to shifts in food procurement by men and women. Blake and Neff, in Chapter 3, assemble the known archaeobotanical data from Archaic through Middle Formative contexts in the Soconusco, tracing the texture of occurrences of different species through time and space. They emphasize the likely importance of tree prod-

16

early social transformations

ucts (avocado, cacao, tropical fruits) to mobile Archaic fisher-hunter-gatherers and find substantial continuity in the archaeobotanical records of estuary locations from Archaic to initial Early Formative. An increasing reliance on annual crops like maize or beans is suspected but still difficult to establish. Though there are continuities in the botanical records at estuary sites from Archaic to Early Formative, there are substantial changes in the overall archaeological records of what were probably special-purpose, seasonally occupied sites. Shell mounds disappear, and sites full of broken tecomates (round-bottomed, neckless jars) appear, the latter contrasting with contemporaneous sites just inland from the estuary, where the repertoire of pottery forms is more diverse. Lesure and Wake, in Chapter 4, propose explaining variation among Early Formative assemblages of faunal remains by referring to adaptive changes that occurred at the Archaic to Formative transition. Part 2 brings together four chapters that advance work on early complex sociopolitical organization in the Soconusco by scrutinizing the archaeological records of important political centers. A recurring issue here is the challenge of grappling with the specificity of events or organizational arrangements at these sites while contributing to generalizing understandings of emerging complexity. Two chapters consider the large, forever surprising, initial Early Formative site of Paso de la Amada. Blake, in Chapter 5, considers the two largest mounds of the site, both of which yielded spectacular finds: a series of superimposed high-status residential structures (“chief’s houses”) in Mound 6 and, in Mound 7, the earliest known Mesoamerican ballcourt. He synthesizes the excavation results and sequences of the two mounds and considers the mutual implications of one for the other. My Chapter 6 complements Chapter 5 by considering small-mound and off-mound excavations at the same site. I frame the chapter as a reaction to Clark’s (2004) reenvisioning of Paso de la Amada as one of the earliest known cere-

monial centers in Mesoamerica. I endorse that assessment and go on to explore the par ticular nature of ceremonialism at the site. The next two chapters in Part 2 consider successively later political centers, the late Early Formative Cantón Corralito and Middle Formative La Blanca. In Chapter 7, Gutiérrez brings his work on the disastrous recent effects of Hurricane Stan to bear on the archaeological record of Cantón Corralito, the proposed Gulf Coast (“Olmec”) colony of Cuadros-phase Mazatán. The result is a new perspective on the potential large-scale impact of singular events—in this case, disasters generated by volcanism or hurricanes— on ancient culture change in the Soconusco. The Cuadros-phase settlement at Cantón Corralito appears to have been destroyed suddenly in a natural calamity rather similar to Hurricane Stan. In Chapter 8, Love and Guernsey report on recent work at the vast Middle Formative site of La Blanca, with its 100 ha ceremonial precinct and substantial surrounding residential area. They consider it the last chiefdom in the Soconusco sequence. Its substantial macroregional impact was the product of several centuries of increase in the scale of political organization, yet evidence for a multitiered system of administration remains weak. Love and Guernsey briefly consider the site’s ceremonial core and evidence for a regional system and then center their attention on household differentiation, finding significant differences in wealth. The chapters in Part 3 deal with emerging complexity and the implications of shared horizon styles; they also contribute to the problem of macroregional synthesis, both within the Soconusco and along the Pacific coast more generally. Morgan, in Chapter 9, presents her work at Chiquiuitan on the Coast of Guatemala. She documents an estuary-focused occupation clearly related to that of the Soconusco but demographically sparser and less complexly organized throughout the Early Formative. She then considers Chiquiuitan in relation to what is known of the early archaeology of the Pacific coast gen-

erally, providing a helpful view from the outside of early developments in the Soconusco itself. In Chapter 10, Pye, Hodgson, and Clark focus on the internal structure of the Soconusco at one crucial period, the Jocotal phase. This is the era between the abandonment of Cantón Corralito and the rise of La Blanca. They see a settlement size hierarchy of multiple tiers and centralized political authority only in the Mazatán area, where the principal site was Ojo de Agua. Their proposed expansion of the Jocotal phase to 200 years (from fifty) has significant implications for larger understandings of the tempo of political change immediately before the emergence of La Blanca. In Chapter 11, Rosenswig, building from his work in the Cuauhtémoc region, grapples with the scale of the basic units of emerging complexity in Early Mesoamerica. On inspiration from Braudel, he envisions “islands” of sedentism and complexity in the Early Formative, separated by large areas still sparsely settled with more mobile peoples. The Soconusco itself would constitute one such island of complexity. He builds his case by emphasizing the internal cultural integrity of the Soconusco and by reviewing available evidence on mobility, complexity, and density of early settlement in adjacent areas. In Chapter 12, I elaborate on the concept of macroregional synthesis, briefly considering its potential contribution to better narratives of early social transformations in Mesoamerica.

REFERENCES Ambrose, Stanley H., and Lynette Norr 1992 On stable isotopic data and prehistoric subsistence in the Soconusco region. Current Anthropology 33: 401–404. Arroyo, Bárbara 1994 The Early Formative in Southern Mesoamerica: An Explanation for the Origins of Sedentary Villages. Unpublished Ph.D. dissertation, Department of Anthropology, Vanderbilt University, Nashville, TN. Arroyo, Bárbara 1995 Early Ceramics From El Salvador: The El Carmen Site. In The Emergence of Pottery

early social transformations

17

Technology and Innovation in Ancient Societies, edited by William K. Barnett and John W. Hoopes, pp. 199–208. Smithsonian Institution Press, Washington, DC. Arroyo, Bárbara 2004 Of Salt and Water: Ancient Commoners on the Pacific Coast of Guatemala. In Ancient Maya Commoners, edited by J. C. Lohse and F. Valdez, pp. 73– 94. University of Texas Press, Austin. Arroyo, Bárbara, Hector Neff, and James Feathers. 2002 The Early Formative Sequence of Pacific Coastal Guatemala. In Incidents of Archaeology in Central America and Yucatán: Essays in Honor of Edwin M. Shook, edited by Michael Love, Marion Popenoe de Hatch, and Héctor L. Escobedo, pp. 35– 50. University Press of America, Lanham, MD. Bellwood, Peter 2005 First Farmers: The Origins of Agricultural Societies. Blackwell, Oxford. Blake, Michael 1991 An Early Formative Chiefdom at Paso de la Amada, Chiapas, Mexico. In The Formation of Complex Society in Southeastern Mesoamerica, edited by W. R. Fowler, Jr., pp. 27–46. CRC Press, Boca Raton, FL. Blake, Michael, Brian Chisholm, John E. Clark, and Karen Mudar 1992a Non-agricultural Staples and Agricultural Supplements: Early Formative Subsistence in the Soconusco Region, Mexico. In Transitions to Agriculture, edited by T. D. Price and A. B. Gebauer, pp. 133– 151. Prehistory Press, Madison, WI. Blake, Michael, Brian Chisholm, Barbara Voorhies, George Michaels, Michael W. Love, and John E. Clark 1992b Prehistoric subsistence in the Soconusco region. Current Anthropology 33: 83– 94. Blake, Michael, and John E. Clark 1999 The Emergence of Hereditary Inequality: The Case of Pacific Coastal Chiapas, Mexico. In Pacific Latin America in Prehistory: The Evolution of Archaic and Formative Cultures, edited by Michael Blake, pp. 55– 73. Washington State University Press, Pullman. Blake, Michael, John E. Clark, George Michaels, Michael W. Love, Mary E. Pye, Arthur A. Demarest, and Bárbara Arroyo 1995 Radiocarbon chronology for the Late Archaic and Formative periods on the Pacific coast of southeastern Mesoamerica. Ancient Mesoamerica 6: 161– 183.

18

early social transformations

Blake, Michael, Richard G. Lesure, John E. Clark, Warren D. Hill, and Luis Barba 2006 The Residence of Power at Paso de la Amada, Mexico. In Palaces and Power in the Americas: From Peru to the Northwest Coast, edited by J. J. Christie and P. J. Sarro, pp. 191–210. University of Texas Press, Austin. Cheetham, David 2009 Early Olmec Figurines From Two Regions: Style as Cultural Imperative. In Mesoamerican Figurines: Small-Scale Indices of LargeScale Social Phenomena, edited by Christina T. Halperin, Katherine A. Faust, Rhonda Taube, and Auore Giguet, pp. 149–179. University Press of Florida, Gainesville. Cheetham, David 2010 Cultural imperatives in clay: Early Olmec carved pottery from San Lorenzo and Cantón Corralito. Ancient Mesoamerica 21(1): 165– 185. Chisholm, Brian, and Michael Blake 2006 Diet in Prehistoric Soconusco. In Histories of Maize: Multidisciplinary Approaches to the Prehistory, Linguistics, Biogeography, Domestication, and Evolution of Maize, edited by John E. Staller, Robert H. Tykot, and Bruce F. Benz, pp. 161– 172. Elsevier Academic Press, Boston. Chisholm, Brian, Michael Blake, and Michael W. Love 1993 More on prehistoric subsistence in the Soconusco region: Response to Ambrose and Norr. Current Anthropology 34: 432–434. Clark, John E. 1981 The Early Preclassic obsidian industry of Paso de la Amada, Chiapas, Mexico. Estudios de la Cultura Maya 13: 265–284. Clark, John E. 1991 The Beginnings of Mesoamerica: Apologia for the Soconusco Early Formative. In The Formation of Complex Society in Southeastern Mesoamerica, edited by W. R. Fowler, Jr., pp. 13–26. CRC Press, Boca Raton, FL. Clark, John E. 1994 The Development of Early Formative Rank Societies in the Soconusco, Chiapas, Mexico. Ph.D. dissertation, Department of Anthropology, University of Michigan, Ann Arbor. Clark, John E. 1997 The arts of government in early Mesoamerica. Annual Review of Anthropology 26: 211–234. Clark, John E. 2004 Mesoamerica Goes Public: Early Ceremonial Centers, Leaders, and Communities.

In Mesoamerican Archaeology, edited by Julia Hendon and Rosemary Joyce, pp. 43– 72. Blackwell, Oxford. Clark, John E. 2007 Mesoamerica’s First State. In The Political Economy of Ancient Mesoamerica: Transformations During the Formative and Classic Periods, edited by Vernon L. Scarborough and John E. Clark, pp. 11–46. University of New Mexico Press, Albuquerque. Clark, John E., and Michael Blake 1989 El origen de la civilización en Mesoamérica: Los olmecas y mokaya del Soconusco de Chiapas, México. In El Preclásico o Formativo: Avances y Perspectivas, Seminario de Arqueología “Dr. Roman Piña Chan,” edited by Martha Carmona Macias, pp. 385–403. Instituto Nacional de Antropología e Historia, Mexico City. Clark, John E., and Michael Blake 1994 The Power of Prestige: Competitive Generosity and the Emergence of Rank Societies in Lowland Mesoamerica. In Factional Competition and Politi cal Development in the New World, edited by E. M. Brumfiel and J. Fox, pp. 17– 30. Cambridge University Press, England. Clark, John E., and David Cheetham 2005 Cerámica del Formativo de Chiapas. In La Producción Alfarera en el México Antiguo I, edited by Breatriz Leonor Merino Carrión and Angel García Cook, pp. 285–433. Instituto Nacional de Antropología e Historia, Mexico City. Clark, John E., and Dennis Gosser 1995 Reinventing Mesoamerica’s First Pottery. In The Emergence of Pottery: Technology and Innovation in Ancient Societies, edited by W. Barnett and J. Hoopes, pp. 209–221. Smithsonian Institution Press, Washington, DC. Clark, John E., and Mary E. Pye 2000 The Pacific Coast and the Olmec Question. In Olmec Art and Archaeology in Mesoamerica, edited by J. E. Clark and M. E. Pye, pp. 217–251. Studies in the History of Art No. 58, National Gallery of Art, Washington, DC. Clark, John E., Mary E. Pye, and Dennis C. Gosser 2007 Thermolithics and Corn Dependency in Mesoamerica. In Archaeology, Art, and Ethnogenesis in Mesoamerican Prehistory: Papers in Honor of Gareth W. Lowe, edited by Lynneth S. Lowe and Mary E. Pye, pp. 23–42. Papers of the New World Archaeo-

logical Foundation No. 68, Brigham Young University, Provo, UT. Coe, Michael D. 1961 La Victoria, an Early Site on the Pacific Coast of Guatemala. Papers of the Peabody Museum of Archaeology and Ethnology No. 53, Peabody Museum, Harvard University, Cambridge, MA. Coe, Michael D., and Richard A. Diehl 1980 In the Land of the Olmec, vol. 1, The Archaeology of San Lorenzo Tenochtitlán. University of Texas Press, Austin. Coe, Michael D., and Kent V. Flannery 1967 Early Cultures and Human Ecology in South Coastal Guatemala. Smithsonian Contributions to Archaeology No. 3, Smithsonian Institution, Washington, DC. Cyphers, Ann 1999 From Stone to Symbols: Olmec Art in Social Context at San Lorenzo Tenochtitlan. In Social Patterns in Pre- classic Mesoamerica, edited by David C. Grove and Rosemary A. Joyce, pp. 155–181. Dumbarton Oaks Research Library and Collection, Washington, DC. Cyphers, Ann 2004 Escultura Olmeca de San Lorenzo Tenochtitlán. Instituto de Investigaciones Antropológicas, Universidad Nacional Autónoma de México, Mexico City. Cyphers, Ann, and Anna Di Castro 2009 Early Olmec Architecture and Imagery. In The Art of Urbanism: How Mesoamerican Kingdoms Represented Themselves in Architecture and Imagery, edited by William L. Fash and Leonardo López Luján, pp. 21– 51. Dumbarton Oaks Research Library and Collection, Washington, DC. de la Fuente, Beatriz 2008 ¿Puede un estilo definir una cultura? In Olmeca: Balance y Perspectivas. Memoria de la Primera Mesa Redonda, edited by María Teresa Uriarte and Rebecca B. González Lauck, pp. 25–37. Instituto de Investigaciones Estéticas, Universidad Nacional Autónoma de México, Mexico City. Ekholm, Suzanne M. 1969 Mound 30a and the Early Preclassic Sequence of Izapa, Chiapas, Mexico. Papers of the New World Archaeological Foundation No. 25, Brigham Young University, Provo, UT. Estrada-Belli, Francisco 1999 The Archaeology of Complex Societies in Southeastern Pacific Coastal Guatemala: A Regional GIS Approach. British Archaeological Reports, Oxford.

early social transformations

19

Flannery, K. V., and J. Marcus 1994 Early Formative Pottery of the Valley of Oaxaca, Mexico. Memoirs of the University of Michigan Museum of Anthropology No. 27, Ann Arbor. Flannery, Kent V. 1986a Adaptation, Evolution, and Archaeological Phases: Some Implications of Reynolds’ Simulation. In Guilá Naquitz: Archaic Foraging and Early Agriculture and Oaxaca, Mexico, edited by Kent V. Flannery, pp. 501– 507. Academic Press, Orlando, FL. Flannery, Kent V. 1986b Ecosystem Models and Information Flow in the Tehuacán- Oaxaca Region. In Guilá Naquitz: Archaic Foraging and Early Agriculture and Oaxaca, Mexico, edited by Kent V. Flannery, pp. 19–28. Academic Press, Orlando, FL. Flannery, Kent V. (ed.) 1986c Guilá Naquitz: Archaic Foraging and Early Agriculture and Oaxaca, Mexico. Academic Press, Orlando, FL. Flannery, Kent V., and Joyce Marcus 1983 The Cloud People: Divergent Evolution of the Zapotec and Mixtec Civilizations. Academic Press, New York. Flannery, Kent V., and Joyce Marcus 2000 Formative Mexican chiefdoms and the myth of the “mother culture.” Journal of Anthropological Archaeology 19: 1–37. García Moll, Roberto, Daniel Juárez Cossio, Carmen Pijoan Aguade, Maria Elena Salas Cuesta, and Marcela Salas Cuesta 1991 Catálogo de entierros de San Luis Tlatilco, México, Temporada IV. Serie Antropología Física-Arqueología. Instituto Nacional de Antropología e Historia, Mexico City. Grove, David C. 1974 The Highland Olmec Manifestation: A Consideration of What It Is and Isn’t. In Mesoamerican Archaeology: New Approaches, edited by Norman Hammond, pp. 109– 128. Duckworth, London. Grove, David C. 1989 Olmec: What’s in a Name? In Regional Perspectives on the Olmec, edited by Robert J. Sharer and David C. Grove, pp. 8– 14. Cambridge University Press, New York. Grove, David C. 2007 Stirrup-Spout Bottles and Carved Stone Monuments: The Many Faces of Interregional Interactions in Formative Period Morelos. In Archaeology, Art, and Ethnogenesis in Mesoamerican Prehistory: Papers in

20

early social transformations

Honor of Gareth W. Lowe, edited by Lynneth S. Lowe and Mary E. Pye, pp. 209–227. Papers of the New World Archaeological Foundation No. 68, Brigham Young University, Provo, UT. Hammond, Norman 1988 Cultura hermana: Reappraising the Olmec. Quarterly Review of Archaeology 9(4): 1–4. Hayden, Brian 1990 Nimrods, piscators, pluckers, and planters: The emergence of food production. Journal of Anthropological Archaeology 9: 31– 69. Hill, Warren D. 1999 Ballcourts, Competitive Games, and the Emergence of Complex Society. Ph.D. dissertation, University of British Columbia, Vancouver, Canada. Hill, Warren D., and John E. Clark 2001 Sports, gambling, and government: America’s first social compact? American Anthropologist 103: 331–345. Hirth, Kenneth G. 1987 Formative Period Settlement Patterns in the Río Amatzinac Valley. In Ancient Chalcatzingo, edited by David C. Grove, pp. 56– 62. University of Texas Press, Austin. Joyce, Rosemary A. 1999 Social Dimensions of Pre- classic Burials. In Social Patterns in Pre- classic Mesoamerica, edited by David C. Grove and Rosemary A. Joyce, pp. 15–48. Dumbarton Oaks Research Library and Collection, Washington, DC. Joyce, Rosemary A. 2000 Gender and Power in Prehispanic Mesoamerica. University of Texas Press, Austin. Kennett, Douglas J., Barbara Voorhies, and D. Martorana 2006 An Ecological Model for the Origins of Maize-Based Food Production on the Pacific Coast of Southern Mexico. In Behavioral Ecology and the Transition to Agriculture, edited by D. J. Kennett and B. Winterhalder, pp. 103– 136. University of California Press, Berkeley. Kennett, Douglas J., Barbara Voorhies, and S. B. McClure 2002 Los Cerritos: An early fishing-farming community on the Pacific Coast of Mexico. Antiquity 76: 631– 632. Lesure, Richard G. 1997 Early Formative platforms at Paso de La Amada, Chiapas, Mexico. Latin American Antiquity 8: 217–235.

Lesure, Richard G. 2000 Animal Imagery, Cultural Unities, and Ideologies of Inequality in Early Formative Mesoamerica. In Olmec Art and Archaeology in Mesoamerica, edited by J. E. Clark and M. E. Pye, pp. 193–215. Studies in the History of Art No. 58, National Gallery of Art, Washington, DC. Lesure, Richard G. 2008 The Neolithic Demographic Transition in Mesoamerica? Larger Implications of the Strategy of Relative Chronology. In The Neolithic Demographic Transition and Its Consequences, edited by Jean-Pierre Bocquet-Appel and Ofer Bar Yosef, pp.107– 138. Springer Science, New York. Lesure, Richard G., and Michael Blake 2002 Interpretive challenges in the study of early complexity: Economy, ritual and architecture at Paso de la Amada, Mexico. Journal of Anthropological Archaeology 21: 1–24. Lesure, Richard G., Aleksander Borejsza, Jennifer Carballo, Charles Frederick, Virginia Popper, and Thomas A. Wake 2006 Chronology, subsistence and the earliest Formative of central Tlaxcala, Mexico. Latin American Antiquity 17(4): 474–492. Lesure, Richard G., Thomas A. Wake, and David W. Steadman 2009 Subsistence in the Estuary: Surplus Production, Expedient Meals, or Something Between? In Settlement and Subsistence in Early Formative Soconusco: El Varal and the Problem of Inter-Site Assemblage Variation, edited by Richard G. Lesure, pp. 203–222. Cotsen Institute of Archaeology Press, University of California, Los Angeles. Love, Michael 2007 Recent research in the Southern Highlands and Pacific coast of Mesoamerica. Journal of Archaeological Research 15: 275–328. Love, Michael 2010 Thinking Outside the Plaza: Varieties of Preclassic Sculpture in Pacific Guatemala and Their Political Significance. In The Place of Stone Monuments in Mesoamerica’s Preclassic Transition: Context, Use and Meaning, edited by Julia Guernsey, John E. Clark, and Bárbara Arroyo, pp. 149– 175. Dumbarton Oaks Research Library and Collection, Washington, DC. Love, Michael W. 2002 Early Complex Society in Pacific Guatemala: Settlements and Chronology of the Río Naranjo, Guatemala. Papers of the New

World Archaeological Foundation No. 66, Brigham Young University, Provo, UT. Love, Michael W., and Julia Guernsey 2007 Monument 3 from La Blanca, Guatemala: A Middle Preclassic earthen sculpture and its ritual associations. Antiquity 81: 920– 932. Love, Michael W., and Julia Guernsey 2008 Sociedad y estilo en la costa del Pacífico en el Preclásico medio. In Olmeca: Balance y Perspectivas. Memoria de la Primera Mesa Redonda, edited by María Teresa Uriarte and Rebecca B. González Lauck, pp. 89– 111. Instituto de Investigaciones Estéticas, Universidad Nacional Autónoma de México, Mexico City. Lowe, Gareth W. 1975 The Early Preclassic Barra Phase of Altamira, Chiapas: A Review of the New Data. Papers of the New World Archaeological Foundation No. 38, Brigham Young University, Provo, UT. Lowe, Gareth W. 2007 Early Formative Chiapas: The Beginnings of Civilization in the Central Depression of Chiapas. In Archaeology, Art, and Ethnogenesis in Mesoamerican Prehistory: Papers in Honor of Gareth W. Lowe, edited by Lynneth S. Lowe and Mary E. Pye, pp. 63– 108. Papers of the New World Archaeological Foundation No. 68, Brigham Young University, Provo, UT. Lowe, Gareth W., Thomas A. Lee, Jr., and E. Martínez 1982 Izapa: An Introduction to the Ruins and Monuments. Papers of the New World Archaeological Foundation No. 31, Brigham Young University, Provo, UT. MacNeish, Richard S. 1964 Ancient Mesoamerican civilization. Science 143(3606): 531– 537. MacNeish, Richard S. 1981 Tehuacan’s accomplishments. In Supplement to the Handbook of Middle American Indians, vol. 1, Archaeology, edited by Victoria Reifler Bricker (General Editor) and Jeremy A. Sabloff (Volume Editor), pp. 31–47. University of Texas Press, Austin. Marcus, Joyce, and Kent V. Flannery 1996 Zapotec Civilization: How Urban Society Evolved in Mexico’s Oaxaca Valley. Thames and Hudson, London. McClung de Tapia, Emily, and Judith Zurita Noguera

early social transformations

21

2000

Las primeras sociedades sedentarias. In Historia Antigua de México, vol. 1, El México Antiguo, sus Àreas Culturales, los Orígenes y el Horizonte Preclásico, edited by Linda Manzanilla and Leonardo López Luján, pp.255–295. Instituto Nacional de Antropología e Historia, Mexico City. Michaels, George H., and Barbara Voorhies 1999 Late Archaic Period Coastal Collectors in Southern Mesoamerica: The Chantuto People Revisited. In Pacifi c Latin America in Prehistory: The Evolution of Archaic and Formative Cultures, edited by Michael Blake, pp. 39– 54. Washington State University Press, Pullman. Neff, Hector, Jeff rey Blomster, Michael D. Glascock, Ronald L. Bishop, and M. James Blackman 2006a Methodological issues in the provenance investigation of Early Formative Mesoamerican ceramics. Latin American Antiquity 17(1): 54– 76. Neff, Hector, Jeff rey Blomster, Michael D. Glascock, Ronald L. Bishop, and M. James Blackman 2006b Smokescreens in the provenance investigation of early formative Mesoamerican ceramics. Latin American Antiquity 17(1): 90– 103. Neff, Hector, Deborah M. Pearsall, John G. Jones, Barbara Arroyo, Shawn K. Collins, and Dorothy E. Freidel 2006c Early Maya adaptive patterns: Mid-Late Holocene paleoenvironmental evidence from Pacific Guatemala. Latin American Antiquity 17(3): 287–315. Niederberger, Christine 2000 Ranked Societies, Iconographic Complexity, and Economic Wealth in the Basin of Mexico Toward 1200 bc. In Olmec Art and Archaeology in Mesoamerica, edited by J. E. Clark and M. E. Pye, pp. 169– 188. National Gallery of Art, Washington, DC. Ortiz, Ponciano, and María del Carmen Rodríguez 2000 The Sacred Hill of El Manatí: A Preliminary Discussion of the Site’s Ritual Paraphernalia. In Olmec Art and Archaeology in Mesoamerica, edited by J. E. Clark and M. E. Pye, pp. 75– 93. Studies in the History of Art No. 58, National Gallery of Art, Washington, DC. Pailles, Maricruz 1980 Pampa El Pajón, an Early Middle Preclassic Estuarine Site, Chiapas, Mexico. Papers of the New World Archaeological Foundation No. 44, Brigham Young University, Provo, UT.

22

early social transformations

Piperno, Dolores R., and Deborah M. Pearsall 1998 The Origins of Agriculture in the Lowland Neotropics. Academic Press, San Diego, CA. Pohorilenko, Anatole 2008 Cultura y estilo en el arte olmeca: ¿un estilo, muchas culturas? In Olmeca: Balance y Perspectivas. Memoria de la Primera Mesa Redonda, edited by María Teresa Uriarte and Rebecca B. González Lauck, pp. 65– 87. Instituto de Investigaciones Estéticas, Universidad Nacional Autónoma de México, Mexico City. Pool, Christopher A. 2007 Olmec Archaeology and Early Mesoamerica. Cambridge University Press, New York. Powis, Terry G., Jeff rey W. Hurst, María del Carmen Rodríguez de Mora, Ponciano Ortiz C., Michael Blake, David Cheetham, Michael D. Coe, and John G. Hodgson 2008 The origins of cacao use in Mesoamerica. Mexicon 30(2): 35–38. Preucel, Robert, and Ian Hodder 1996 Contemporary Archaeology in Theory. Blackwell, Oxford. Pye, Mary E. 1995 Settlement, Specialization, and Adaptation in the Río Jesus Drainage, Retalhuleu, Guatemala. Ph.D. dissertation, Department of Anthropology, Vanderbilt University, Nashville, TN. Pye, Mary E., and Arthur A. Demarest 1991 The Evolution of Complex Societies in Southeastern Mesoamerica: New Evidence From El Mesak, Guatemala. In The Formation of Complex Society in Southeastern Mesoamerica, edited by W. L. Fowler, Jr., pp. 77– 100. CRC Press, Boca Raton, FL. Rosenswig, Robert M. 2006 Sedentism and food production in early complex societies of the Soconusco, Mexico. World Archaeology 38: 329–354. Rosenswig, Robert M. 2007 Beyond identifying elites: Feasting as a means to understand early Middle Formative society on the Pacific Coast of Mexico. Journal of Anthropological Archaeology 26: 1–27. Rosenswig, Robert M. 2008 Prehispanic settlement in the Cuauhtémoc region of the Soconusco, Chiapas, Mexico. Journal of Field Archaeology 33(4): 389–411. Rosenswig, Robert M. 2010 The Beginnings of Mesoamerican Civilization: Inter-Regional Interaction and the Olmec. Cambridge University Press, New York.

Sanders, William T., Jeff rey R. Parsons, and Robert S. Santley 1979 The Basin of Mexico: Ecological Processes in the Evolution of a Civilization. Academic Press, New York. Serra Puche, Mari Carmen, Jesús Carlos Lazcano Arce, and Manuel de la Torre Mendoza 2004 Cerámica de Xochitécatl. Instituto de Investigaciones Antropológicas, Universidad Nacional Autónoma de México, Mexico City. Sharer, Robert J., Andrew K. Balkansky, James H. Burton, Gary M. Feinman, Kent V. Flannery, David C. Grove, Joyce Marcus, Robert G. Moyle, T. Douglas Price, Elsa M. Redmond, Robert G. Reynolds, Prudence M. Rice, Charles S. Spencer, James B. Stoltman, and Jason Yaeger 2006 On the logic of archaeological inference: Early Formative pottery and the evolution of Mesoamerican societies. Latin American Antiquity 17(1): 90– 103. Shook, Edwin M., and Marion Hatch 1979 The Early Preclassic sequence in the Ocós/ Salinas La Blanca area, South Coast Guatemala. Contributions of the University of California Archaeological Research Facility 41: 143– 195. Smalley, John, and Michael Blake 2003 Sweet beginnings: Stalk sugar and the domestication of maize. Current Anthropology 44: 675– 703. Smith, Michael E., and Marilyn A. Masson 2000 The Ancient Civilizations of Mesoamerica: A Reader. Blackwell, Oxford. Stark, Barbara L. 2007 Out of Olmec. In The Political Economy of Ancient Mesoamerica: Transformations During the Formative and Classic Periods, edited by Vernon L. Scarborough and John E. Clark, pp. 47– 63. University of New Mexico Press, Albuquerque. Symonds, Stacey, Ann Cyphers, and Roberto Lunagómez 2002 Asentamiento Prehispánico en San Lorenzo Tenochtitlán. Instituto de Investigaciones Antropológicas, Universidad Nacional Autonoma de México, Mexico City. Taube, Karl A. 2000 Lightning Celts and Corn Fetishes: The Formative Olmec and the Development of Maize Symbolism in Mesoamerica and the American Southwest. In Olmec Art and Archaeology in Mesoamerica, edited by John E. Clark and Mary E. Pye, pp. 297–337. Studies in the History of Art No. 58, National Gallery of Art, Washington, DC.

Tolstoy, Paul 1989 Coapexco and Tlatilco: Sites With Olmec Materials in the Basin of Mexico. In Regional Perspectives on the Olmec, edited by R. J. Sharer and D. C. Grove, pp. 85– 121. Cambridge University Press, New York. Voorhies, Barbara 1976 The Chantuto People: An Archaic Period Society of the Chiapas Littoral, Mexico. Papers of the New World Archaeological Foundation No. 41, Brigham Young University, Provo, UT. Voorhies, Barbara 2004 Coastal Collectors in the Holocene: The Chantuto People of Southwest Mexico. University Press of Florida, Gainesville. Voorhies, Barbara, Douglas J. Kennett, J. G. Jones, and Thomas A. Wake 2002 A Middle Archaic archaeological site on the West Coast of Mexico. Latin American Antiquity 13: 179–200. Wake, Thomas A. 2004a A Vertebrate Archaeofauna From the Early Formative Period Site of Paso de la Amada, Chiapas, Mexico: Preliminary Results. In Maya Zooarchaeology: New Directions in Method and Theory, edited by Kitty F. Emery, pp. 209–222. Cotsen Institute of Archaeology Press, University of California, Los Angeles. Wake, Thomas A. 2004b Prehistoric Fisheries of the Greater Soconusco Region, Pacific Coastal Chiapas and Adjacent Guatemala. In Presencia de la Arqueoictiología en México, edited by A. F. Guzmán, O. J. Polaco, and F. J. Aguilar, pp. 161– 166. Proceedings of the Twelfth Meeting of the Fish Remains Working Group of the International Council for Archaeozoology, Instituto Nacional de Antropología e Historia e Museo de Paleontología de Guadalajara “Federico A. Solórzano Barreto,” México. Wake, Thomas A., and Lady R. Harrington. 2002 Appendix 2: Vertebrate Faunal Remains From La Blanca, Guatemala. In Early Complex Society in Pacific Guatemala: Settlements and Chronology of the Rio Naranjo, Guatemala, edited by M. W. Love, pp. 237–252. Papers of the New World Archaeological Foundation No. 66, Brigham Young University, Provo, UT. Wake, Thomas A., and David W. Steadman 2009 Fishing in the Mangroves at Formative Period El Varal. In Settlement and

early social transformations

23

Subsistence in Early Formative Soconusco: El Varal and the Problem of Inter-Site Assemblage Variation, edited by Richard G. Lesure, pp. 99– 112. Cotsen Institute of Archaeology Press, University of California, Los Angeles. Webster, David 2011 Backward bottlenecks: Ancient teosinte/ maize selection. Current Anthropology 52(1): 77– 104.

24

early social transformations

Wing, E. S. 1981 A Comparison of Olmec and Maya Food Ways. In The Olmec and Their Neighbors, edited by E. P. Benson, pp. 20–28. Dumbarton Oaks Research Collection and Library, Washington, DC.

part one

Archaic to Formative Transformations in Subsistence

TWO

A Gender-Based Model for Changes in Subsistence and Mobility During the Terminal Late Archaic Period on the Coast of Chiapas, Mexico Barbara Voorhies and Douglas J. Kennett

O

ur purpose in this chapter is to

investigate in detail the changes in subsistence and settlement mobility manifested over four millennia by the prehistoric Chantuto people. These ancient people occupied what is now the south Pacific coast of Mexico from approximately 7,500 to 4,000 years ago (Voorhies 2004:14). They are best known for the large, highly visible shell mounds that they formed within the wetlands along the outer coast, and this site type is the most thoroughly investigated of what we have inferred to be a settlement system comprising multiple site types formed by residentially mobile people (Voorhies 2004). The basic thesis of this chapter is that during the final phase (around 2700–2000 b.c.) of the Chantuto people’s adaptation to coastal resources that resulted in the formation of substantial shell mounds, women continued to make forays into the coastal wetlands to procure clams and small fish while men concentrated their activities on the inner coast rather than undertaking fishing expeditions to the coastal lagoons as they did in earlier times. This interpretation of the archaeological evidence rests on an analogy with the

Lacandón Maya, a recent historic people who occupy an ecological niche similar to that reconstructed for the Chantuto people. Ethnographic data for gender1 distinctions in seasonal mobility and resource procurement among the Lacandón provide a strong explanatory framework for the archaeological record as currently known for the Chantuto people.

LIFEWAYS OF THE CHANTUTO PEOPLE The fact that the lifeways of the Chantuto people are known principally to archaeologists from investigations of shell mounds inevitably makes it tempting to infer that these sites, formed initially as islands within the wetlands, were once occupied permanently. This seductive inference is also encouraged by the sheer size of these features. The largest shell mound of the six that we have studied, for instance, has a diameter of 160 m and is presently 11 m high.2 Such a substantial accumulation of deposits, principally shell, might seem to represent a sustained effort of the sort often expended by permanently settled peoples.

27

However tempting it may seem to infer permanent residency at the shell mound sites, the actual archaeological record suggests a very different interpretation. Most notable is the fact that the stratified deposits of shell that constitute the matrix of the deposits formed by the Chantuto people lack all of the expected archaeological signatures of permanent settlements. As we have discussed previously (for example, Kennett, Voorhies, and Martorana 2006; Voorhies 2004), consideration of both the site structure and site contents leads to the inference that the shell mound sites were not locations where the Chantuto people were permanently residing. The wellbedded stratigraphy indicates that the processes of deposition were repetitive and that postdepositional processes of disturbance were minimal. Our interpretation is that the repeated couplets of beds of whole shell valves and shell gravel are the archaeological signatures of ancient cooking practices principally for small marsh clams, but also for fish and probably shrimp. Permanently settled people ordinarily produce different sediment, usually consisting of a soil that has high organic content. In addition, if people constantly walk over well-bedded sediments, those layers would become disturbed rather than retain the undisturbed and uniform horizontal bedding that is encountered at the Chantuto shell mounds. Moreover, the shell mounds significantly lack the features that are expected at residential sites, such as signatures of permanent buildings (postholes, rock alignments, and so forth); fire pits, hearths, or both; refuse middens containing a diversity of debris; and burials. Admittedly, one singular discovery at the Tlacuachero shell mound (see Voorhies 2004:51ff ) does have many of these features, but we interpret this as a unique signature of changing lifeways rather than being typical of the overall processes of site formation practiced at the shell mounds by the Chantuto people throughout the entire Middle and Late Archaic periods.* Finally, the nota-

* See Figure 1.2 for site locations and Figure 1.3 for regional chronological chart. All dates are in calendar years.

28

archaic to formative

ble scarcity of artifacts in general and their extremely low diversity are not compatible with the expectations at a site where multiple activities are carried out, as at a residential base. Rather, this is what would be expected if a limited range of activities was being pursued at a location dedicated to specialized activities only. These considerations are spelled out more thoroughly in Voorhies (2004). If the shell mounds were processing stations for aquatic resources and not residential locations, where exactly did the people live? We propose that they lived farther inland on the coastal plain and that we have explored one such apparent base camp that we named Vuelta Limón. This site is coeval in time with the terminal dates from shell mounds located farther downstream in the wetlands in Acapetahua, the municipality where the Late Archaic Period shell mounds are located. Although our limited excavations did not produce evidence of many features usually associated with a settled lifestyle, such as permanent structures or hearths, we did find a dense concentration of discarded artifacts consisting of more tool types than in the shell mounds, along with abundant fire-cracked rock. This feature appears to be a refuse midden or dump. Such a wide diversity and quantity of artifacts is precisely what is expected for a residential base. We suspect that if we were to conduct further excavations at this site we would find evidence of residential features. Accordingly, the evidence for the Chantuto people being logistical foragers (collectors, in Binford’s 1980 terminology) or semi-mobile horticulturalists (or possibly even slash and burn agriculturalists as suggested by Kennett et al. 2010 and Neff et al. 2006) seems to us compelling. The known archaeological record, as we have mentioned, spans four millennia, and it is now possible to detect subtle shifts in the record that reflect temporal changes in the basic adaptation that we have described. The first shift to become apparent to us was in the  record of the seasonal collection of marsh clams that was studied via oxygen isotopes at

the Late Archaic Period Tlacuachero shell mound (Kennett and Voorhies 1995, 1996). This was later complemented by an additional set of isotopic data from the Middle Archaic Period Cerro de las Conchas shell mound (Voorhies et al. 2002). A discussion of the methods for studying season of death using oxygen isotopes may be found elsewhere (for example, Kennett and Voorhies 1996) but the basic principle is that the ratio of O16 to O18 changes predictably with salinity changes in this aquatic environment and registers in the incremental growth of marsh clams. Salinity changes, in turn, are linked to rainfall patterns, which on the coast of Chiapas are strongly bimodal. Accordingly, the ratio of O16 to O18 detected analytically at the margin of an individual shell reflects the salinity of the individual’s environment just before the time of death, which in turn is used to estimate the season of harvesting. The pattern that emerged from this study is that throughout the Middle Archaic Period record and during the early part of the Late Archaic Period record, the Chantuto people collected clams throughout the year and had a slight preference for dry season collection (Voorhies 2004:fi gure 3.13). However, contrary to our initial expectations that are based on the seasonal availability of resources, during the later part of the Late Archaic Period, dry season clam harvesting diminished until fi nally clams were collected only during the wet season. Although we have long suspected that this change in seasonality of clam harvesting resulted from scheduling confl icts associated with an increasing commitment to an agricultural production on the adjacent coastal plain (see Kennett et al. 2006), exactly how and why this transformation took place initially eluded us. Once we detected seasonality changes over time in the archaeological record at Tlacuachero, we recognized discernible changes in other aspects of the site’s contents. An early discovery in the upper Archaic deposits at Tlacuachero is a constructed floor3 with features suggestive of

a settled lifestyle (Stark 1981; Voorhies 1976, 2004:51ff ), but it was only after the seasonality data became available that we fully appreciated its significance. This singular complex of features dates to late in the Late Archaic Period occupation. Among the discovered features were postholes, imprints from circles of rocks, possible traces of small shelters or windbreaks, and a human burial (Voorhies 2004:figure. 2.14). This is the earliest archaeological indication of substantial labor investment in infrastructural features that we have found in the shell mounds, or in fact in the region. As an indication of the amount of labor invested just to bring the construction sediment to the site, Voorhies calculated that between 7,500 and 8,650 loads of material, thought to be clay, were carted to the shell mound, depending on the actual shape of the floor.4 This suggests to us that the construction probably was beyond the capabilities of a small family (Voorhies 2004:54) and was more likely the result of a communal effort. What also became more apparent from our accumulating data is that artifact diversity is greater above the constructed floors at Tlacuachero than below them. The artifact density probably increases concomitantly, although the density numbers are so low that they are not statistically reliable. The increased artifact diversity above the floors results mainly from the addition of choppers and milling stones to the artifact inventory. These artifacts signal increasing reliance on plants. The choppers probably were used for forest clearing, and the appearance of milling stones in the deposits signals the advent of technology likely designed for the processing of hard grains. This probability is strengthened by the concurrent appearance of phytoliths from domesticated maize (Zea mays; Voorhies 2004:6.4) associated with the uppermost prepared floor. Maize phytoliths also are present in deposits above the floor, but not below it. In short, it is now possible to detect changes in the archaeological record toward the end of the occupation during the Late Archaic Period that show slow, incremental,

a gender- based model for changes

29

and subtle shifts away from the general Archaic pattern of logistical foraging. In this chapter we explore one of the puzzling questions that has emerged from our prior work: why was the Archaic pattern of clam harvesting and shell mound formation selectively retained during the wet season in the terminal Late Archaic before it was abandoned entirely? Or, reversing the question, why was the long stable economic pattern of the Archaic forsaken first during the dry season, the most productive part of the year in the littoral zone? Although, as we mentioned earlier, we have long thought that the increasing demands of cultivation were in some way causally responsible for this pattern, the ecological details of this transition remained unclear to us. One reason that this issue has been so difficult for us to address is that the ecology of modern day residents on the coastal plain inland of the Acapetahua wetlands is very different from that of the ancient Chantuto people, despite the fact that the basic habitat is largely the same. More precisely, many current inhabitants of the region are cattle ranchers, or they work in industrialscale farming activities (for example, banana and oil palm production), or they are urban dwellers. In addition, there are some people who are small-scale farmer-ranchers but the penetration of modernity is so great that they exhibit minimal, if any, dependency on wild resources. This means that archaeologists can learn little from the modern residents of the inner coastal plain that will help in modeling the lifeways of the ancient Chantuto people5 when they were residing away from the wetlands. In fact, the coastal plain is now completely denuded of the forest cover that our palaeobotanical data show was present at the time the Chantuto people were living there. Accordingly, we have had to look elsewhere to gain a deeper understanding of these processes of ecological change. We have turned to the Lacandón Maya, a fairly well-studied native people whose environment and ecological adaptation are strongly similar to those of the ancient Chantuto people.6

30

archaic to formative

THE LACANDÓN MAYA The Lacandón live in the northeastern part of Chiapas, adjacent to the border with Guatemala (Baer and Merrifield 1971:ix). The region has a tropical climate in which the average monthly temperature never falls below 18° C (classed as “A” in the Koeppen system of climate classification; Vivó Escoto 1964:205), as is also the case for the coast of Chiapas. There is some rain throughout the year, but in January through March— during the coolest time of the year— rain is very scarce (Amw’ in the Koeppen system). These three months constitute the dry season. The heaviest rains occur from June through August. Speaking of the Maya forest generally, including the Lacandón area, Nations (2006:50) states that the rainy season occurs between late May and late November, “a six month period when up to 3 meters (118 inches) of rain fall in some parts of the forest.”7 On the coast of Chiapas there is also a definite and long dry season in the winter (low sun period) typical of tropical wet-and- dry climate (Vivó Escoto 1964:212) (Aw in the Koeppen system), followed by a season of heavy rain. In summary, the climate of the two regions is broadly similar but rainfall is more sharply seasonal on the Pacific coast than in the Lacandón region. The vegetation of the Lacandón habitat consists of tropical evergreen forest grading from north to south to tropical deciduous forest. Because the Lacandons practice a long-fallow system of agriculture, their habitat also is characterized by a patchwork of fields (milpas) and  old fields (acahuales) in various stages of vegetational succession. We have reason to believe that the Chiapas coastal plain also supported the two types of climax tropical forest during the Archaic Period, but the spatial patterns, of course, would not be exactly the same in these two regions. Like the Lacandons, the pioneer cultivators on the Chiapas coastal plain would have created fields as agriculture became an increasingly important economic pursuit. Another strong parallel between the two habitats under consideration is the spatial pat-

terning of aquatic resources. In the Lacandón forest, large freshwater lakes provide critical food resources, especially fish8 and other aquatic organisms (birds, turtles, crustaceans, and so forth). The analogous resource patches on the coast of Chiapas are the tidal lagoons that occur sporadically along the coastline. Despite the obvious fact that the Lacandón habitat is not a coastal one, as is that of the Chantuto people, our research has shown that the ancient Chantuto people did not exploit the open marine environment or the sandy beach but rather focused their attention on the resources available in the coastal lagoons (Voorhies 1976, 2004) and the wetlands generally. Accordingly, both the Chantuto people and the Lacandons have exploited aquatic habitats that are comparable in their spatial characteristics: they are large circumscribed aquatic resource patches (lakes vs. lagoons) hosting a similar array of potential food resources. Moreover, the Lacandón forest is well drained by many permanent rivers, a situation also characteristic of coastal Chiapas. In the Lacandón forest the drainages form a dendritic pattern (Nations and Nigh 1980:figure 1), whereas on the Chiapas coast drainages are linear and perpendicular to the coastline. The Lacandón procure fish and other resources from the rivers and the Chantuto people probably did the same, but the lack of bone preservation at archaeological sites on the inner coastal plain prevents us from confirming this. In addition to broad similarities between these two habitats, we think that the ecological niches (that is, ecological roles) of the two groups of people being compared are quite similar in that both have practiced a mixed economy involving cultivation and foraging. That is, their place in their respective ecosystems is similar. Another way of expressing this is that the ecological adaptation of the Lacandons involved significant dependency on wild resources despite the fact that they were principally farmers. Moreover, their cultivation practices were those of low technology and involved field rotation as is generally modeled for the earliest cultiva-

tors of the Chiapas coastal plain and elsewhere in Mesoamerica (for example, Kennett, Voorhies, and Martorano 2006). So, if we want to learn more about agroforestry 9 in Mesoamerica (Nations 2006), this is where we are likely to fi nd the best native instructors. This is not to deny the fact that the Lacandons have a long history of influences from outsiders, principally demographic alterations from introduced plagues beginning in the fi fteenth century and the arrival of metal tools several centuries later. Despite this, when the Lacandons began to be studied they remained “as isolated from western culture and as conservative in their retention of indigenous lifeways as any people native to Middle America” (Baer and Merrifield 1971:ix) and are considered to have been largely autosufficient until the second half of the twentieth century (Marion Singer 1991:95). Nations (2006:134), speaking of the Lacandón Maya, forcefully makes this same point: “Until the middle 1960s, the immensity of the lowland tropical forest protected the Lacandones from many of the changes that rolled over other indigenous groups in Mexico. This isolation allowed the group to preserve environmental adaptations that are intimately tied to the forest ecosystem.” In summary, the Lacandons certainly are not perfect analogs to the Chantuto people, but we are convinced that they provide the best available opportunity to learn more about the ecology of tropical forest groups practicing a mixed economy involving cultivation and foraging in Mesoamerican tropical forests.

Sources on the Seasonality and Mobility of the Lacandons The most fine-grained analysis of seasonality and mobility of the Lacandons is that of MarieOdile Marion Singer (1991), whose work has been our principal source of information about their ecology. This author unfortunately does not provide much information about her actual fieldwork other than stating that most of her observations pertain to the group living at

a gender- based model for changes

31

Lacanjá (Marion Singer 1991:31) and that she began her investigations in 1986 (Marion Singer 1991:11). This seminal work is not an ecological study per se, resulting in frustrating omissions in the data. We consulted several other publications in order to supplement Marion Singer’s information. These include an early publication by Alfred Tozzer (1907), who compared the lifeways of the Lacandón and Yucatec Maya. We also consulted the publication of Phillip Baer and Merrifield (1971). Baer, an evangelical missionary, lived with his family for fifteen years in the northern part of the Lacandón area beginning in 1944. In 1955 Baer began to have greater contact with the southern Lacandón. These experiences were later supplemented in the 1960s by work with informants (Baer and Merrifield 1971:xi). Baer’s coauthor, Merrifield, assisted Baer in his work with informants. We also rely on James Nations’ (2006) recent publication, which is extremely useful for fine-grained observations of forest ecology, and on Nations and Nigh (1980) for details concerning long-fallow agriculture of the Lacandón Maya.

Lacandón Annual Cycle of Procurement and Mobility According to Marion Singer’s analysis, the Lacandón exhibit three basic patterns of economic activities and mobility throughout the year. She refers to these as cycles of technoeconomic activities (ciclos de actividades tecnoeconómicas). These patterns, which we shall describe below, coincide with three periods within the annual climatic cycle: the dry season, the wet season, and the intermediate season. The spatial arrangement of the traditional Lacandons may be visualized as consisting of widely dispersed settlements each consisting of one or two related families10 who lived close to rivers (the Southern Lacandón) or the edges of  lakes (the Northern Lacandón) (Nations, personal communication; Nations and Nigh 1980:14). The houses were surrounded by fields (milpas). However, since about 1970 most fami-

32

archaic to formative

lies have moved into settlements and their fields are more distant from the community. Interspersed spatially with the milpas are old fields in various stages of fallow. Nations and Nigh (1980) have detailed the importance of these fallow fields as locations for resources for the Lacandón population. Finally, there is the deep forest that lies even farther from the residential base than the active and fallow fields. So the mobility pattern of the Lacandons radiates successively from their residential bases to their active and old fields, and then farther into the forest. In the following paragraphs we describe the spatial and temporal patterns of these people. Dry Season This period begins around the New Year when insolation is low, the temperature is relatively cool, and there is very little rain. The period generally extends from January through March (Marion Singer 1991:171) or April (Baer and Merrifield 1971:166). During this short dry season Lacandón men invest a great deal of labor preparing their fields for the main crop of the year (Nations and Nigh 1980:9).11 Preparation involves, in sequence, cutting underbrush, felling trees, and burning vegetation prior to planting. Planting takes place as soon as the rains are imminent. The demands of agriculture require that at this time of the year the men stay close to their fields, which means that extended hunting expeditions into the deep forest and lake fishing expeditions requiring relatively long forays from the residential base are infrequent. When men do undertake hunting expeditions in the forest, they hunt wild pigs (Tayassu peccary, T. tajacu), deer (Odocileus virginianus, Mazama americana), hocofaisán (Crax rubra), agoutis (Agouti paca), and other mammals and birds. At this time of the year some animals congregate near water sources; this facilitates hunting success. Game animals are also obtained in fields and fallow fields (acahuales), where they can be captured readily (Marion Singer 1991:138). Hunting in the fallow fields is often done using traps for smaller animals and ambush for the larger animals (Marion Singer

1991:101). Nations and Nigh (1980:17) observe that a large portion of the dietary protein of the Lacandons is provided by animals that feed in the acahual. In addition, the low water level in the rivers near the residential bases favors all three types of fishing practiced by the Lacandons. Adult men catch fish from canoes using hook and line, projectiles (spears, harpoons, and arrows), and nets. Boys employ fish poison in river bends, a technique that is practiced only during the dry season when water levels are low. Adult women procure small fish using fish traps along riverbanks. These various fishing activities satisfy most of the group’s need for protein without requiring that people forage far from the home base during the dry season. The dry season, moreover, is not a good time for collecting wild plant foods in the forest since it is the season when most plants are flowering rather than fruiting. This means that women do not forage widely from their home base during this season. In addition, in normal years the granaries are full of recently harvested maize so the incentive for women to collect wild plant foods is diminished.12 In summary, during the dry season neither men nor women tend to range very far from the residential base. Instead, their ranges extend to nearby rivers and to fields and fallow fields, located relatively close to the residences. However, men range farther than women during this season, mostly on occasional fishing expeditions to the lakes (Marion Singer 1991:174), which seem to take place toward the end of this season, and occasional hunting expeditions to the forest when agricultural tasks permit. During the dry season the bulk of animal protein consumed by the group is acquired by men, who hunt in their milpas and old fields as well as take fish from nearby rivers. Rainy Season At the onset of the rains Lacandón men continue to be intensively occupied with farming activities. First they plant the principal maize crop13 along with a wide variety other inter-

planted root, tree, grain, and vegetable crops (see Nations and Nigh 1980:table 1), after which they must protect and nurture the young plants.14 Planting of different crops can occur throughout April and May at specific times that are signaled by the flowering of par ticular forest plants (Baer and Merrifield 1971:174; Nations and Nigh 1980:9– 11). Once sown, the fields must be weeded periodically15 and the plants must be continually protected. Many animals, such as pacas, peccaries, deer, and other edible mammals are attracted to the young plants in the milpa, where they often fall prey to Lacandón hunters. At the onset of the rainy season the farmers harvest the winter maize crop as well. Because of these farming demands during the rainy season, men tend to remain spatially tethered to their fields. This limitation in men’s ranges becomes most pronounced when rains become heaviest, between April and August. In April, however, before the rains become too heavy, expeditions to the lakes are sometimes organized. These fishing parties usually consist of related men and boys and last for several days (Marion Singer 1991:122). Men fish from canoes using nets, spears, harpoons, bow and arrow, and hook and line.16 The Lacandons consider these fishing techniques as exclusively masculine (Marion Singer 1991:129). During these trips men also hunt large animals, such as crocodiles (Marion Singer 1991:155) and turtles by combing through the reeds with their canoes. Tozzer (1907:54) mentions that turtles and turtle eggs form an important part of the diet and that as many as 400 to 500 eggs were sometimes procured in a single hunt. Crocodile eggs are also collected at this time of year (Marion Singer 1991:155). Men also hunt faisánes and howler monkeys at the edges of the lakes (Marion Singer 1991:121). These hunting and fishing expeditions must be mounted before the forest is flooded and the rivers overflow their banks. Fishing trips often last two days (Baer and Merrifield 1971:233, 245). Travel time may be as long as four or five hours to distant lakes (Marion Singer 1991:121). The fishing party usually spends the night on an island in

a gender- based model for changes

33

the middle of the lake in order to be safe from predators. In general, however, forest hunting opportunities for the men diminish substantially as the rainy season progresses. At this time the game animals that can be procured in fields and fallow fields during the dry season disperse into the forest. One reason for this is that the winter maize crop has been removed from the fields, lessening the opportunity for animals to forage. At the same time many animals are attracted deeper into the forest by the seasonal availability of fruits. This is particularly the case of chicozapote (Manilkara zapota), zapote mamey (Pouteria sapota), ramón (Brosimum alicastrum), and ramón colorado (Trophis racemosa), whose fruit attracts peccaries, pavo de monte (Penelope purpurascens), monkeys (Ateles geoffroyi, Alouatta pigra), and pacas (Agouti spp.). Also, many prey animals tend to congregate less during this season than in the dry season, making them more difficult to locate and necessitating a greater expenditure of energy to procure them. Finally, and perhaps most significantly, forest flooding hinders hunting activities by making it difficult for hunters to track their prey (Marion Singer 1991:138). The maturation of fruits in the ecotones at the edge of the forest does allow for some hunting activities in ranges close to the home base. This is especially true of several prey species: faisánes, pavo de monte, and tepescuintles (Cuniculus paca). Also, in recent times the Lacandons rely heavily on their domestic poultry during the relatively lean time of the rainy season. As mentioned, the rainy season is a time of fruition in the forest when there is an abundance of edible fruits and plants. Zapote mamey and other zapotaceas fruit at this time are consumed in abundance by the Lacandons, as well as by many of their prey species. Women forage intensively during this season, usually in small groups accompanied by their older children (Marion Singer 1991:172– 173). In addition to plants, women and children collect crustaceans, insects, and all kinds of eggs in order to complement the largely plant-based diet of the long rainy season. Animal protein drops off during this season and

34

archaic to formative

is substituted somewhat by proteinaceous plant foods such as the seeds of zapote mamey and ramón. Also, the heart of the corozo palm, now a commodity in short supply, is sometimes procured when meat is scarce because of the rains (Marion Singer 1991:159). Nevertheless, in the tropical forest vegetal protein sources tend to be of low quality (Nations and Nigh 1980:17), putting a premium on animal protein. Another source of protein that takes on greater importance at this time of year is trapped fish, a resource principally procured by women (Marion Singer 1991:127, 129) and sometimes by boys (Nations, personal communication). Women trap fish throughout the year, and often daily (Marion Singer 1991:127,129) in conjunction with other activities such as bathing, fetching water, and washing clothes. The relative importance of the diet of trapped fish increases when other sources of protein diminish during the period of heaviest rains. The traps, woven from a flexible vine by women (Figure 2.1; Marion Singer 1991:127) or rarely men (Nations 2006:80) are placed in river branches near Lacandón residences, near the family’s milpa (Marion Singer 1991:129), or at lake shores (Nations, personal communication). The traps are different sizes (Marion Singer 1991:127) but are probably most effective in capturing smaller fish. Although Marion Singer (1991:129) asserts that only women use fish traps, Baer and Merrifield (1971:245) describe a fish trap made from a bottle gourd that was used by a man.17 This observation serves to remind us that in small hunter-gatherer and horticultural societies economic specializations by gender are not rigid but necessarily must be flexible to ensure group survival and reproduction. Nevertheless, the sources are clear that women usually trap small fish, whereas large fish are hunted with projectiles, hook and line, or are netted by men. The freshwater snail (Pachychilus indiorum) is another resource that becomes more important dietarily when meat and fish are scarce (Marion Singer 1991:168), especially during the wet season. These snails are collected by women and children from streambeds and

FIGURE 2.1 A Lacandón woman from Naha weaving a basketry fish trap. Photograph by Gertrude Duby. Reprinted by permission of the Asociación Cultural Na Bolom, A.C.

appear to have been especially important for food and as a source for lime18 in the recent historical past (Nations 2006:82 ff ). This is also the period when craft activities are practiced, especially during the period of heaviest rains and strong winds, which keep people housebound. In summary, during the rainy season the mobility of Lacandón men is generally restricted to agricultural plots, whereas women tend to range farther from the home base, even if only for brief excursions to zones that do not require more than an hour’s walk from their residences (Marion Singer 1991:174). In addition, because game animals are difficult to procure as this season progresses there is a tendency for the Lacandons to depend more on animals that occupy lower trophic levels (for example, mollusks, crustaceans, small primary consumer fish) than they do at other times in the annual cycle. These are the animals that are preferentially procured by Lacandón women. Transitional Season In the Lacandón environment the months September through December are considered by Marion Singer as a transitional season between the heavy rains and the dry season beginning

in January. During this period, rains are intermittent and it is possible to move more freely through the high forest (Marion Singer 1991:174) than it is during the period of heaviest rains. However, this is not a time when forest fruits are maturing, and so the opportunities for collecting wild plant foods are reduced. In contrast, this is the optimal time for hunting in the forest. This is because many game animals have fattened and reproduced during the preceding rainy season when hunting pressure on the prey species is lessened. Simultaneously, it is not necessary for the men to work every day in their fields. This allows them to go on longer hunting and fishing expeditions. Also, nowadays, this is the time in the annual cycle when people take trips to distant towns such as San Cristóbal de las Casas, Palenque, and Villahermosa (Marion Singer 1991:175) in order to sell craft items and purchase supplies. These excursions are alternated with work in the fields such as the doubling of maize stalks (an infield storage technique), construction of infield granaries, and the harvest of the principal maize crop. Winter maize is planted at the end of this period. Marion Singer (1991:175) stresses that for men this is the season of maximum mobility.

a gender- based model for changes

35

Men range between their milpas and the high forest during this time of year. According to Marion Singer, the long-distance fishing and hunting expeditions into the deep forest are scheduled during slack times in the agricultural cycle (Marion Singer 1991:129). Women, in contrast, are more tethered to their home bases than at other seasons of the year and as compared with the men during the intermediate season. In summary, the data for the Lacandón indicate that during the dry season neither men nor women range far from their home base. In par ticular, men are tethered to fields because of the heavy workload associated with field preparation, and at the same time they are able to hunt productively in the milpas and acahuales. Women, in contrast, are not drawn far from the home base because opportunities for collecting are not particularly favorable in the forest. During the wet season men also tend not to range far from the residential base, although there is a window of opportunity for longer expeditions just before the onset of the heavy summer rains. During the heavy rainy season, hunting and fishing opportunities are poor and men are occupied with craft production and agriculture. This is the time of year when there is a sharp decline in the contribution of meat from game animals (principally hunted by men) to the Lacandón diet. Women, in contrast, tend to range relatively far from their home base at this time of year, when they penetrate deeply into the forest on collecting expeditions for plant foods as well as for various protein-rich resources. Finally, during the intermediate season men tend to range far from the home base to take advantage of good hunting and fishing conditions. Women, in contrast, remain relatively close to home. As we have discussed, during the terminal Late Archaic Period the Chantuto people had incorporated domesticated plants into their diet and they were engaged in cultivation.19 These observations are well documented by phytolith, pollen, and charcoal frequencies in sediment records reported elsewhere (Jones and Voorhies

36

archaic to formative

2004). The artifact record also supports this interpretation. Also, during the terminal Late Archaic Period the Chantuto people continued to visit the wetlands but did so almost exclusively during the wet season. In other words, by late in the Late Archaic our data indicate that dry season forays into the wetlands diminished considerably. Using the behavioral ecology of the Lacandons as a guide, we can predict that during the wet season women would have been more likely than men to journey to the wetlands, since this is the season of maximum mobility for Lacandón women. How can we test this hypothesis using the archaeological data available to us in coastal Chiapas? One promising line of evidence comes from the examination of fish bones from shell mounds.

ANNUAL CYCLE OF THE CHANTUTO PEOPLE In a previous study of the archaeoithyofauna from Zapotillo, one of five shell mound sites studied by us and dating to the Late Archaic Period, Cooke et al. (2004:298, figure 5.7a) found that although bones of big and small fish were found throughout the stratigraphic section, the bones of small, low trophic level fish and the bones of large, high trophic level fish were strikingly out of phase with each other (Figure 2.2). That is, in the microstrata where large fish bones were prevalent, small fish bones were scarce and vice versa. This led the authors to suggest that these two groups of fish were not captured synchronously, but the reasons for this were not clear. Cooke et al. suggested that a gendered division of labor was one of several possible explanations for this evidence. We pursue this idea further here. In Cooke et al.’s analysis of fish bones from Zapotillo,20 sleepers (Eleotrididae) constitute the most abundant group of fish represented, followed by marine catfish (Ariidae). Both sleepers and catfish are prevalent in the Acapetahua estuary system today, and their presence in the archaeological record is a reliable bioindicator

FIGURE 2.2 Frequency (in percent of total diagnostic bones) of two groups of fish at Zapotillo. The frequency of the most common fish, Pacific fat sleeper, is shown by a solid line, whereas the frequency of corvina, snapper, and snook combined is shown by the dashed line. The vertical line indicates the contact between Archaic deposits to the right of the line and later ceramic-bearing deposits on the left. Source: Voorhies 2004:fi gure 5.7a.

that the ancient estuary was basically similar to the modern one. It is also highly signifi cant that the majority of individuals were small (that is, juveniles). In fact, although sleepers can weigh as much as 2 kilograms (kg) (Cooke et al. 2004:269), the majority of sleepers from one of these shell mounds reported by Cooke and colleagues weighed only 60 grams (g)21 (Cooke et al. 2004:292). Sleepers are low trophic level taxa within the aquatic food chain. They are considered to be either primary or secondary consumers (Cooke et al. 2004:269), and their place in the food chain from primary to secondary likely shifts through the individual life cycle. They are bottom feeders and abundant in brackish water (Allen and Robertson 1994:256). Dormitator latifrons, the Pacific fat sleeper, which is particularly abundant in the assemblage, has a maximum depth range of only 2 m. Studies of

modern fishing technologies show that various kinds of aquatic barriers such as weirs, coupled with nets, are effective means of capturing both sleepers and catfish (Cooke et al. 2004:292). It is possible also to capture these fish in small shallow pools left by receding waters. Thus the technological requirements for capturing these fish are relatively low. Cook et al. found that at Zapotillo the bones of the fat sleeper were out of phase with snook (Centropomus), snapper (Lutjanus), and corvina (Cynoscion), all high trophic level predatory fish. Bones of these three fish are among the most abundant fish bones at Zapotillo (Cooke et al. 2004:258–259), after sleepers and catfish. They are tertiary consumers and travel between the marine and brackish water environments throughout their life cycles. Although the archaeofauna discussed here suggest that many of the captured fish were not full grown,

a gender- based model for changes

37

they are significantly larger than the sleepers from the same archaeological sites (Cooke et al. 2004:figure 5.5). For example, Cooke et al. (2004:264) mention that snook weighing as much as 8 kg were captured by the people who formed the Zapotillo site. One possible method of procuring these fish, according to Cooke et al. (2004:294), would be by spearing or harpooning them in the upper lagoon or mangrove channels. If this is the method used by the Chantuto people, it is possible that boats were employed. These fish can also be taken with hook and line, but the apparent22 absence of hooks in the Late Archaic archaeological record, plus consideration of the pattern of recovered species, prompted Cooke et al. (2004) to infer that angling was unlikely. Finally, fishing technology involving weirs coupled with devices for removing the fish (such as nets, harpoons, and spears) could be used for these top predatory fish. The foregoing considerations suggest the possibility that different kinds of fish were procured with distinct technologies and by people of different ages and genders within the Chantuto people’s society. Women, probably accompanied by children, might have used traps or hand nets for capturing the small sleepers (primary consumers) hiding near the lagoon shores, whereas men might have used boats, spears, harpoons, arrows, or a combination of any of these to procure the larger predatory species that inhabit deeper water. If such a gendered division of fishing was practiced by the Chantuto people, then we should predict that by the terminal Late Archaic, when the commitment to cultivation was intensifying and people were visiting the wetlands mostly during the wet season, the archaeoithyofauna should be dominated by  the bones of small, low trophic level fish rather than those of large, high trophic level fish (given the assumption that men were engaged during this season in cultivation activities, as with the Lacandón). This prediction is based on the likelihood that women would range farther from the home base than men during the wet season and that the dietary need for proteinaceous resources acquired by women would

38

archaic to formative

be the most critical at this time in the annual cycle, when access to large game is diminished. To perform an initial test of this proposition, we turn to the archaeofaunal data from the El Chorro shell mound, which dates to the terminal Late Archaic Period, a time when Chantuto people were becoming more committed to maize cultivation. For comparative purposes we will also consider the archaeofaunal data from Cerro de las Conchas shell mound, which dates to the Middle Archaic Period.

El Chorro: A Terminal Late Archaic Site The El Chorro shell mound shares many characteristics with the other five shell mounds that we have investigated (Voorhies 2004). Its contents consist of bedded shell deposits dating to the Archaic Period overlain by a soil mantle that is younger in age. We know from a sediment core below the water table that the site formed initially on a substrate of mangrove mud and that it developed as an artificial island. El Chorro is the smallest of the group of shell mounds that we have investigated, being only 0.20 hectares (ha) in area. It also is more apical in shape and less eroded, which initially suggested to us that it might be younger in age than the other shell mounds of the area. We worked at El Chorro in 2005 and at that time secured five radiocarbon dates, three from the excavation and two from a core taken below the excavation level (Kennett et al. 2007:11, Cuadro 3). The dates when arranged in their stratigraphic order are not statistically different, leading us to infer that the site accumulated rapidly between approximately 2700 and 2200 b.c. In other words, as we suspected, this site formed during the terminal Late Archaic Period, at a time when the commitment to agriculture was well under way in the region. The study of seasonality by means of oxygen isotopes in the margins of clamshells sampled from excavation levels shows that at this site clams were preferentially collected during the wet season (Figure 2.3), except for the uppermost level (0.86 m), when clam procurement

Season Wet

Dry

Seasonal Profiles

Nivel (m)

–4 –6 –8 δ18O

0.86

–10

20

16

12

8

4

–12 –4 –6 –8 δ18O

2.00

–10

20

16

12

8

4

–12 –4 –6 –8 δ18O

3.00

–10

20

16

12

8

4

–12 –4 –6 –8 δ18O

4.00

–10

20

16

12

8

4

–12 –4 –6 18 –8 δ O

5.00

–10

20

–11

–10

–9

–8

–7

–6

–5

–4

16

12

8

4

–12

Distance (mm)

δ O Isotope (PDB) ( = Margin of one clam) 18

FIGURE 2.3 Ratio of oxygen isotopes from the margins of marsh clamshells taken from five excavation levels in Unit 1, El Chorro

once again was carried out all year round. The data also reveal that some limited clam procurement took place during the intermediate and dry seasons, but clearly the wet season was the preferred season to procure clams from the coastal lagoons. Accordingly, the seasonality study from El Chorro supports the results from coeval deposits at the Tlacuachero shell mound: during the terminal Late Archaic, clam collection in the Acapetahua wetlands became much more restricted to the wetseason, whereas a more year-round procurement pattern manifested earlier in the Archaic Period. Using the ecology of the Lacandons as a guide, we predict that the El Chorro site, which formed late in the Late Archaic Period and was used primarily in the wet season, should have been visited more frequently by Chantuto women than by Chantuto men. This prediction is based on the fact that among the Lacandón during the wet season it is the women who are most mobile and their protein contribution to the diet is most significant. If this pattern of mobility and gendered

division of labor is applicable to the ancient Chantuto people, then we expect that the faunal remains from the terminal Late Archaic deposits at El Chorro should have many more bones from small, primary consumer fish suspected as being procured through the use of basket traps, nets, or both by women, and comparatively fewer bones of large top predatory fish suspected as being procured through the use of other fishing technologies by men. Thomas A. Wake identified 5,384 bones from the excavated Archaic Period deposits at El Chorro (Kennett et al. 2007:40). Of these, 97.1 percent are from fish, 2.3 percent from turtles, 0.3 percent from birds, and 0.3 percent from mammals. This makes it clear that during the terminal Late Archaic Period the principal function of the El Chorro site was as a fishing station, at the same time that it was used for the processing of marsh clams. Significantly, the procurement of vertebrate animals other than fish was evidently a relatively  rare occurrence, because their bones

a gender- based model for changes

39

are conspicuous in their near total absence, which is not the case at all other shell mounds in the region. It is also significant that 90.1 percent of the identified fish are those that prefer waters with low salinity. These are the sleepers (eleotrids), gar, cichlids, and poeciliids. These freshwater-loving fish are found in relatively high abundance in the uppermost reaches of a lagoonal- estuarine system during the dry season and somewhat farther downstream during the wet season. Accordingly, the archaeoithyofauna support the results from the oxygen isotope study of clamshells in suggesting that fishing, like clamming, may have been carried out preferentially during the wet season, when water salinity is low in the coastal lagoons. A second important observation about the fish fauna at El Chorro is that the recovered bones are overwhelmingly from small fish rather than large fish. This was abundantly clear to us when we were excavating the site. We recovered a huge number of tiny fish vertebrae during screening, whereas large fish vertebrae were virtually absent. To quantify this field observation, we compared the sizes of fish vertebrae from the terminal Late Archaic Period El Chorro site with the size of fish vertebrae from the Middle Archaic Period (c. 5500–3500 b.c.) site of Cerro de las Conchas. Cerro de las Conchas was formed throughout the annual cycle, according to our seasonality data (Voorhies et al. 2002), so we suspect that both men and women were using that site for resource procurement and processing. Accordingly, the faunal assemblage at Cerro de las Conchas should exhibit evidence of large fish as well as evidence of small fish.

Cerro de las Conchas: A Middle Archaic Site The Cerro de las Conchas site and the research conducted at it is presented in Voorhies (2004:81ff ) and Voorhies et al. (2002). The lowest of three strata (Stratum III) consists of the familiar bedded marsh clamshells so typical of the Archaic Period shell mounds on the coast of Chiapas. Another Archaic Period stratum

40

archaic to formative

(Stratum II) overlies the clamshell stratum on one flank of the site. It consists of a matrix of limpets and mussels. This stratum is unique for two reasons: the contents of the shell matrix and the presence of what appear to be fishhooks and a fish gorge. The uppermost stratum (Stratum I) is a dark soil that was deposited after the Archaic Period and is not pertinent to the present discussion. Thomas A. Wake examined 770 bones from Stratum III, the lower of the two Middle Archaic strata. Fish bones dominate (number of identified specimens, or NISP = 94.29 percent), followed by reptiles (NISP = 4.68 percent), and mammals (NISP = 1.03 percent). Amphibians and birds were absent. Wake also examined 5,225 bones from Stratum II at Cerro de las Conchas, the upper stratum also dating to the Middle Archaic Period. Once again fish dominate (NISP = 98.57 percent), followed by reptiles (NISP = 1.28 percent), and then mammals (NISP = 0.13 percent). Birds represent a negligible proportion of the assemblage, and amphibians are absent (Wake et al. 2004:172). The above data make clear that, like El Chorro, Cerro de las Conchas is a site where the resource focus heavily emphasized fish, as well as mollusks, the shells of which form the bulk of the archaeological deposits.

Comparison of Fish Captured at Two Sites According to our expectations based on the Lacandón model, there should be more top predatory fish in the earlier, Middle Archaic site than in the later, Late Archaic site, and there should be larger fish in the assemblage from the early site than in the later one. It is important to note that these data were collected from deposits that were screened in the field at both sites using one- quarter-inch screens. It is well documented that the finer the screen the better the recovery of small bones, which means, of course, the more accurate the assessment of the role of small fish in the paleodiet. Quarter-inch screens are not very fine, which means that small fish probably are underrepresented in our

TABLE 2.1 A Comparison of Selected Fish Taxa From Two Shell Mounds on the Coast of Chiapas, Mexico VALUES ARE PERCENTAGES OF TOTAL BONES (NISP) FROM EACH ANALYTIC UNIT

Fish

El Chorro

Cerro de las Conchas

Cerro de las Conchas

Stratum C (Terminal Late Archaic)

Stratum II (Middle Archaic)

Stratum III (Middle Archaic)

High Trophic Level Predators Snook (Centropomus)

1.45

4.34

7.27

Snapper (Lutjanus)

0.59

20.15

17.00

Corvina (Cynoscion)

0.29

0.51

3.51

Total

2.33

25.00

27.78

Low Trophic Level Herbivores Sleepers (Eleotridae)

55.84

samples. However, we are most interested here in determining the relative importance of large fish at these two sites, which we can do reliably with the available data. To compare taxa of two different trophic levels for the two sites, we used NISP of snook, snapper, covina, and sleepers as percentages of total NISP for each analytic unit. Snook, snapper, and corvina are tertiary consumers that we suspect may have been procured by Chantuto men. Sleepers, in contrast, occupy a lower level on the food chain, and we suspect that women may have procured them. The results of this comparison are shown in Table 2.1. These data make it clear that the three top predatory fish make up a greater percentage of total bone specimens in the two Middle Archaic Period strata at Cerro de las Conchas than in the terminal Late Archaic Period deposit at El Chorro. Specifically, at Cerro de las Conchas bones of snook, snapper, and corvina constitute approximately one-quarter or more of all bones in the Archaic deposits, whereas at El Chorro the bones from these three taxa are about only 2 percent of total bone specimens. In contrast, bones of sleepers are less than 1 per-

0.17

0.13

cent of total bones at Cerro de las Conchas but approximately 56 percent of total bones at El Chorro. Accordingly, the comparison of fish from different trophic levels meets the expectations regarding a hypothesized change in fishing practices over time, with the Middle Archaic deposits yielding significantly more top predatory fish than the terminal Late Archaic deposits. Conversely, as predicted, the terminal Late Archaic deposits have significantly more of the low trophic level fish than the earlier Middle Archaic deposits. We have also predicted that the sizes of fish would be smaller at the late site than at the early site. To test this expectation we compare the widths of fish vertebrae from Archaic Period deposits at the two sites under consideration. Previously Voorhies has shown (2004:figure 5.2) that there is a direct linear relationship between the diameters of vertebrae and the overall length of fish. This means that it is possible to estimate the length of a fish from a determination of the width of its vertebrae. Erik Marsh measured the widths of fish vertebrae from the two faunal assemblages under consideration. Marsh used digital calipers that

a gender- based model for changes

41

800

Cerro de La Conchas El Chorro

700 600

Count

500 400 300 200 100 0 1

2

3

4

5

6

7

8 9 10 11 Diameter (mm)

12

13

14

15

16

17

18

FIGURE 2.4 Graph of diameters of fish vertebrae from Cerro de las Conchas, a Middle Archaic Period shell mound (N = 1379), and El Chorro, a terminal Late Archaic Period shell mound (N = 2598)

allowed the data to be automatically recorded into an Excel computer database. Then, the frequency of diameters of vertebrae sorted by size increments was plotted on a graph (Figure 2.4). As may be seen in Figure 2.4, there is a striking difference in the size ranges of fish vertebrae, used as a proxy for overall fish length, for the two sites. Marsh measured the diameters of 1,379 vertebrae from the Middle Archaic deposits at Cerro de las Conchas. These ranged in diameter from 2.61 mm to 24.18 mm, with a mean of 6.40 mm. These data show that the people who formed the Cerro de las Conchas site were catching a wide range of big and little fish. The mean whole fish length as estimated from the mean diameter is approximately 35 cm, using data reported by Voorhies (2004:figure 5.2) on the relationship of diameter of vertebrae to whole length of fish. Compare these data with those from El Chorro, where Marsh measured 2,598 vertebrae from Late Archaic Period deposits. The range of vertebrae from this site is 0.97 mm to 11.82 mm, with a mean of 2.74 mm (Figure 2.4). The mean whole fish length as estimated from the mean diameter using Voorhies’ table is approximately 18 cm, almost half

42

archaic to formative

the estimated length of mean fish size from Cerro de las Conchas. In summary, our evidence shows clearly that the large top predators that were fished during the Middle Archaic when the Cerro de las Conchas site was formed are conspicuously absent in the terminal Late Archaic period El Chorro site, where smaller, lower trophic level fish were procured almost exclusively.

CONCLUSIONS We propose that during the terminal Late Archaic Period (2700–2000 b.c.) the economic activities of Chantuto men and women may have been changing at different rates, with women retaining their traditional Archaic pattern of exploiting coastal wetlands longer than men. The basis for this argument is not linked to suppositions about the inherent characteristics of people of either gender but rather to the behavioral ecological perspective that women and men in hunter-gatherer societies pursue different subsistence strategies based on different goals relating to reproductive fitness (Zeanah 2003). In such societies, “men usually hunt

larger game whereas women gather plants and smaller animals (Bliege Bird 1999)” (Zeanah 2003:3). Men are driven by the goal of maximizing mating opportunities, best realized by the sharing of game, whereas women are driven by the goal of assuring survival of their offspring, best realized by provisioning their children. In this view, women’s subsistence strategy encourages them to pursue smaller, reliable prey rather than larger, more unreliable prey. Our model is also based on an understanding of the ecological adjustments that might come into play as a group of hunter-gatherers becomes more committed to a farming lifestyle and moves away from the procurement of wild resources within a specific ecological setting. Our seasonality data for shell mound formation in the terminal Late Archaic show that these sites formed principally during the wet season rather than the dry season. Hunting and fishing by Lacandón men decline during the wet season when the men are spatially tethered to their residences and agricultural fields. Accordingly, this is the season when men make the least contribution to the daily diet, and in particular their procurement of protein resources declines. In contrast, Lacandón women range farthest from the home base during the wet season, and their economic contribution to diet is greatest during this season because of their elevated contribution of protein. Lacandón women achieve this in various ways, including the procurement of small fish, which they trap, and the collection of crustaceans and mollusks. A similar scenario may have played out during the terminal Late Archaic Period for the Chantuto people. Analysis of fish remains from one Archaic Period shell mound indicates that different fishing strategies were in effect throughout the Archaic, and we have pursued the idea that these may be gender based. Specifically we propose, based on the analogy with the Lacandons, that men procured large, top predator fish and that women procured small, low trophic level fish (we also assume that field preparation and hunting large animals were men’s work, based on the Lacandón case). If

this association is correct, then women’s fishing strategies outlasted men’s practices during the terminal Late Archaic. We acknowledge freely that these archaeological data can be interpreted in other ways, but we argue this model is the most comprehensive and robust available. For example, it could reasonably be proposed that the observed differences in the prey species at Cerro de las Conchas and El Chorro derive simply from locational differences between upper and lower wetlands. That is, Cerro de las Conchas might have formed closer to the open ocean, where large predatory fish were more prevalent than in the upper estuary, where El Chorro might have formed. Or perhaps it could be shown that the large predatory fish are present in the estuary only during the dry season, so any site that formed only during the wet season (as at El Chorro) would not include the bones of such animals. While these considerations might explain the previously noted differences in fish caught at the two sites, they fail to account for other changes in the archaeological record, such as the evidence for increased commitment to farming and the lack of synchrony in the procurement of small, low trophic level fish and large, high trophic level fish in the shell mounds of the Archaic Period. In our preferred interpretation, both women and men made forays into the coastal wetlands as early as the Middle Archaic but they engaged in different procurement pursuits. Women procured clams and small fish whereas men procured large fish. Over the course of the four millennia of the archaeological record, as the commitment to farming gradually increased, men abandoned their dry season fishing expeditions but women continued to make wet season forays to the coastal lagoons in order to procure protein-rich aquatic resources, which otherwise were missing from the group’s diet.

ACKNOWLEDGMENTS The National Science Foundation (BCS- 0211215) funded our fieldwork at El Chorro with a grant to Kennett, and the H. John Heinz III Fund

a gender- based model for changes

43

Grant Program of Latin American Archaeology funded the work at Cerro de las Conchas with a grant to Voorhies. Voorhies conducted the excavations at both sites and was assisted at El Chorro by Fumie Izuka and at Cerro de las Conchas by Linda A. Brown and Michelle Woodward. Kennett conducted the stable isotope work at the Archaeometry Facility at the University of Oregon. The New World Archaeological Foundation, under the directorship of John E. Clark, provided critical logistical support during the field seasons, and the University of California, Santa Barbara provided support during the analysis phase of this research. We are also indebted to the Consejo de Arqueología, Instituto Nacional de Antropología e Historia, Mexico for granting us permission to carry out this research. Thomas A. Wake analyzed the vertebrate faunal remains from both sites, and Brendan Culleton assisted with the oxygen isotope study at El Chorro. Erik Marsh measured the fish vertebrae, and we thank him for the many tedious hours that he spent doing so. We also thank the late Phillip Walker for the loan of the electronic calipers that Marsh used for this task. We are grateful to Richard Lesure for giving us a forum to present this work and to all the participants in the Cotsen Institute of Archaeology symposium for their healthy skepticism that prompted our strengthening of the text. We also received useful comments on a preliminary draft of this article from Michael Jochim, Richard Lesure, Madonna Moss, and James D. Nations. NOTES 1. The preferred social roles associated with people of a specific physical sex are known as gender roles and are always specific to a par tic u lar society. The features known to have a genetic basis are considered a person’s physical or phenotypic sex (Martin and Voorhies 1975:3). Thus, from an analytic point of view it is useful to distinguish phenotypic sex, with its genetic foundation, from gender, with its social foundation. 2. The upper stratum on this mound is soil. It was deposited after the shell deposit and by people with very different lifeways.

44

archaic to formative

3. We know now that this is one of several floors that were constructed under the mound’s summit. 4. The apparent dimensions of this floor, which lies 4.60 m below the current surface, were determined only at four points along two axes. 5. This situation is sharply different for the wetlands, where modern inhabitants have adaptations much more like those reconstructed for the ancient Chantuto people in that microenvironment. See Voorhies (2004) for a discussion of this topic. 6. It could be proposed that the Huaves of coastal Oaxaca would provide a better ethnographic analogy because of their intense focus on resources from coastal lagoons (Millán and García Souza 2003:17), like that reconstructed for the Chantuto people. Although it is true that the lagoons respectively used by the present- day Huaves and the ancient Chantuto people both have abundant populations of marsh clams, shrimp, and fish (Contreras Espinosa 1993), other aspects of the two environments are strikingly different. For example, there are no extensive wetlands in the Huave habitat, whereas the Chantuto habitat has a mature mangrove forest and large patches of herbaceous swamp. Also, we note that the prolonged dry season, xerophytic vegetation, and windswept landscape of the Isthmus of Tehuantepec are markedly different from the reconstructed tropical evergreen and deciduous forests in which the Chantuto people lived. In fact, the aridity, high salinity, propensity for flooding, and climatic extremes all conspire against the Huave adopting farming as a major subsistence activity (Millán and García Souza 2003:20). 7. Baer and Merrifield (1971:166) say that the rainy season starts in May and continues through August, during which time the rain is carried by south winds. Then heavier rains arrive and rivers overflow their banks until the end of December. 8. One ecological assessment program found fifty-five species of freshwater fish in the northwest corner of the Guatemalan Petén, whereas another project identified sixty-seven species of freshwater fish in the Selva Lacandona of Chiapas (Nations 2006:79). 9. This traditional system of farming produces and manages food crops, trees, and animals in the same plot (Nations 2006:138). 10. These families are usually headed by married sons in the north, where residence patterns are patrilocal, and married daughters in the south, which is organized matrilocally (Nations, personal communication). 11. This is called the summer maize crop.

12. However, this is the ideal time for people to gather raw materials to be used later in the year for manufacturing tools and other craft items. 13. Lacandón men are in charge of maize and they plant many other crops, but women do not plant or harvest maize. Women, not men, tend the plants in the household garden, and women do plant and harvest noncorn crops in the distant milpas. Girls and boys also help with this (Nations, personal communication). 14. A Lacandón plot may have “up to 79 varieties of food and fiber crops on a single hectare (ha) (2.47 acres)” (Nations 2006:138). 15. Between two and four weedings are carried out between June and August (Marion Singer 1991:99). The number of weedings required depends critically on whether the plot is surrounded by primary forest or secondary growth (Nations and Nigh 1980:14). Nowadays, men and postpubescent boys do the weeding in the distant milpas. In traditional families living within the milpa, men pulled weeds by hand so as not to scatter seeds, and women may have been involved in weeding as they went about planting and harvesting noncorn crops (Nations, personal communication). 16. Apparently, net and bow and arrow fishing has declined over the last century and hook and line fishing has increased. Nations (personal communication) observes that Lacandón men, to kill fish, formerly used a long thin wooden arrow with no barb. 17. Susanna Ekholm (personal communication, 2008) notes that today the Lacandón use baited glass bottles to collect small fish. 18. Lime is a critical ingredient in rehydrating dried maize kernels (Nations and Nigh 1980:24). 19. According to Neff et al. (2006) and Kennett et al. (2010), maize cultivation began much earlier, but perhaps at this time there was more intensified use or more people were practicing slash and burn agriculture. 20. The authors also analyzed faunal remains from the Tlacuachero shell mound and found that sleepers dominate the assemblage of the Late Archaic Period. 21. This estimated biomass figure was calculated from deposits at Tlacuachero that were fi nescreened. In that study, the Pacific fat sleeper represents 92 percent of the number of individual specimens, 56 percent minimum number of individuals (MNI), and 41 percent estimated biomass (Cooke et al. 2004:280). 22. During fieldwork in 2009, after Cooke et al.’s publication, Voorhies found fish hooks in early Late Archaic Period deposits at Tlacuachero. These same levels also contained the bones of large fish.

REFERENCES Allen, Gerald R., and D. Ross Robertson 1994 Fishes of the Tropical Eastern Pacific. Crawford House Press, Bathurst, Australia. Baer, Phillip, and William R. Merrifield 1971 Two Studies on the Lacandones of Mexico. Summer Institute of Linguistics of the University of Oklahoma, Norman. Binford, Lewis R. 1980 Willow smoke and dogs’ tails: Huntergatherer settlement systems and archaeological site formation. American Antiquity 45: 4–20. Contreras Espinosa, Francisco 1993 Ecosistemas costeros mexicanos. Universidad Autonóma Metropolitana, Unidad Iztapalapa, México. Cooke, Richard G., Máximo Jiménez, Conrado Tapia, and Barbara Voorhies 2004 A Closer Look at the Late Archaic Fish Fauna. In Coastal Collectors in the Holocene: The Chantuto People of Southwest Mexico, edited by Barbara Voorhies, pp. 207–299. University Press of Florida, Gainesville. Jones, John G., and Barbara Voorhies 2004 Human and Plant Interactions. In Coastal Collectors in the Holocene: The Chantuto People of Southwest Mexico, edited by Barbara Voorhies, pp. 300–343. University Press of Florida, Gainesville. Kennett, Douglas J., Dolores R. Piperno, John G. Jones, Hector Neff, Barbara Voorhies, Megan Walsh, and Brendan J. Culleton 2010 Pre-pottery farmers on the Pacific coast of Southwest Mexico. Journal of Archaeological Science 37: 3401–3411. Kennett, Douglas J., and Barbara Voorhies 1995 Middle Holocene periodicities in rainfall inferred from oxygen and carbon isotopic fluctuations in prehistoric tropical estuarine mollusc shells. Archaeometry 37: 161– 183. Kennett, Douglas J., and Barbara Voorhies 1996 Oxygen isotopic analysis of archaeological shells to detect seasonal use of wetlands on the southern Pacific coast of Mexico. Journal of Archaeological Research 23: 689– 704. Kennett, Douglas J., Barbara Voorhies, John G. Jones, Hector Neff, Dolores R. Piperno, Thomas A. Wake, Karry L. Blake, Brendan Culleton, Josué Augusto Gómez García, and Natalia Martínez Taqueña

a gender- based model for changes

45

2007

Informe técnico final del Proyecto Arcaico-Formativo: Costa de Chiapas. Submitted to the Consejo de Arqueología, Instituto Nacional de Antropología e Historia, México. Kennett, Douglas J., Barbara Voorhies, and Dean Martorana 2006 An Ecological Model for the Origins of Maize-Based Food Production on the Pacific Coast of Southern Mexico. In Behavioral Ecology and the Transition to Agriculture, edited by Douglas J. Kennett and Bruce Winterhalder, pp. 103– 136. University of California Press, Berkeley. Marion Singer, Marie- Odile 1991 Los hombres de la selva: Un estudio de tecnología cultural en medio selvático. Colección Regiones de Mexico, Instituto Nacional de Antropología e Historia, México. Martin, M. Kay, and Barbara Voorhies 1975 Female of the Species. Columbia University Press, New York. Millán, Saúl, and Paola García Souza 2003 Lagunas del tiempo: Representaciones del agua entre los Huaves de San Mateo del Mar. Colección Etnografía de los Pueblos Indígenas de México, Serie Estudios Monográficos, Instituto Nacional de Antropología e Historia, México. Nations, James D. 2006 The Maya Tropical Forest: People, Parks, and Ancient Cities. University of Texas Press, Austin. Nations, James D., and Ronald B. Nigh 1980 The evolutionary potential of Lacandon Maya sustained-yield tropical forest agriculture. Journal of Anthropological Research 36: 1–30. Neff, Hector, Deborah M. Pearsall, John G. Jones, Bárbara Arroyo, Shawn K. Collins, and Dorothy E. Freidel 2006 Early Maya adaptive patterns: Mid-Late Holocene paleoenvironmental evidence from Pacific Guatemala. Latin American Antiquity 17: 287–315.

46

archaic to formative

Stark, Barbara L. 1981 The Rise of Sedentary Life. In Supplement to the Handbook of Middle American Indians: Archaeology, edited by Jeremy A. Sabloff, pp. 345–372. University of Texas Press, Austin. Tozzer, Alfred M. 1907 A Comparative Study of the Mayas and the Lacandones. Archaeological Institute of America. Macmillan, London. Vivó Escoto, Jorge A. 1964 Weather and Climate of Mexico and Central America. In Handbook of Middle American Indians, vol.1, Natural Environment and Early Cultures, edited by Robert C. West, pp. 187–215. University of Texas Press, Austin. Voorhies, Barbara 1976 The Chantuto People: An Archaic Period Society of the Chiapas Littoral, Mexico. Papers of the New World Archaeological Foundation No. 41, Brigham Young University, Provo, UT. Voorhies, Barbara 2004 Coastal Collectors in the Holocene: The Chantuto People of Southwest Mexico. University Press of Florida, Gainesville. Voorhies, Barbara, Douglas J. Kennett, John G. Jones, and Thomas A. Wake 2002 A Middle Archaic archaeological site on the west coast of Mexico. Latin American Antiquity 13: 179–200. Wake, Thomas A., Natalie Anikouchine, and Barbara Voorhies 2004 Food Procurement and Processing: Fish and Game Remains at the Shellmound Sites. In Coastal Collectors in the Holocene: The Chantuto People of Southwest Mexico, by Barbara Voorhies, pp. 300–343. University Press of Florida, Gainesville. Zeanah, David W. 2003 Sexual division of labor and central place foraging: A model for the Carson Desert of western Nevada. Journal of Anthropological Archaeology 23: 1–32.

THREE

Evidence for the Diversity of Late Archaic and Early Formative Plant Use in the Soconusco Region of Mexico and Guatemala Michael Blake and Hector Neff

THE PROBLEM: MAKING A LIVING IN A RICH ENVIRONMENT All of us who have worked and lived in the Soconusco region would agree that it is a land of plenty. As in many tropical regions in the Americas, it is a land where fence posts take root and sprout branches and leaves (Budowski 1987). The daily and weekly markets display a bewildering array of locally produced plants and animals. Baskets of fish from the oceans and estuaries provide a colorful array of tempting dining possibilities. Local residents often remark that the countryside is so productive that one would have to be a fool to go hungry. But this plentitude requires labor and organization in order to be realized, and this is, of course, true not just of the inhabitants of the Soconusco but of all peoples everywhere. People are always faced with the question of where to invest time and labor in the productive process The basic decisions about when to plant some seeds, tend some fruit trees, hunt deer or turtles, catch some fish, harvest palm leaves to thatch a new roof, gather a medicinal herb, and so on all require careful thought and planning. Some of the resources people

require are always available and close to hand and so take less planning to use than others. Some can be taken opportunistically—for example, walking along a trail to show a couple of archaeologists a shell mound deep in the estuary, one may find a large turtle and easily scoop it up, lash some vines around it to convert it into a living handbag, and carry it home to the family’s stew pot.1 Other resources need a greater investment in technology: fishing, for example, often requires expertly made nets, a dugout canoe or two, and the coordination of a fishing crew of five or more friends and relatives. And some resources, such as plant crops, require a sustained effort and investment in land clearing, seeding, weeding, and protection from browsers if one is to have any expectation of benefiting from all this work and planning. A perennial question in Mesoamerican archaeology has been this: what was the changing nature of this mix of productive activity over the entire course of human history in the region? This broad question has remained compelling because its answer has been tantalizingly close but exasperatingly elusive at the

47

same time. We do know that the earliest peoples to occupy Mesoamerica were hunter-fishergatherers who generally were mobile for most of the year. We also know that, at the time of the Spanish Conquest, most Mesoamericans were farmer-hunter-fisher-gatherers. The last three of these four productive components are often left out of our general descriptions and economic characterizations of these peoples, but every historic description, ethnographic account, and even present-day observation by the most casual observer shows that this truncation from farmer-hunter-fisher-gatherer to simply “farmer” is a mistake. No doubt farming was an essential part of the daily life of almost every Mesoamerican by the early 1500s when the Spanish arrived. But, it was not the only part and, even more important, the nature and mix of farming activities was highly variable from region to region. And this variation in the mix—the relative importance of different plants used and the products obtained from them, and the relative importance of farming (and in par ticular, horticulture) as opposed to hunting, fishing, and gathering— has changed in myriad ways over the course of the past 10,000 years. The archaeological record from the Soconusco is far from complete, but the parts that are emerging allow us to see that this productive diversity varied through time and throughout the region. Although our record goes back only about 6,000 years, we can see that the first known occupants of the Soconusco were concerned with the widest possible range of resources that they had at their disposal. Focusing on the plant resources, we will start by summarizing the types of plant remains that have been recorded so far. We use a time frame based on calibrated radiocarbon dates and presented in Chapter 1 (see Figure 1.3).

ARCHAEOLOGICAL EVIDENCE FOR PLANT USE At the time of the Spanish Conquest and during the first century of colonization, a very wide

48

archaic to formative

range of economically important plants was recorded for the coastal region. John Clark’s (1994:78– 79) summary of René Acuña’s (1982) analysis of various Relaciones Geográficas lists the plants that were in common use in Guatemala. Table 3.1 combines this list with those from several botanical surveys and archaeobotanical analyses along the Pacific coast of Chiapas and Guatemala.2 Research carried out in the last two decades has added considerably to our understanding of ancient human-plant interactions in the Soconusco. In summarizing the results of this work, we look at several periods and phases spanning the Middle Archaic to the beginning of the Middle Formative period where archaeological and paleoenvironmental data are now available. The individual results are fascinating, and the combined picture shows that there are clear signs of agricultural activity, particularly with the introduction of nonlocal domesticates by the end of the Middle Archaic period.3 Beginning in the Late Archaic period this activity picks up and then becomes much more prevalent in the Early Formative period. We can see these trends in the macrobotanical remains, the evidence for land clearing and burning, and the microremains of plant pollen and phytoliths. Another indicator comes from the tools used to cook, process, and serve various foods. And finally, there is evidence in the form of stable carbon and nitrogen isotope ratios in human bone remains.4 Proceeding period by period, starting with the Middle Archaic, we will summarize the data and then explore their implications. Of course, some sites have more than one class of data whereas others have limited amounts of even one type of data. Complicating matters further is the fact that some periods are much better represented in terms of both numbers of sites excavated and the diversity of remains recovered. These limitations notwithstanding, it is still instructive to examine the current information we have at hand, looking at all the lines of botanical data.

TABLE 3.1 Plant Species of Economic Importance in the Soconusco

Spanish common name

English common name

Scientific name

Plant type and Economic use

Nopal

Prickly pear cactus

Opuntia ficus-indica

Cactus: food

Chili

Chili pepper

Capsicum spp.

Small plant: food

Calabaza

Squash

Cucurbita spp.

Small plant: food

Frijol

Bean

Phaseolus vulgaris

Small plant: food

Tomatillo

Ground cherry

Physalis philadelphica

Small plant: food

Chayote

Chow chow, vegetable pear

Sechium edule

Small plant: food

Tomate

Tomato

Solanum lycopersicum

Small plant: food

Tulares

Cattail

Typha sp.

Small plant: food

Maíz

Corn

Zea mays

Small plant: food

Chía

Chia, lime-leaved sage

Salvia hispanica

Small plant: herb

Hierba mora

Nightshade

Solanum americanum

Small plant: herb

Tobaco

Tobacco

Nicotiana tabacum

Small plant: herb: medicinal

Algodoncillo

Common milkweed

Asclepias syriaca

Small plant: herb: medicinal, technological

Epazote

Wormseed, Mexican tea

Chenopodium ambrosioides

Small plant: herb: medicinal, culinary

Achiote

Anotta

Bixa orelleana

Small plant: herb: medicinal, technological

Algodón

Cotton

Gossypium hirsutum and barbadense

Small plant: technological

Hoja blanca

White leaf

Calathea lutea

Small plant: technological (cooking)

Platanillo

Lobster claw, wild plantain

Heliconia latispatha and other ssp.

Small plant: technological (cooking)

Camote, batata

Sweet potato

Ipomoea batatas

Small plant: tuber: food

Yuca

Manioc

Manihot esculenta

Small plant: tuber: food

Jicama

Yam bean

Pachyrhizus erosus

Small plant: tuber: food

Papaya

Papaya

Carica papaya

Tree fruit: food

Cacao

Chocolate

Theobroma cacao

Tree: food

Vanilla

Vanilla

Vanilla planifolia

Tree: food, spice

Coccoloba

Seagrape

Coccoloba uvifera

Tree: food, technological

Chicozapote

Sapodilla

Achras zapota

Tree: fruit

Anona blanca

Ilama

Annona diversifolia

Tree: fruit

Guanábana

Soursop

Annona muricata

Tree: fruit

Anona

Custard apple

Annona reticulata L.

Tree: fruit

Ramón

Breadnut

Brosimum alicastrum

Tree: fruit (continued )

TABLE 3.1 (continued) Spanish common name

English common name

Scientific name

Plant type and Economic use

Nanche

Nance

Byrsonima

Tree: fruit

Zapote Colorado

Mamey

Calocaroum zapota

Tree: fruit

Matasano

White sapote

Casimiroa sapota

Tree: fruit

Amate prieto

Black fig

Ficus glaucescens

Tree: fruit

Chicozapote

Sapodilla, beef apple

Manilkara zapota

Tree: fruit

Aguacate

Avocado

Persea americanca

Tree: fruit

Sapote

Mammee zapote

Pouteria sapota

Tree: fruit

Guayaba

Guava

Psidium guayaba

Tree: fruit

Jobo, ciruela

Hog plum

Spondias mombin

Tree: fruit

Jocote

Plum

Spondias purpurea

Tree: fruit

Chaya

Spurge

Cnidoscolus chayamansa

Tree: herb-shrub: medicinal

Palo de hormiga

Mexican alvaradoa

Alvaradoa amorphoides

Tree: medicinal

Zarzaparilla

Sarsaparilla

Smilax ornata

Tree: medicinal

Raíz de Michoacán

Michoacan root

unidentified

Tree: medicinal (root)

Matapiojo

Jamaican dogwood, fish poison tree

Piscidia piscipula

Tree: medicinal, technological

Pacaya, tepejilote

Pacaya palm

Chamaedorea tepejilote

Tree: palm: food

Pejibaye

Peach palm

Bactris sp.

Tree: palm: food, technological

Palma real

Fan palm

Sabal mexicana

Tree: palm: technological

Corozo

Corozo nut

Scheelea preussi

Tree: palm: technological

Coyol

Coyol palm

Acrocomia mexicana

Tree: palm: technological, food

Espino blanco

Fernleaf acacia

Acacia pennatula

Tree: technological

Palo mulato

Gumbo-limbo

Bursera simaruba

Tree: technological

Hule

Rubber tree

Castilla elastica

Tree: technological

Cupapé

Siricote

Cordia dodecandra

Tree: technological

Jicaro

Calabash, gourd tree

Crescentia cujete

Tree: technological

Roble

Oak

Quercus oleoides

Tree: technological

Guanacaste

Elephant ear tree

Enterolobium cyclocarpum

Tree: technological, food

Cacahito

Sandpaper tree

Curatella americana

Tree: technological, medicinal

Guapinol, Jatoba

Locust, Algarrobo

Hymenaea courbaril

Tree: technological, medicinal

Bejuco

Climbing vine

Tetracera volubilis

Vine: medicinal, technological

sources: Breedlove 1981; Clark 1994; Coe and Flannery 1967; Eccardi and Álvarez del Toro 1987; Feddema 1993; Helbig 1964; Hernández Bermejo and León 1994; Jones and Voorhies 2004; Miranda 1976; Otero-Arnaiz et al. 1999; Paillés H. 1980; Voorhies 1976.

Middle Archaic (pre-3500 cal b.c.) Cerro de las Conchas We have a glimpse of the Middle Archaic period thanks to the excavations at Cerro de las Conchas, a shell mound site located on the edge of Cantileña (or Hueyate) swamp in the Huixtla region of the Soconusco (Voorhies 2004). The site dates from about 5500 cal b.c. to about 4000 cal b.c. when, according to Voorhies (2004:97), “the deposition of pure shell by Middle Archaic Period people ceased here.” At Cerro de las Conchas a series of soil samples from test pits and trenches recovered phytoliths (pollen was entirely absent) representing heavy forest coverage, especially by members of the families Chrysobalanaceae (many genera of flowering trees and shrubs), Fabaceae (leguminous plants including trees and shrubs), and Moraceae (with at least forty genera, including trees such as fig and mulberry) (Jones and Voorhies 2004:303–307). Jones and Voorhies also note indeterminate phytoliths from the Marantaceae (arrowroot) family, Bactris-type palms, Sabal-type palms, and various grasses. They suggest that during the course of the Middle Archaic at the site, there was some mechanism, probably natural, that led to a reduction in the forest species and an increase in palms and grasses. No phytoliths that could be unambiguously identified as cultigens were recovered from the Middle Archaic deposits at this site.5

Late Archaic The Late Archaic period is much better represented than the previous period. This may be partly because of natural deposition processes or simply because the Middle Archaic peoples were few and far between— and they trod much more lightly on the ground than their Late Archaic successors. But the actual number of Late Archaic occupation sites is still very low. In the Acapetahua zone of Chiapas, Voorhies (2004) documents five shell mounds that have been discovered since the mid-1940s. In addi-

tion, there is one inland, non-shell-mound site. And that is all. The state of Late Archaic affairs in Guatemala is somewhat more dismal. Neff and colleagues (2006a, 2006b) recently conducted a core sampling program of nonarchaeological estuary zones along the Pacific coast of Guatemala in part because “as yet not a single Archaic period archaeological site has been located on the Guatemalan Pacific coast, despite extensive exploration both within the mangrove zone . . . and on the coastal plain. . . .” (Neff et al. 2006a:293). Thanks to paleoenvironmental testing at several locations along the coast of Guatemala as well as at a small number of excavated archaeological sites in Chiapas, we now have some idea of the patterns of human-plant interaction in the Soconusco after 3500 cal b.c. In the Acapetahua zone at the northern end of the Soconusco, Jones and Voorhies (2004) provide a detailed summary of all of the paleoethnobotanical materials recovered to this point, including “exceedingly scarce” macrobotanical remains, even rarer pollen, and the (fortunately) much more abundant phytolith remains. Their summary of the pollen and phytoliths from the shell mound sites of Islona Chantuto, Tlacuachero, and Zapotillo and the inland site of Vuelta Limon show a number of fascinating patterns. Islona Chantuto The only pollen samples came from a core at the base of the Islona Chantuto mound. As one might expect, Jones and Voorhies found that the pollen assemblage was dominated by species common in a coastal mangrove estuary system, including Rhizophora and Avicennia (mangroves), Nymphaea (water lily), Typha (cattail), Cyperaceae (sedges), Alismataceae (water plantains), and Combretaceae-Melastomataceae (large families of flowering trees, shrubs, lianas, and herbs) (Jones and Voorhies 2004:309). They also found pollen from various families of tropical trees that they think may have been living on some of the islands located within the

evidence for the diversity of plant use

51

wetlands: Moraceae (mulberry or fig), Fabaceae (legumes— a large family that ranges from trees to beans), Coccoloba (seagrape), and Arecaceae (palms). Some long-distance wind-borne pollen from several genera of trees was also recovered: Pinus (pine), Quercus (oak), and CTC (Taxodiaceae, Cupressaceae, and Thuja— all genera of evergreen conifers). In the lowest zone of the core (Zone 1a), dating from about 2000 cal b.c. to 1850 cal b.c., the researchers found a relatively low frequency of Cecropia pollen (a tree genus that is thought to be part of the Urticaceae or nettle family). Beginning about 1850 cal b.c. and extending to about 1690 cal b.c., the frequency of Cecropia pollen increased, suggesting to the analysts a corresponding increase in forest disturbance and clearing activity (see the Early Formative section in this chapter).

bance and some evidence of one cultigen: Zea mays. There are more phytoliths of grasses, Cyperaceae (sedge), and Marantaceae (arrowroot family) and fewer Chrysobalanaceae (flowering trees and shrubs) and Sabal-type palm— all indicators of increased forest clearance. Three instances of large cross-shaped phytoliths that are interpreted to be Zea mays were found in samples above the clay floor, and one of the floor samples had a single maize phytolith (Jones and Voorhies 2004:323). Apart from these few phytoliths identified as Zea, however, there were no other phytoliths that were identified as coming from other domesticated plants. But several taxa that have economic importance are represented: palms for thatching, woody species for fuel and building, and a possibly imported Chusquea or bamboo. Vuelta Limon

Tlacuachero At the Tlacuachero shell mound site, Jones and Voorhies also looked for pollen, but because of oxidation, found none. Instead, they were able to recover a robust phytolith data set from fourteen samples spanning 7 m of deposits. In the earlier levels that pre- date a large and elaborate clay floor construction event, the phytolith assemblage led the authors to suggest the site was located within a relatively intact evergreen tropical forest (Jones and Voorhies 2004:319). This is indicated by high counts of Chrysobalanaceae-type (flowering trees and shrubs) and Sabal-type (palm) phytoliths, many of which were burned. They suggest that these plants may have been used as fuel. Most important, in the layers below the clay floor there are few grass and Cyperaceae (sedge) phytoliths represented. This is consistent with the idea that the forest was little disturbed at this time. Even more noteworthy is the complete lack of phytoliths that could be considered to have come from cultigens. The deposits on and above the clay floor structure, which dates to between about 3600 cal b.c. and 3100 cal b.c. (Voorhies 2004:66), contain much more evidence of forest distur-

52

archaic to formative

Phytolith samples were also collected and analyzed from the Late Archaic Stratum E at the inland site of Vuelta Limon. Very small numbers of Zea mays phytoliths were found in seven samples across Stratum E, along with phytoliths from grasses, shrubs and trees, sedges, bamboo, and palms (Jones and Voorhies 2004:334). The small numbers of Zea along with the lack of any other cultigens in the Late Archaic deposit present a picture of marginal agricultural activity at that site. This, combined with the fairly disturbed nature of the contact zone between Stratum E and the layer above it (Voorhies 2004:107, figure 2.45) call into question the nature of the association of the Zea phytoliths with the Late Archaic deposits: could some of the phytoliths in the Stratum E samples be intrusive as a result of mixing caused by animal burrowing? This possibility is amplified by the one radiocarbon date from a piece of charred wood found in Stratum E that dated between 2030 cal b.c. and 1530 cal b.c. (2 sigma) (Voorhies 2004:115). Voorhies points out that this time range spans the dating hiatus that Blake et al. (1995) had noted for the Soconusco region. Clark’s recent adjustments to the dating of the Early Formative phases puts the

beginning of the Barra phase at approximately 1900 cal b.c. (Chapter 1, Figure 1.3), and this now suggests that the charred wood sample from Vuelta Limon could come from any time spanning the last century of the Late Archaic period to the end of the Locona phase. It seems possible then that Vuelta Limon is a candidate for a site that spans the transition between Late Archaic and Early Formative, as people started to make more intensive use of the landscapes inland from the estuary zone (Clark and Cheetham 2002). Voorhies (2004:404) and Kennett and Voorhies have suggested that during the Middle Archaic and early part of the Late Archaic, people were episodically occupying shell mounds in the estuary during both the wet and dry seasons. After about 3500 cal b.c., stable oxygen isotope ratios in clam shells at Tlacuachero suggest that there was a gradual shift to wetseason- only use of the site. This might correspond to increasing use of inland sites and an increased reliance on plant foods, “leading to the eventual development of agroeconomies in the region” (Kennett, Voorhies, and Martorana 2006; Voorhies 2004:404). Guatemala Recent work by Neff et al. (2006a, 2006b) documents the changing climate and depositional history in wetland locations along the coast of Guatemala (two in the southern end of the Soconusco and one farther south). Three core locations are all in mangrove forests within 20 km of the Mexico- Guatemala border: Lower Río Naranjo (cores OC001and TIL016) and Manchón (MAN015). Another set of cores was taken from the Sipacate locale about 120 km southeast of the border along the coast (SIP001, SIP014, and SIP99E). None of the core sequences are from archaeological sites, although it is quite likely that settlements, both seasonal and permanent, were located nearby. Pollen and phytoliths were recovered from most of the cores and produced a detailed radiocarbon sequence dating the deposits as early as 4753 cal b.c. and extending up to about

410 cal b.c. (although there were some earlier and later dates in some sequences that can be ignored in this analysis). lower río naranjo In the Lower Río Naranjo cores Neff and colleagues noted a wide range of taxa represented in both pollen and phytoliths, including the following broad classes: mangroves, aquatic species, herbs, and arboreal species. In core TIL016, phytoliths of Zea mays were much more common than those of Zea pollen (one grain only), and phytoliths of other economically important taxa were also identified: Cucurbitaceae (squash), Cannaceae (roots), and Marantaceae (arrowroot). At Manchón similar economic plants were also represented in the phytolith assemblage (with only a few traces of maize but no squash). Manchón’s pollen assemblage included one grain of Gossypium (cotton), but no maize or squash. Both Lower Río Naranjo and Manchón had a number of economically important tree species present, including possible fruit trees in the Annonacaea family (Neff et al. 2006a). sipacate Among the Sipacate cores, two had sequences extending for the duration of the Late Archaic period (ca. 3700 cal b.c.– 1900 cal b.c.). The pollen from these cores included three grains of Zea mays as well as many genera of herbs and trees that could have had significant economic roles, such as Cheno-Ams, Typha, and Coccoloba. The phytoliths, on the other hand, showed the same range of cultigens and economic plants that were recorded for the cores from the Lower Río Naranjo and Manchón and included Cucurbitaceae, Cannaceae, and Marantaceae phytoliths. In their summary of the Sipacate data (particularly cores SIP001 and 014), Neff and colleagues note that mangrove forests began to dominate by about 4000 cal b.c. and “persisted until around 3400 or 3500 cal b.c., during which time tropical forest with palms (Arecaceae) and other trees, such as Coccoloba, predominated on the dry land adjacent to the mangroves” (Neff et al. 2006a:297).

evidence for the diversity of plant use

53

One of the most dramatic findings in the Sipacate cores is that at the beginning of the Late Archaic, around 3500 cal b.c. (as defined for the Chantuto sequence already mentioned), there is a “dramatic” environmental change. Mangroves and other tree taxa declined rapidly and disturbance indicators such as cattails and sedges increased markedly (Neff et al. 2006a). Even more striking is the massive increase in charcoal fragments. This persisted for many centuries until mangroves started to rebound toward the last part of the Late Archaic, after about 2500 cal b.c. By the beginning of the Early Formative period, mangroves may have returned to previous levels only to face another bout of deforestation corresponding to the increasing settlement of the area during that period. Neff et al. (2006a:298) argue that this episode of deforestation is associated with early horticulture—the first wave of which began just after 3500 cal b.c. and a second just after 1700 cal b.c. While this is a useful hypothesis, independent lines of evidence—for example, Late Archaic archaeological sites with additional traces of horticultural activity—would be valuable at this point for making the case. At Manchón, the decline in mangrove forest around 3000 cal b.c. and the corresponding rise in open forest plant taxa are interpreted as evidence of a drying period rather than evidence of human land clearing— especially because there is only a slight increase in charcoal and very little evidence of cultigens in the pollen or phytolith assemblages. In the Lower Río Naranjo cores, the general pattern is one of increasing human impact, including the presence of cultigens and mangrove forest decline beginning around 2600 cal b.c. As mentioned for Sipacate to the south, this sequence in the Lower Río Naranjo cores suggests that there were environmental changes that may have been caused by human clearing— especially burning— and that this is a signal of horticulture during the Late Archaic. The undisputed evidence of traces of Zea mays pollen and phytoliths confirms that people were doing more than just burning the mangrove forests

54

archaic to formative

to increase hunting potential in cleared landscapes—they were also cultivating nonnative plants, including at least maize and squash. But again, the nature and intensity of this activity cannot be firmly evaluated without excavations of the settlements where these peoples lived— even if just seasonally. In the face of a lack of such evidence we are left with the impression of small populations of hunting-fishinggathering people who practiced horticulture in favorable locations along the inland edges of the mangrove swamps. Populations living along the coast of Guatemala must have been fairly small in number and remained so for the entire Late Archaic. Population growth must have been slow for millennia; this is perhaps related to the nature of seasonal availability of plant and animal foods, their low densities, or their unpredictability from year to year. In areas where human activity produced a visible record, such as the shell mound sites in the Acapetahua and Cantileña zones, the traces of Late Archaic peoples are easier to spot. Without these mounds of shell as testaments to their ancient economies, they too would be as invisible as the Late Archaic peoples in the southern part of the Soconusco.

Early Formative The appearance of pottery in the archaeological record of the Soconusco marks a major increase in the visibility of the region’s ancient inhabitants. Pottery is correlated with settlements, both coastal and inland, that have evidence of houses in village clusters with the remains of a much wider range of activities represented than we can perceive for the Late Archaic. Excavations at several sites have recovered botanical remains that confirm some of the plant use  patterns that are apparent for the Late Archaic—increasing evidence of horticulture, land clearing, and use of cultigens, many of which were originally domesticated outside the region. We will summarize some of the findings from the Mazatán zone where excavations

at several Early Formative sites have recovered macrobotanical remains of economically important plant species. We will also continue with the sequences of microfossil remains for Acapetahua and Guatemala. Mazatán Zone One of our best glimpses of Early Formative human-plant interaction comes from the paleoethnobotanical analysis of charred macrobotanical remains from the Mazatán zone carried out by Vicki Feddema (1993) for her Master’s thesis research.6 Here we can divide the samples into phases within the Early Formative sequence and look at a much more fine-grained chronology than has been possible for the Late Archaic—thanks almost entirely to the presence of pottery. Feddema’s analysis presents the macrobotanical data from 132 flotation samples from hearths, trash pits, floors, and midden deposits at four sites (Aquiles Serdán, Paso de la Amada, Chilo, and San Carlos) spanning five phases from 1900 cal b.c. to 1200 cal b.c. (Barra, Locona, Ocós, Cherla, Cuadros) (see Chapter 1, Figure 1.3).7 Only six taxa were identified in the flotation samples: Zea mays, Phaseolus spp. (beans), Persea americana (avocado), Mollugo sp. (carpetweed), Polygonum sp. (knotweed), and Brassica sp. (mustard). The fi rst three genera (maize, beans, and avocado) were the only cultigens recovered and were also the most common to occur in the samples. The other three genera are represented by charred seeds that occurred in very small numbers and were probably of little, if any, economic importance—instead, they are probably related to weedy invasion in and around the settlements (Mollugo = 19; Polygonum = 1; Brassica = 2). The most ubiquitous of the cultigens was maize, which occurred in all four sites and 81 of 132 samples. A total of 2,279 fragments of maize was identified; most of this consisted of kernels and cupules, but several cob fragments were also recovered. Table 3.2 shows the distribution of the samples through time in all of the four sites combined. We calculated the ratio of

the number of maize fragments to the total weight of charcoal fragments (excluding the unidentified charred seed fragments) in order to control for the variable size of different flotation samples. Using this measure, rather than the raw counts, we see that there is a wide range in the ratios of maize to charcoal in the five phases. There is clearly no pattern of change— no doubt a matter of small sample sizes and differential preservation of specimens in different contexts. As previously reported (Clark 1994:547), Accelerator Mass Spectrometry (AMS) radiocarbon dating was used to check the absolute ages of seven maize samples from relatively secure contexts. The oldest of the group, from a Barra phase feature at San Carlos, dated to between 1869 cal b.c. and 1508 cal b.c. (95.4 percent; Beta-62911, median = 1654 cal b.c.). The most recent, from a Cuadros-phase level in a trash pit feature at Aquiles Serdán, dated to between 1400 cal b.c. and 1056 cal b.c. (95.4 percent; Beta-62920, median = 1243 cal b.c.).8 The other five maize AMS dates correspond nicely with their estimated ages, which are based on ceramic associations. The next most ubiquitous cultigen recovered from the flotation samples was Phaseolus sp. (beans). As Table 3.2 shows, 257 bean fragments were recovered from Locona, Ocós, Cherla, and Cuadros deposits, and none were recovered from the Barra phase deposits. The ratio of beans to charcoal is rather constant, except for the Cherla-phase samples, where there is a threefold increase. Again, the fluctuations in frequency probably do not say as much about changing use through time as they do about differences in the conditions of deposition and preservation. So far, only one bean fragment has been AMS dated: a specimen from a Locona midden deposit in Mound 6 at Paso de la Amada (Clark 1994:547). It dated to between 1516 cal b.c. and 1269 cal b.c. (95.4 percent; Beta- 62914, median = 1411 cal b.c.)—placing it with the greatest likelihood in the Ocós phase, slightly later than its inferred Locona phase ceramic

evidence for the diversity of plant use

55

TABLE 3.2 Maize, Bean, and Avocado Fragments From Four Early Formative Sites in the Mazatan Region

Phase

No. Samples (all sites)

No. maize frag.

maize/gm charcoal

No. bean frag.

bean/gm charcoal

No. avocado frag.

avocado/gm charcoal

grams charcoal

Cuadros

10

12

3.40

6

1.70

2

0.57

3.53

Cherla

22

153

43.84

21

6.02

4

1.15

3.49

Ocos

55

1521

17.69

198

2.30

12

0.14

85.96

Locona

39

510

36.07

32

2.26

1

0.07

14.14

Barra

6

83

218.42

0

0.00

0

0.00

0.38

Totals

132

2279

257

19

107.5

note: After Feddema 1993.

association. It is worthwhile pointing out that the frequency of bean recovery from the four Mazatán zone sites is one of the highest we know of for the Early Formative period anywhere in Mesoamerica—in fact, these are the earliest Phaseolus specimens reported for Mesoamerica to date. And, the co- occurrence of maize and beans as the most ubiquitous identified plant remains at these sites reinforces the notion that these two of three members of the famous triumvirate (the third is squash, which is missing from the Mazatán assemblage but present elsewhere in the Soconusco) were certainly in place by the Early Formative period. Finally avocado remains (Persea americana) represented the other most frequent occurrence of economically important plant species. However, with only nineteen fragments of charred avocado pits in the Locona through Cuadros deposits, these are considerably rarer than maize and bean remains. As with beans, avocado pit fragments were somewhat more common than charcoal fragments in Cherla phase deposits (Table 3.2)—but almost all of the specimens came from just one site: Aquiles Serdán. Again, this is likely a question of preservation, although it is possible that people at this site were growing more avocado trees than their neighbors were. The presence of charred avocado pit fragments in household refuse deposits is impor-

56

archaic to formative

tant because it confirms the use of tree fruits (whether or not the trees were actually domesticated) in the subsistence economy of the region. As Table 3.1 indicates, there are many tree genera, several of which bear edible fruits that have become important in the region in recent times. Avocados, as a member of the Lauraceae family, are native to the tropical regions of Middle America, and their seeds have been found in many Early Formative contexts in Mesoamerica— Oaxaca, Tehuacan, and elsewhere— so it is not unexpected that they would be present in the early village remains of the Soconusco. In fact, they are so common that we can propose that the triumvirate be expanded to Mesoamerica’s “Big 4.” We should also point out, however, that no Lauraceae pollen or phytoliths were identified in either the Early Formative or Late Archaic samples previously discussed for the Acapetahua, Lower Río Naranjo, Manchón, or farther afield, the Sipacate zones—but this may be because the pollen does not travel far. Evidence has been reported recently that suggests cacao (Theobroma cacao) may have been present in the Soconusco during the Early Formative period. Powis and colleagues (2007, 2008) tested sixteen ceramic potsherds from Early Formative site of Paso de la Amada in the Mazatán zone for residues of T. cacao and found a clear chemical signal on one of the sherds. The single sherd is in a style that dates

to the Barra phase and was recovered in early Locona phase construction fill between Structures 3 and 4 in Mound 6 at Paso de la Amada. Theobroma is a genus in the Sterculiaceae family of tropic flowering trees and, as with the previously mentioned Lauraceae family, it also has not been reported in the Archaic or Early Formative pollen and phytolith samples collected in the Soconusco region—again, in the case of pollen, probably because it does not travel very far. The issue of whether cacao plants are native only to South America and domesticated varieties were introduced into Mesoamerica or whether some remnant wild stands existed in the tropical forests of southern Mesoamerica and were eventually domesticated locally is still unresolved (Henderson et al. 2007; McNeil 2006). We think that if cacao trees were cultivated, or at least tended, as part of the Late Archaic tropical forest complex in the Soconusco, we might expect to find phytoliths and pollen evidence of the Sterculiaceae family in association with settlements of that period. El Varal Richard Lesure’s excavations at El Varal, located in an estuary setting about 3 km southeast of the mouth of the Río Coatan, also recovered plant macroremains (Lesure 2009; Popper and Lesure 2009). The deposits date to the Cuadros and Jocotal phases and so are contemporary with CAP-78 (see the section on CAP-78 in this chapter). Virginia Popper analyzed five flotation samples from El Varal and compared these remains to samples from three other sites in the Mazatán zone, analyzed for the first time: Cuadros-phase Canton Corralito (San Carlos), Locona- or Ocós-phase Paso de la Amada, and site Mz-250. In all these sites very few preserved macroremains were recovered, and few could be identified with certainty. However, Popper and Lesure do note the recovery of Zea mays from El Varal (two cupules), Canton Corralito (one cupule and three kernel fragments), and Paso de la Amada (one cupule). Two other notable discoveries were made: at Canton Corralito they report a possible Fabaceae specimen

(legume) and at Mz-250 there were 0.24 grams of nut fragments. The rest of the botanical material from these sites was unidentifiable. One of the important aspects of this analysis is that, as with the previous discussion of the Mazatán sites, Zea mays, although represented by only seven identified examples in more than 300 liters of floated soil samples, was the most common identified plant. Cuauhtémoc Robert Rosenswig’s excavations at the Cuauhtémoc site, located approximately 40 km southeast of the Mazatán zone, near the Suchiate River, which forms the border between Mexico and Guatemala, have yielded a wide range of plant remains spanning the Early Formative period and extending into the Middle Formative (Rosenswig 2010). The analysis of the plant materials is still under way, but preliminary identifications of maize show similar patterns as already noted for the Mazatán sites. Amber VanDerwarker, who is currently working on the Cuauhtémoc plant materials, has recovered maize in all “fifteen flotation samples from pit features and middens dating to the Locona through Conchas phases” (Rosenswig 2010:137). The trend that shows up in these samples is a constant low-level presence of maize from the Locona to Ocós phases, then an abrupt increase in the ratio of maize fragments to excavated volume in the Cherla phase (Rosenswig 2010:144, figure 5-4). All of the later Conchas phase sample units had higher ratios of maize than most of the preceding Early Formative units, suggesting that maize growing (and perhaps reliance on agricultural production in general) became more important in Middle Formative times, as Blake et al. (1992) found in their stable isotope study of human bone samples from other Soconusco sites. Islona Chantuto Returning to the microbotanical samples from the Chantuto sites, Jones and Voorhies (2004) report on a range of samples that date to the Early Formative period. As already discussed in the Late Archaic section, the earliest zone in the

evidence for the diversity of plant use

57

Islona Chantuto pollen core (labeled the Chantuto 4 sediment core) was subdivided into an upper and lower zone (Zone 1a-lower and Zone 1b-upper). The upper zone had a much higher frequency of forest disturbance indicators than the lower zone. Beginning at about 1850 cal b.c.—just after the start of the Barra phase— pollen from Cecropia, Cheno-Ams (amaranths including flowering herbs and small shrubs), sedges, and grasses all increased. Jones and Voorhies (2004:310–311) suspect that this change, probably the result of forest disturbance, indicates increased reliance on cultigens. There was, however, no pollen from cultigens in this zone, so it is possible that the clearing and disturbance might have resulted from a number of activities unrelated to agriculture per se. Vuelta Limon The rich phytolith record from Vuelta Limon spans part of the Early Formative period. Voorhies (2004:114– 115) estimates, on the basis of ceramic associations, that the lower portion of Stratum C was deposited during the Early Formative Barra and Locona phases. This stratum contains some instances of Zea mays and Cucurbita phytoliths, the only definite cultigens, but also present are some forest phytolith types (trees, shrubs, herbs) that may have had economic value: Annonaceae, Arecaceae, Bactris, Chrysobalanaceae, Chusquea, Heliconia, Marantaceae, and various palms (Jones and Voorhies 2004:328–329). CAP-78 Voorhies (Jones and Voorhies 2004) collected phytolith samples from the site of CAP-78, about 1 km downstream from Vuelta Limon on the Río Cacaluta. On the basis of ceramic associations, Voorhies (2004:117) estimates that this occupation dates to the Cuadros and Jocotal phases.9 Just as at Vuelta Limon, the range of plant phytoliths recovered indicates cultivation of Zea mays, and Cucurbita. But, unlike Vuelta Limon, the site environs appear to have been relatively deforested, and they have a much higher percentage of disturbance indica-

58

archaic to formative

tors (grasses and sedges). After abandonment, forest species return— even though many of them are of potential economic use. Salinas la Blanca In Guatemala there are very few examples of sites that have produced well-preserved macrobotanical remains. One notable exception comes from the excavations at Salinas La Blanca carried out in the 1960s by Michael Coe and Kent Flannery (1967). They reported finding the impressions of maize cobs, stalks, and leaves both on and in clay floors dating to the Cuadros phase (1967:71). In all, more than fifty impressions of cobs were found, and some of these contained the actual mineralized remains or “fossils” of the cobs. Paul Mangelsdorf, who analyzed these samples, found that they were most like the Chapalote and Nal-Tel races of maize, common to Mexico and Guatemala, and characterized them as a pod corn (Mangelsdorf 1967). Other plant impressions were also recovered from Salinas la Blanca. All of them appear to be seeds of tree fruits. Coe and Flannery (1967:72– 73) report finding a seed (possibly two) of jocote (Spondias purpurea), one matasano (Casimiroa sapota), three avocado pits (possibly four) (Persea americana), and at least three other species of tree fruit seeds, all unidentifiable. These findings are often overshadowed by the mineralized maize remains that are undeniably significant in their degree of preservation and their quantity. However, the tree fruits are also significant because they are species common in other regions of Mesoamerica and continue to be used in house orchards and garden plots throughout the Soconusco region today. These examples are not carbonized, so we would likely never have known of their existence if it were not for the unique conditions that led to their impressions being left in the clay floors. Guatemala Core Samples Finally, many of the paleoenvironmental cores reported by Neff and colleagues (2006a, 2006b) extended well beyond the Late Archaic and into the Early and Middle Formative periods. Sipa-

cate 001, and 99E, Manchón 015, and TIL016 all had similar assemblages represented for the Early Formative period and already described herein for the Late Archaic. In some cases, evidence of forest clearing that began around 3500 cal b.c. (for example, SIP001) continued on throughout the core sequence. In other cases, charcoal frequencies increased during the Late Archaic (for example, TIL016), while in yet others the increase in charcoal appeared near the beginning of the Early Formative (MAN015) or partway through it (SIP99E). Maize is represented either in pollen or phytolith remains in most samples, and squash is present in some (Neff et al. 2006a).

THE CHANGING ROLE OF AGRICULTURE The bulk of this chapter so far has been devoted to presenting summaries of the basic information on known plant remains from archaeological and paleoenvironmental studies in the Soconusco. The data, as presented, highlight the incomplete, sporadic, and differentially preserved nature of the archaeobotanical record. This, however, is no different from any class of archaeological remains (with the exception perhaps of ceramics and lithics—the evidence of which gives most archeologists the impression that analysis of these two classes of data can answer three-quarters of all questions we may have about past cultures). So, in spite of the desultory nature of the archaeobotanical remains, we are bound to seek and, ultimately find, interpretable patterns. The data that we have presented up to this point are summaries of rather fine-grained analyses of samples collected in precise archaeological contexts dispersed throughout the region and through time. Some of the welldocumented sequences highlight the microbotanical (pollen and phytolith) evidence for cultivation and plant use by nonsedentary people centuries before the appearance of the first permanent villages during the Barra Phase (ca. 1900 cal b.c.). Other examples document the macrobotanical evidence for cultigens, includ-

ing maize, beans, and avocados among the earliest villagers. Taken together, these fine-grained data suggest the cultivation, management, and use of a wide range of plants that formed part of the complex coastal subsistence system. Cultivation of tree species such as avocado, cacao, and various tropical fruits along with the use of a wide array of other native plant species may have allowed Late Archaic hunter-fishergatherers a high degree of mobility, while at the same time permitting the gradual introduction of and then the increasing reliance on annual crops such as maize and beans during the Early Formative period. To perceive larger-scale patterns, we aggregate the material into broader categories of spatial and temporal units and discuss the presence or absence of plants by general type. To see if there are any intraregional trends, we divide the sample into north, central, and south coast and within each of these we further subdivide into the simple categories of estuary vs. inland. The northern group includes the Chantuto zone, the central group consists of the Mazatán zone, and southern group includes the sites collection locales in Guatemala. Within each group we separate the sites located in or on the edge of estuaries near the ocean from those located farther away from the ocean and not adjacent to estuary systems. These are somewhat arbitrary distinctions but may be useful for a first-run comparison. The next step in aggregating the data is to group sites by temporal affiliation. The Middle and Late Archaic deposits remain separated, but the Formativeperiod deposits are grouped into initial Early Formative (Barra through Cherla) and later Early Formative (Cuadros and Jocotal). Using these categories, Table 3.3 presents the summary of plant families, genera, and in some cases species represented in the archaeological record of the Soconusco. Admittedly, we are painting the picture with broad brush strokes, but the aim here is to see what the picture looks like when all the data are combined. The main pattern that emerges is the remarkable number of gaps still in the data. There

evidence for the diversity of plant use

59

may be various reasons for the gaps: no sites have yet been found in a subregion for a given time period, no excavation has taken place, no archaeobotanical samples were collected during excavation, or no archaeobotanical samples were preserved. As serious as these gaps are, they do point to specific time periods and regions where more sampling would be extremely productive (for example, Later Early Formative phases, Middle Archaic period, and all inland subregions except the central zone, which is well represented). We would also point out that while there are excellent pollen and phytolith samples for the northern and southern zones, we do not have comparable samples from the central region. Fortunately this is compensated for, in part, by the relatively large numbers of plant macroremains from four sites in the inland subregion of the central zone (the Mazatán zone). One of the reasons it will be important to recover pollen, phytoliths, and macroremains from all zones and time periods is that they provide very different perspectives on the presence and absence of plants in each area and on the level of specificity. While the microremains give us a good picture of the general sweep of plant varieties present, it is often difficult to know— except in the case of nonnative cultigens such as maize, beans, and squash—whether the plants that are present were actually being used by the ancient inhabitants of the region. One of the key patterns apparent in Table 3.3 is that maize first appeared in the Late Archaic period. And, although it is not clear from the summaries presented in Table 3.3, the finegrained analyses from both the northern estuary zone (Chantuto—Jones and Voorhies 2004) and the southern estuary zone (Guatemala— Neff et. al. 2006a, 2006b) show that the evidence for maize in this period consisted of very small numbers of pollen grains or phytoliths. We can assume, based on this, that people were either growing or importing maize into these estuary zones. The evidence for other cultigens is even sparser than for maize: there are cucurbit (squash) microremains and, in one sample,

60

archaic to formative

there is cotton. Clear evidence of beans is absent. The relatively small numbers of cultigen microremains in these samples may not give an accurate representation of how economically important these plants were. Pollen from these species does not necessarily travel very far from its source— although maize and teosinte pollen can become windborne and travel well over 100 to 200 m or more in turbulent wind conditions (Aylor, Baltazar, and Schoper 2005). In the case of the Guatemalan core samples that were collected off-site, in areas that were swamps at the time, it is remarkable that any pollen from these cultigens made it into the record. That it did indicates that there must have been much more agricultural activity in the zone than we can at present perceive directly from other kinds of remains— such as those found in the Chantuto zone. Neff and colleagues (2006a:305) suggest that these Late Archaic occupants of the Guatemalan zone were relatively mobile and may have repeatedly occupied locations, burning and clearing forests as they used the region’s resources. It is quite likely— given the low level of agriculture at this time, the sparse population densities, and the complete lack of permanent settlements—that people practiced burning to encourage the growth of economically useful wild plants and to attract animals, rather than primarily to plant crops. These tactics are reported for almost all nomadic hunting and gathering peoples in the ethnographic literature, as well as for part-time horticulturalists. Burning is a particularly effective and efficient way to promote an abundance of economically useful plants and animals in dense forest environments (Piperno 1994; Piperno and Pearsall 1998). At the same time, it is very likely that people would have been interested in a wide range of tree species that could provide useful products: fruit and leaves, bark (medicines, dyes, fibers), roots (medicines), building materials (roofing, dwelling posts, dugout canoes, fish traps, tools of all sorts). They would not have wanted to deforest a region. Instead they

TABLE 3.3 Aggregated Plant Data by Time Period and Subregion

North

Central

South

Period

Estuary

Inland

Estuary

Inland

Estuary

Inland

Later Early Formative (Cuadros and Jocotal)

No data

No data

Maize

Maize, legumes

Maize, avocado, No data jocote, matasano, (and all families listed below)

Maize Initial Early Formative (Barra, Locona, Ocos, Cherla)

No data Maize, cucurbita, annona family, Sabal and Bactris palms, flowering trees and shrubs, bamboo, water plantain, arrowroot family

Maize, beans, avocados, nuts, cacao

(All families listed below)

No data

Late Archaic (Chantuto (A)

Maize, arrowroot family, palms, water plantain, water lily, cattail, flowering trees and vines, bamboo, mulberries and figs, legumes, seagrape

Maize, arrowroot No data family, palms, water plantain, water lily, cattail, flowering trees and vines, bamboo, mulberries and figs, legumes, seagrape

No data

Maize, palms, Cheno-Ams, cattails, seagrape, squashes, Canna root family, arrowroot family, cotton, annona family

No data

Middle Archaic (Chantuto (B)

No data

No data

No data

No data

No data

would have been most interested in changing the ecological balance to foster the mix of plants and animals that were most useful to them (Ford and Nigh 2009; VanDerwarker 2006:110– 111). In the Chantuto zone, the pattern is somewhat different because the huge swamp systems

Legumes, figs and mulberries, arrowroot family, Bactris and Sabal palms

favored the exploitation of aquatic resources, including shellfish, the most visible if not the most nutritious of these resources during the Archaic period. Voorhies (2004) and Kennett, Voorhies, and Martorana (2006) have addressed the issue of integrating resources such as shellfish into emerging agricultural systems, so we

evidence for the diversity of plant use

61

will not review their arguments here. But it is worth noting that the evidence for maize or any other cultigens providing significant contributions to the Late Archaic diet is still minimal. One possible exception to this is the stable carbon isotope analysis of one individual associated with the clay floor at Tlacuachero—the high-value stable carbon value suggested that C4 plants (possibly maize) did contribute significantly to the diet (Blake et al. 1992; Chisholm and Blake 2006; Voorhies 2004), although high marine food (for example, fish) can also lead to elevated stable carbon isotope ratios.10 Following this line of analysis, another key pattern apparent in Table 3.3 is that during the initial Early Formative, the estuary zones, both north and south, preserve a continuing record of diversity in plant species present in and near settlement locations. These records, based almost entirely on pollen and phytolith evidence, show that there was a great deal of continuity in the kinds of plant resources that were available and probably in use between the Late Archaic and the initial Early Formative. This is a significant, and often overlooked, pattern. It suggests that the estuary environments may have continued to be seasonally or sparsely occupied during the Early Formative, as they likely were during the Archaic. Neff and colleagues (2006a:308) point out that the Early Formative mounds along the Guatemalan Coast show discontinuous occupation: periods of building are interspersed with episodes of abandonment, suggesting a relatively mobile subsistence-settlement system. This sharply contrasts with the archaeological record at many of the Early Formative sites in the inland portion of the Mazatán zone, where all sites show evidence of continuous occupation, building, and rebuilding of structures (Blake 1991 and Chapter 5 in this book; Clark 1994; Lesure 1995). It may well be that during the Early Formative period more attention was paid to agricultural production in the inland regions than to that in the estuary zones, a pattern that certainly continues into the present. A portion of the inland population may have

62

archaic to formative

moved periodically, perhaps seasonally, to the coast for a few days or weeks at a time, while the bulk of the population remained in the permanent inland communities. Remember that communities like Paso de la Amada, although classed here as inland, are relatively close to the ocean—Paso is located only 7 km from the sea. It would hardly be an onerous trip to the estuary, either on foot, or by canoe during times of high water. We might expect people to take agricultural foods and other products with them to the coast from inland villages and return from their sojourn with fish and game and perhaps plant products that are available only on the coast. The pattern in Table 3.3 of a relatively impoverished plant record at inland sites in the Mazatán zone, as compared with the wide range of plants revealed by the phytolith and pollen analyses at estuary sites, does not reflect the actual range of plants that would have been present and used at inland communities. This is certainly a function of the lack of microremain sampling at inland communities. If we refer to the record from the later Early Formative estuary site of Salinas la Blanca, we get a view of plant use similar to the one we get from the inland sites in the Mazatán zone. In other words, there is a strong representation of maize (the mineralized casts) and a range of at least three tree fruit species present. The tree fruits— avocado, matasano, and jocote— are all unlikely to preserve under most conditions. In the Mazatán zone avocados are preserved only when subjected to fi re and charred without burning up completely. Since avocados and other tree fruits would never likely be cooked over a fire, their preservation is highly unlikely in most circumstances. That any avocado pits have been preserved is a minor miracle. Dried maize, on the other hand, is likely to be cooked, and seeds could easily be exposed to fire as part of their processing. Furthermore, dried cobs can be used as fuel and burned in the fi re along with other fuels, thereby increasing their chances of accidental charring and preservation. So, although the paleoethnobotanical

analysis of the Early Formative sites in the Mazatán zone produced relatively few classes of plant material (six species, only three of which were economically significant—maize, beans, and avocados), the chances of two of these, maize and beans, being preserved through charring is actually quite high. Almost all of the other plant products used for foods, medicinal, or technological purposes would have little chance of being preserved. The one exception to this is, of course, wood or other plants used for cooking fuel— and there is abundant evidence of charcoal in most deposits throughout the region. We present these data as a way to systematically assess the evidence for plant use in general and the changing role of agricultural production in par tic u lar in the Soconusco during Archaic and Early Formative times. It is clear that we have evidence for diverse and complex human-plant interactions on the one hand but a spotty evidentiary record on the other. This, we believe, has allowed for a certain degree of wishful thinking to color our understandings of both the general and the specific question. Plants obviously were important in the ancient economy, just as they are today— this is not really at issue. What is at issue is the question of how we construct a model of ancient plant use that allows for a reasonable fit between our theoretical models (be they cultural ecology, behavioral ecology, material-historical, political economy, or other) and the physical remains that we have to work with. At present it is difficult to argue for the increasing importance of agriculture in the ancient economy because we do not have systematic evidence of presence or absence, relative quantities, range of types used, manner of growing, tending, harvesting, consumption, and changes in the nature of all of these through time. How important was the adoption, spread, and use of Theobroma cacao? Were other tree fruits besides cacao and avocados used on a regular basis? How did people ensure a supply of a wide range of plants that were used in technological activities (buildings, tools, utensils, arts, clothing, and so on)?

When did maize become a staple food, and how do we measure its changing dietary importance? What about root crops such as sweet potatoes and manioc? Questions such as these have been of primary concern to archaeobotanists and archaeologists studying the origins of tropical agriculture for decades, and they have been central to some of the most interesting and productive research in the Americas. Scholars such as the geographer Carl O. Sauer laid out the directions and premises for answering questions like these beginning in the 1950s (Sauer 1952). As Dolores Piperno and Deborah Pearsall (1998) have pointed out, to answer these questions and many more, we need to make use of all the data sets we have at our disposal because each can provide a separate window onto the issue at hand. In the coming years we will turn increasingly to a suite of analyses beyond macrobotanical and microbotanical collections and identifications, important though these may be. Increasingly we will be able to test our hypotheses about the changing nature of humanplant interactions with other methods as well, including molecular analysis, residue analyses, stable isotope measures of human remains as well as plant and animal specimens, and more precise and accurate dating techniques. We will also need a good deal of luck: somewhere out there is a perfectly preserved midden that contains just the right mix of archaeological debris that can answer all our questions— but we will not hold our breath. NOTES 1. This actually happened one day when Blake, along with John Clark first visited the estuary site of Cerro de las Conchas. 2. Numerous sources provide detailed lists of the economic plant species that occur in the region. Not all of these are necessarily native to the Soconusco, and in the case of some species it is difficult to determine if plants were introduced in preconquest times or during the colonial period. This is an area of research that needs a great deal more effort.

evidence for the diversity of plant use

63

3. A note on dating: we follow the broad temporal divisions outlined by Voorhies (2004) and Clark (personal communication 2007). However, these two sequences now leave a 400-year overlap between the end of the Late Archaic period at 1500 cal b.c. and Clark’s new dating of the Barra and Locona phases between 1900 cal b.c. and 1500 cal b.c. Since the Barra and Locona phases are calculated using many dates, all in association with ceramics, their dating provides a more solid framework than the relatively few Late Archaic dates at the end of that period, which are not associated with any diagnostic artifacts (indicating that they could be associated with Barra phase occupation). Therefore, for the purposes of this paper, we will assume that the Late Archaic Chantuto B phase ends about 1900 cal b.c. with the beginning of the Early Formative Barra phase. 4. In this chapter we restrict our observations to the microbotanical and macrobotanical remains that have been recovered and analyzed so far. Summaries of the other available lines of evidence (for example, plant processing and cooking technologies, chemical residue analyses, stable isotope analyses of human and other remains, and starch grain analysis) will have to wait for another paper. 5. Phytolith and pollen grain counts and percentages, as well as the methodologies for identifying the sample specimens, are detailed in the primary reports consulted for this chapter. Rather than reproduce these data here, we recommend the reader examine the original publications. The methods are generally standardized, though there are slight variations from study to study. For both generalized and detailed descriptions of the procedures used to collect, analyze, and report these types of plant microremains, see Pearsall (2000) and Piperno (2006). 6. Vicki Feddema’s M.A. thesis research was carried out in the early 1990s and examined the plant remains recovered in flotation samples from four sites in the Mazatán region excavated by John Clark and Michael Blake between 1985 and 1990. The methods, sample descriptions, and detailed analysis results are presented in the thesis (Feddema 1993). 7. One Jocotal sample that contained no identifiable macrobotanical remains is omitted. 8. These are new calibrations using OxCal 4.0 with IntCal 04 and reporting the probability ranges at 95.4 percent (Bronk Ramsey 2009). 9. Voorhies includes the Cuadros and Jocotal in the Middle Formative, but we leave them in the Early Formative period. 10. However, in this case the stable nitrogen measurement suggests that it was likely plant-based,

64

archaic to formative

rather than aquatic-based, protein that contributed to the diet.

REFERENCES Acuña, René 1982 Relaciones Geográficas del Siglo XVI: Guatemala. Universidad Nacional Autónoma de México, Mexico City. Aylor, Donald E., Baltazar M. Baltazar, and John B. Schoper 2005 Some physical properties of teosinte (Zea mays Subsp. parviglumis) pollen. Journal of Experimental Botany 56(419): 2401–2407. Blake, Michael 1991 An Emerging Formative Chiefdom at Paso de La Amada, Chiapas, Mexico. In The Formation of Complex Society in Southeastern Mesoamerica, edited by W. L. Fowler, pp. 27– 46. CRC Press, Boca Raton, FL. Blake, Michael, Brian S. Chisholm, John E. Clark, Barbara Voorhies, and Michael W. Love 1992 Prehistoric subsistence in the Soconusco region. Current Anthropology 33(1): 83– 94. Blake, Michael, John E. Clark, Barbara Voorhies, George Michaels, Michael W. Love, Mary E. Pye, Arthur A. Demarest, and Barbara Arroyo 1995 Radiocarbon chronology for the Late Archaic and Formative periods on the Pacific coast of southeastern Mesoamerica. Ancient Mesoamerica 6: 161– 183. Breedlove, Dennis E. 1981 Flora of Chiapas, Part 1: Introduction to the Flora of Chiapas. California Academy of Sciences, San Francisco. Bronk Ramsey, Christopher 2009 Bayesian analysis of radiocarbon dates. Radiocarbon 51(1): 337–360. Budowski, G. 1987 Living Fences in Tropical America, a Widespread Agroforestry Practice. In Agroforestry: Realities, Possibilities and Potentials, edited by H. L. Gholz, pp. 169– 178. Martinus Nijhoff, Dordrecht, Netherlands. Chisholm, Brian, and Michael Blake 2006 Isotope Analysis and Subsistence in the Soconusco Region, Mexico. In Histories of Maize: Multidisciplinary Approaches to the Prehistory, Biogeography, Domestication, and Evolution of Maize, edited by J. E. Staller, R. H. Tykot, and B. F. Benz, pp. 162– 172. Academic Press, Amsterdam.

Clark, John E. 1994 The Development of Early Formative Rank Societies in the Soconusco, Chiapas, Mexico. Unpublished Ph.D. dissertation, Department of Anthropology, University of Michigan, Ann Arbor. Clark, John E., and David Cheetham 2002 Mesoamerica’s Tribal Foundations. In The Archaeology of Tribal Societies, edited by W. A. Parkinson, pp. 278–339. Archaeological Series 15. International Monographs in Prehistory, Ann Arbor, MI. Coe, Michael D., and Kent V. Flannery 1967 Early Cultures and Human Ecology in South Coastal Guatemala. Smithsonian Contributions to Anthropology, vol. 3. Smithsonian Press, Washington, DC. Eccardi, F., and R. Álvarez del Toro 1987 Aspectos Generales de La Ecología en el Estado de Chiapas. Institutio de Historia Natural, Mexico, DF. Feddema, Vicki L. 1993 Early Formative Subsistence and Agriculture in Southeastern Mesoamerica. Unpublished M.A. thesis, Department of Anthropology, University of British Columbia, Vancouver. Available at https:// circle.ubc.ca/handle/2429/2254 . Ford, Anabel and Ronald Nigh 2009 Origins of the Maya forest garden: Maya resource management. Journal of Ethnobiology 29(2): 213–236. Helbig, Karl M. 1964 El Soconusco y Su Zona Cafetalera en Chiapas. Institutio de Ciencias y Artes de Chiapas, Tuxtla Gutierrez, Chiapas, Mexico. Henderson, John S., Rosemary A. Joyce, Gretchen R. Hall, W. Jeff rey Hurst, and Patrick E. McGovern 2007 Chemical and archaeological evidence for the earliest cacao beverages. Proceedings of the National Academy of Sciences USA 104(48): 18937– 18940. Hernández Bermejo, J. E., and J. León (eds.) 1994 Neglected Crops 1492 From a Different Perspective. Plant Production and Protection Series No. 26, Food and Agriculture Organization of the United Nations, Rome. Jones, John G., and Barbara Voorhies 2004 Human and Plant Interactions. In Coastal Collectors in the Holocene: The Chantuto People of Southwest Mexico, edited by B. Voorhies, pp. 300–343. University Press of Florida, Gainesville.

Kennett, Douglas J., Barbara Voorhies, and Dean Martorana 2006 An Ecological Model for the Origin of Maize-Based Food Production on the Pacific Coast of Southern Mexico. In Behavioral Ecology and the Transition to Agriculture, edited by D. J. Kennett and B. Winterhalder, pp. 103– 136. University of California Press, Berkeley. Lesure, Richard G. 1995 Paso de La Amada: Sociopolitical Dynamics in an Early Formative Community. Unpublished Ph.D. dissertation, Department of Anthropology, University of Michigan, Ann Arbor. Lesure, Richard G. (ed.) 2009 Settlement and Subsistence in Early Formative Soconusco: El Varal and the Problem of Inter-Site Assemblage Variation. Cotsen Institute of Archaeology Press, University of California, Los Angeles. Mangelsdorf, Paul C. 1967 Report on Mineralized Corncobs and Other Prehistoric Specimens From Salinas La Blanca. In Early Cultures and Human Ecology in South Coastal Guatemala, edited by M. D. Coe and K. V. Flannery, pp. 127– 128. Smithsonian Contributions to Anthropology, vol. 3. Smithsonian Press, Washington, DC. McNeil, Cameron (ed.) 2006 Chocolate in Mesoamerica: A Cultural History of Cacao. University of Florida Press, Gainesville. Miranda, Faustino 1976 La Vegetación de Chiapas. Segunda Parte, 2nd ed. Gobierno del Estado, Departamento de Prensa y Turismo, Tuxtla Gutierrez, Chiapas, Mexico. Neff, Hector, Deborah M. Pearsall, John G. Jones, Bárbara Arroyo, Shawn K. Collins, and Dorothy E. Freidel 2006a Early Maya adaptive patterns: Mid-Late Holocene paleoenvironmental evidence from Pacific Guatemala. Latin American Antiquity 17(3): 287–315. Neff, Hector, Deborah R. Pearsall, John G. Jones, Bárbara Arroyo, and Dorothy E. Freidel 2006b Climate change and population history in the Pacific lowlands of southern Mesoamerica. Quaternary Research 65: 390–400. Otero-Arnaiz, Adrians, Silvia Castillo, Jorge Weave, and Guillermo Ibarra-Manríquez 1999 Isolated pasture trees and the vegetation under their canopies in the Chiapas Coastal Plain, Mexico. Biotropica 31(2): 243–254.

evidence for the diversity of plant use

65

Paillés H., Maricruz 1980 Pampa El Pajón, an Early Middle Preclassic Estuarine Site, Chiapas, Mexico. Papers of the New World Archaeological Foundation, No. 44, Brigham Young University, Provo, UT. Pearsall, Deborah M. 2000 Paleoethnobotany: A Handbook of Procedures, 2nd ed. Academic Press, San Diego. Piperno, Dolores R. 1994 Phytolith and charcoal evidence for prehistoric slash-and-burn agriculture in the Darien rainforest of Panama. Holocene (4): 321–325. Piperno, Dolores R. 2006 Phytoliths: A Comprehensive Guide for Archaeologists and Paleoecologists. AltaMira Press, Walnut Creek, CA. Piperno, Dolores R., and Deborah M. Pearsall 1998 Origins of Agriculture in the Lowland Neotropics. Academic Press, San Diego. Popper, Virginia S., and Richard G. Lesure 2009 Macrobotanical Remains From El Varal, With a Comparison to Inland Sites. In Settlement and Subsistence in Early Formative Soconusco: El Varal and the Problem of Inter-Site Assemblage Variation, edited by R. G. Lesure, pp. 113– 114. Cotsen Institute of Archaeology Press, University of California, Los Angeles. Powis, Terry G., W. Jeff rey Hurst, María del Carmen Rodríguez, Ponciano Ortíz C., Michael

66

archaic to formative

Blake, David Cheetham, Michael D. Coe, and John G. Hodgson 2007 Oldest chocolate in the New World. Antiquity 314 (December). Available at www.antiquity.ac.uk/projgall/powis/index .html. Powis, Terry G., W. Jeff rey Hurst, María del Carmen Rodríguez, Ponciano Ortíz C., Michael Blake, David Cheetham, Michael D. Coe, and John G. Hodgson 2008 The origins of cacao use in Mesoamerica. Mexicon 30(2): 35–38. Rosenswig, Robert M. 2010 The Beginnings of Mesoamerican Civilization: Inter-Regional Interaction and the Olmec. Cambridge University Press, England. Sauer, Carl O. 1952 Agricultural Origins and Dispersals. American Geographical Society, New York. VanDerwarker, Amber M. 2006 Farming, Hunting, and Fishing in the Olmec World. University of Texas Press, Austin. Voorhies, Barbara 1976 The Chantuto People: An Archaic Period Society of the Chiapas Litoral, Mexico. Papers of the New World Archaeological Foundation No. 41, Brigham Young University, Provo, UT. Voorhies, Barbara 2004 Coastal Collectors in the Holocene: The Chantuto People of Southwest Mexico. University Press of Florida, Gainesville.

FOUR

Archaic to Formative in Soconusco the adaptive and organizational transformation Richard G. Lesure and Thomas A. Wake

R

ecent work in the early archaeology of the Soconusco leaves unresolved the nature of local adaptive transformations between the Archaic and Early Formative periods and the distinctiveness of those transformations within the larger picture of Mesoamerica. Three important elements of the local Early Formative adaptive reorientation—greater sedentism, the adoption of pottery, and the initiation of a trajectory of demographic expansion—were shared with regions that were environmentally very different, including the Gulf Coast, the Valley of Oaxaca, and the Basin of Mexico. Those similarities of pattern, together with the simultaneity of transitions to the Early Formative (1800 ± 100 b.c.) in diverse regions, would seem to indicate that explanation for the changes should be sought in general processes not tied to the details of local ecologies. Yet the status of a fourth element routinely linked to the other three— a step increase in the significance of agriculture—is less clear in the Soconusco than in the other cases. Available data on Early Formative subsistence in the

Soconusco do not add up in any single, coherent way. On the one hand, macrobotanical remains are dominated by the expected Mesoamerican domesticates from the Locona phase (1700 b.c. through 1500 b.c.)* at the latest (Feddema 1993). Sedimentary and pollen records of human impact on the environment in three different locations of the Guatemalan coast all register disturbance during the Early Formative. Since Archaic disturbances are by contrast uncoordinated and followed in one case by reforestation, a possible implication is for a transition from shifting, extensive land use in the Archaic to greater sedentism and more intensive cultivation in the Early Formative (Neff et al. 2006), much as traditional understandings of the Archaic-Formative transition would lead us to expect. On the other hand, stable isotope analyses of human bone from Early Formative sites yield carbon values inconsistent with a diet based on maize (Blake et al.

* See Figure 1.2 for site locations and Figure 1.3 for regional chronological chart. All dates are in calendar years.

67

1992a, 1992b; Chisholm and Blake 2006). Although it is worth remembering that the reliability of those data has been the subject of debate (Ambrose and Norr 1992; Chisholm, Blake, and Love 1993), it is relevant as well that settlement pattern changes between Late Archaic and Early Formative seem to have involved concerns other than agricultural intensification. In the reconstructed settlement system of the Archaic, residential bases were located on the interior coastal plain, on the best agricultural lands of the entire coastal strip (Kennett, Voorhies, and Martorana 2006; Michaels and Voorhies 1999; Voorhies 2004). Although the evidence for those interior residential bases is not overwhelming, we do not expect that it will be radically augmented in the near future and therefore assume here that the Archaic system reconstructed by Voorhies and colleagues is essentially correct. Although the Early Formative archaeology of the interior coastal plain is also not as developed as one might hope, the more permanent residential bases or villages of that era were not concentrated there but instead closer to the estuaries, on lands that would have been less desirable for agriculture (Kennett Voorhies, and Martorana 2006:134; Neff et al. 2006:308). Given that somewhat discordant evidence, investigators have tended to emphasize the importance of wild estuary resources to Early Formative adaptations in the Soconusco. They have simultaneously downplayed the significance of agriculture, suggesting, for instance, that maize was at that point cultivated primarily for beverages like beer rather than as a dietary staple (Clark and Blake 1994; Smalley and Blake 2003). Permanent villages, pottery, and rapid demographic expansion would have characterized local Early Formative adaptations without any significant increase in the dietary importance of cultivated plants. The latter might have increased gradually in importance from the later Archaic through Late Formative (Kennett et al. 2006), or a shift toward maize as a staple might have occurred at some point within the Formative—in par ticular at the tran-

68

archaic to formative

sition from the Early to Middle Formative around 1000 cal b.c. (Blake et al. 1992a, 1992b; Clark, Pye, and Gosser 2007; Love 1999; Rosenswig 2006). Is it reasonable to portray the local Early Formative as sharing some aspects of contemporaneous adaptive shifts (sedentism, pottery, population growth) with diverse other regions in Mesoamerica but not a reorientation toward agriculture? How distinctive were local adaptations? Could the Soconusco Early Formative, after a suggestion by Blake et al. (1992a:145– 146), even represent a transitory shift away from agriculture? Beyond any debate over the relative importance of different foods, how was the Early Formative subsistence system organized? Did the social organization of food procurement, processing, and consumption then differ from that of the Late Archaic, and, if so, how? Neff et al. (2006) provide a possible solution to the puzzle of simultaneous but locally distinctive Early Formative readaptations by identifying climatic drying as a causal factor in the early second millennium b.c. With simultaneity explained by climate, the inhabitants in each area can be imagined to have responded to subsistence stress in a manner appropriate to the local ecology. For the Soconusco, Neff et al. propose that innovations in technology allowed an expanded breadth of diet and intensified production of estuary foods; drawing on Blake et al. (1992a), they point specifically to net weights and fishhooks—in other words, to recovery technology. We find the basic scheme proposed by Neff et al. (2006) appealing, especially their emphasis on the Archaic-Early Formative shift as a demographically highly successful adaptive transformation founded on technological innovation. We are unconvinced, however, that innovations in recovery of aquatic animals could have kicked off Formative-scale population expansion. Recourse to climatic drying as an ultimate causal factor seems at first glance to allow readaptations across Mesoamerica that were different (in their articulation with local ecological conditions) and yet simultaneous. But

the pervasive success of these readaptations— the initiation in Soconusco, Oaxaca, Central Mexico, and the southern Gulf Coast of trajectories of demographic expansion—is enough to give pause. If Early Formative readaptations were locally specific responses to climatic stress and each therefore different one from another, one would have expected their demographic consequences to have been equally diverse. We are inclined, instead, to take a careful look at shared elements of Early Formative adaptations in any attempt to account for similarities in demographic consequences. Here we propose a refined understanding of Early Formative adaptive patterns in the Soconusco, identifying three familiar and widely shared elements as critical to the local readaptation: agricultural intensification, sedentism, and pottery. We suspect that there was an increase in the dietary significance of maize agriculture in the initial Early Formative, although we do not present new data or attempt to resolve the discordances already noted. As Clark, Pye, and Gosser (2007:37) suggest, maize may have “made sufficient difference in a mix of exploited resources to be a significant factor in people deciding to settle down.” Perhaps the Archaic-Formative transition corresponds to the point at which the focus of human use of maize shifted from stalk sugar to kernels (Webster 2011; see Lesure 2011). Our evidence pertains to the role of pottery in the Early Formative adaptive shift. In the Soconusco, we suggest, pottery proved an important technological innovation for local reasons. One vessel form in particular, the rounded-bottom, neckless jar (tecomate), allowed direct heating of contents. It was this innovation in processing technology (rather than changes in recovery technology) that, together with sedentism and agricultural intensification, fueled demographic expansion in the Early Formative. Those suggestions, offered here essentially as a hypothesis, stem from recently completed analyses for a monograph on El Varal, a specialpurpose, Early Formative estuary site in the Mazatán region (Lesure 2009b). Although exca-

vated materials from El Varal date to the end of the Early Formative (1250–1000 b.c.), we propose that the site fit into an adaptive pattern that first appeared in the Locona phase and persisted into the Middle Formative. Special-purpose estuary sites are also known from the Archaic, but they differ in character from those of the Early Formative. The altered contents and structure of such special-purpose sites are significant clues to local adaptive and organizational changes at the threshold of the Formative, but it is necessary to grapple with a tendency to formulate the “special” purposes of estuary sites on either side of the Archaic-Formative divide by reference to divergent theoretical formulations. The hunting and gathering system of the Archaic has been understood (after Binford 1983) in terms of logistically organized collecting strategies, while the Early Formative cases have been formulated as instances of occupational specialization in societies increasing rapidly in complexity. Comparisons between El Varal and nearby sites provide an empirical basis for exploring the appropriateness of those theoretical formulations for understanding Early Formative adaptive organization. Our results complement previous suggestions by Arnold (1999) for the Tuxtla Mountains and Arroyo (2004) for Pacific coastal Guatemala envisioning significant mobility in Early Formative settlement systems.

EL VARAL, MATERIALS RECOVERED, AND CASES FOR COMPARISON El Varal is located in the Mazatán region of the Soconusco near of the mouth of the Coatán River, 2 km inland from the ocean. It is today in a seasonally flooded zone of the Pampa Cabildo estuary near a salt flat (Clark 1994: figure 9) that likely marks the location of an ancient lagoon (after Coe and Flannery 1967:12). Stratigraphic evidence suggests that the site was an island in that lagoon during the Cuadros phase, with siltation proceeding rapidly so that by late Jocotal the site was in a seasonally inundated but vegetated pampa (see Lesure 2009a).

adaptive, or gan i za tion al transformation

69

FIGURE 4.1 View from the canal cut toward the west profile of the Vásquez Mound, El Varal, with excavations in progress, 1992. Note the artificial steps cut to record the profile. Visible atop the mound is the huge heap of dirt removed by the bulldozer. Photo by Richard Lesure.

The site consists of two large mounds with a total surface of approximately 1 hectare. The Martínez Mound is 80 × 50 m and 7 m high. Immediately beside it is the Vásquez Mound, 5 m high and vaguely round, with a diameter of 90 to 100 m. The 1992 investigations were confined to the latter mound, an entirely human construction built up gradually during the Cuadros and Jocotal phases. A surface collection from the Martínez Mound indicates that it dates to the Locona phase. Sometime during the fall of 1991, persons unknown dug a drainage canal through the Vásquez Mound (Figure 4.1). The cut was 30 m wide in its maximum extent at ground surface toward the middle of the mound, with exposed archaeological deposits sloping down from those upper edges to the water. The canal alone was about 11 m wide. Archaeological fieldwork was a salvage project arranged by John Clark under the auspices of the Mazatán Early Formative Project, and it was conducted by Lesure and Tomás Pérez Suárez. Investigations were confined to the two 100-m-long profiles—5 m high in the middle of the mound— exposed by the bulldozer. The field crew cleaned and recorded the west profile by means of artificial steps (see Figure 4.1); selec-

70

archaic to formative

tively tested middens exposed in the west profile; and, finally, excavated a 30-m-long segment of one of the steps created during profile cleaning to obtain a continuous sequence of refuse deposits from the earliest to latest occupations (the Step Excavation). Given the extensive deposits exposed by the bulldozer, it is noteworthy that only a single possible formal structure was identified. The bulk of exposed deposits consisted of slanting layers of sediment and artifacts. They derive from materials dumped by Early Formative occupants off the surface of the mound. The inclination of these layers ranged up to a high of 30°. It is due to this curious internal structure that a continuous stratigraphic sequence was obtained in the Step Excavation by digging horizontally along a single step created in the profi le cleaning. The stratigraphy of the dump zone tended toward an alternation between layers of yellow-brown silty sediment with few artifacts and layers containing mainly broken sherds and chunks of fire-reddened earth. We believe that the sediment layers are the debris of salt production. The excavated deposits can be divided into four Stratigraphic Periods. The Early Period

(late Cuadros or early Jocotal) corresponds to informal surfaces and sandy outwash deposited before the appearance of thick refuse dumps. The Middle Period (early Jocotal) includes inner refuse-dump layers as well as associated surfaces and outwash zones. The Late Period (late Jocotal) includes the outer refuse dumps. The Terminal Period (end of Jocotal and transitioning to Middle Formative) is distinguished by dramatic changes in artifact assemblages noted at the north and south extremities of the western profile. It appears to correspond to a relatively brief episode just before the site was abandoned, when Varal became a typical habitation site rather than a special-purpose location for the production of estuary resources. Broken pottery was the most abundant artifact. Sherds were large and, in the Early-Late Periods, overwhelmingly from a single vessel form: large, thick-walled, rounded-bottom tecomates. Mollusk shells were also common, particularly in the Late Period. Other items recovered include obsidian flakes, grinding stones, shell artifacts, figurines and other ceramic artifacts, crab claws, and animal bones (predominantly from fish). Except for a rise in mollusk shells in the Late Period, artifact occurrences were more or less stable from Early through Late. Changes in the Terminal Period include a switch from a pottery assemblage dominated by tecomates to one dominated by dishes as well as dramatic rises in the frequency of obsidian flakes, figurines, worked sherds, and grinding stones. Numerous individuals have contributed to analyses. Here, we draw mainly on Lesure’s work with stratigraphy and site contents and Wake’s identifications of the archaeofauna. Our data on crabs are from Dietler and Wake (2009); our data on birds are from Steadman, Tellkamp, and Wake (2003), as updated in Lesure, Wake, and Steadman (2009). We will make brief reference to relevant analyses of other materials (described in detail in the site monograph, Lesure 2009b). Particularly important for the analysis of faunal remains are comparable materials from habitation sites. Unfortunately, no

faunal assemblage from an exact contemporary of El Varal is available. We make use of two Cherla-phase (1400– 1300 b.c.) samples from other sites in the Mazatán region, one from Aquiles Serdán reported by Blake et al. (1992a) and the other from Paso de la Amada (three units of lot 11, Mound 1), analyzed by Wake (2004). Detailed faunal comparisons between sites of the sort described here have to our knowledge never before been conducted on materials from the Early Formative Soconusco. The choice of samples, while not ideal, is based on what is currently available in analyzed form. We justify the use of materials from different phases by claiming basic adaptive stability throughout the Early Formative, a claim that is supported by preliminary results of an expanded analysis of remains from Paso de la Amada, currently in progress.

ARCHAIC SPECIAL-PURPOSE SITES: PATTERNS AND INTERPRETATION Archaic special-purpose estuary sites are large mounds composed almost entirely of shells. Documented at Cerro de las Conchas adjacent to the Hueyate Swamp and numerous sites in the Acapetahua Estuary by Voorhies and colleagues (Kennett, Voorhies, and Martorana 2006; Michaels and Voorhies 1999; Voorhies 1976, 2004; Voorhies et al. 2002), the contents are in most cases marsh clams of a single species (Polymesoda radiata). Voorhies (2004:42– 51) notes patterns in the bedding of the shells involving paired layers of burnt and unburnt shells. The lack of disturbance in these extensive layers suggests that people collected large numbers of shellfish during short visits. Absence of structures and a limited range of tool types further support the idea that these were not long-term habitation sites. Drawing on Binford’s (1983) analysis of hunter-gatherer settlement systems, Voorhies (2004; see also Michaels and Voorhies 1999) interprets the Archaic shell mounds as specialpurpose locations for the extraction of estuary resources. They are the most archaeologically

adaptive, or gan i za tion al transformation

71

visible manifestation of a logistically organized settlement system centered at residential bases on the interior coastal plain. Task groups dedicated to a narrow range of goals would have visited the sites for short periods. They collected large numbers of marsh clams, steamed them open, and dried the meat. The product would then have been transported inland for consumption by some larger social unit. Voorhies (2004:400–403) identifies other possible products as well, including shrimp and fish, although the argument for those as products processed for transfer inland is hypothetical. This adaptive pattern was resilient over several thousand years (although see Chapter 2, in which changes toward the end of the Archaic are discussed). By the Locona phase, however, the shell mounds had been abandoned and a new pattern of estuary use appeared. Some sherds of the Barra phase mixed into ceramic levels at Tlacuachero and the absence of markers from the rest of the Early Formative raise the intriguing possibility that the Archaic adaptations in the Acapetahua Estuary continued into the incipient Early Formative (Lowe 1978:346; see Voorhies 1976: figure 57). Such a pattern would make sense from the standpoint of the argument developed here, since rounded-bottom tecomates appear only with the Locona phase.

EARLY FORMATIVE SPECIAL-PURPOSE SITES: PATTERNS In the Early Formative case we have the advantage of good evidence from both special-purpose and long-term habitation sites, although the contemporaneous presence of two types of site is well documented only in the Mazatán area. There, the basic pattern is as follows. Long-term habitation sites are located somewhat inland from the estuary and contain evidence of architecture. A variety of domestic artifacts is found, with pottery assemblages dominated by dishes, which constitute usually 50 percent or more of the assemblage from the later Locona phase. Special-purpose sites are small in areal extent but highly mounded; they were islands

72

archaic to formative

in the estuary. Three such sites investigated in the Mazatán region—Los Alvarez, Sandoval, and El Varal— are close to each other in the Pampa Cabildo near the mouth of the Coatán (see Chapter 10, Figure 10.2). Mollusk shells are present but come nowhere near forming the bulk of the deposits. There is, instead, considerable evidence of burning in the form of chunks of fire-reddened earth and much pottery. The vessel assemblages are tecomatedominant, with 60 percent to 90 percent being tecomates. Domestic assemblages are impoverished, and evidence of architecture is rare. The dish-dominant and tecomate-dominant distinction characterized contemporaneous sites a few kilometers from each other. For example, in the Locona phase we can point to dishdominant Paso de la Amada and tecomatedominant Los Alvarez; for Cuadros-Jocotal we  have dish- dominant Cantón Carralito and tecomate-dominant El Varal. The pattern of intersite assemblage differences thus spans most of the Early Formative from 1700 b.c. through 1000 b.c., suggesting certain stability in the activities generated. Certain clarifications and complications require discussion. Relevant data are assembled in Figure 4.2, which charts tecomates as a percentage of Pacific coast vessel assemblages from the Locona phase through the Middle Formative. Tecomate-dominant assemblages appear in the upper ellipse and dish-dominant assemblages in the lower ellipse. The most obvious point to be made is that the distinction between special-purpose assemblages and those from permanent communities is complicated by a steady reduction, over the Early Formative, in the percentage of tecomates at the permanent communities and perhaps in the specialpurpose sites as well. At any one time there were estuary sites with discard assemblages containing high proportions of tecomates and inland sites with lower proportions. However, tecomates in general became a less important component of Formative vessel assemblages, probably as people used pots for more diverse

100

Seasonal and/or specialpurpose occupations

90

Tecomates as a Percentage of Assemblage

Paso de la Amada

80

Chilo Canton Corralito

70

La Calentura La Blanca

60

El Varal Sandoval

50

Cuauhtemoc Chiquiuitan

40

El Mesak (minus Mesak jars)

30 20

Permanent communities 10 0 –1800

–1600

–1400

–1200

–1000

–800

–600

–400

–200

0

Years B.C.

FIGURE 4.2 Tecomates as a percentage of vessel assemblages (based on rim sherds) from ten Early Middle Formative sites along the Pacific coast of Chiapas and Guatemala. Data from the following sources: Clark 1994; Pye, Hodgson, and Clark, Chapter 10, this book; Lesure 1995; Lesure and López 2009; Love 2002; Morgan, Chapter 9, this book; Rosenswig 2006.

tasks and, eventually, as jars (with necks) began to replace tecomates.1 A second point of clarification concerns the high percentages of tecomates (55 percent to 70 percent) observed in the earlier Locona phase (c. 1700– 1600 b.c.) at permanent, inland communities in the Mazatán area (Cantón Corralito [San Carlos Mound], Paso de la Amada, Chilo, La Calentura). Such assemblages are dish- dominant in the sense that there were other, contemporaneous sites with 90 percent to 95 percent tecomates; no simple labels for the phenomena in question are entirely satisfactory, as we will see for the alternatives of “estuary” and “inland.” Inclusion of the Barra phase (1900– 1700 b.c.)—left out of Figure 4.2  because it predates the emergence of the special-purpose estuary sites—would have made the pattern of high proportions of tecomates at early inland sites even more dramatic. Clark and Gosser (1995) interpret the Barra tecomates as vessels for serving and consum-

ing beverages. In contrast, tecomates at estuary, special-purpose sites are round-bottomed and unslipped below an exterior band. We interpret them as being preparation rather than service vessels. Many were used for the heating of contents over direct heat. Thus, two phenomena in Early Formative Soconusco involve high frequencies of tecomates. The first is early and ephemeral: the Barra and Early Locona use of smaller, often flat-bottomed tecomates as a serving vessels. The second is more persistent: special-purpose estuary sites with assemblages dominated by larger, round-bottomed, utilitarian tecomates. Our concern here is with the second phenomenon. A third point is that it was possible for a site to shift from being tecomate-dominant to dishdominant. Such a shift is best documented at El Varal, as already described here. In Figure 4.2, the four data points from El Varal correspond to the Early, Middle, Late, and Terminal assemblages. The Terminal Period shift realigned the

adaptive, or gan i za tion al transformation

73

Varal assemblage with that of contemporaneous inland villages. It is because of such shifts at individual sites that we have chosen to supplement the estuary-inland distinction (see Chapter 10 in this book) with the categories of dish-dominant and tecomate-dominant. Finally, although the pattern of Early Formative intersite assemblage differences has been known in the Mazatán area for several decades (Lowe 1977:210–211) and estuary sites with tecomate-dominant assemblages have been documented in other areas of the Pacific coast (Pajón [Pailles 1980] and Salinas la Blanca [Coe and Flannery 1967]), evidence of the contemporaneity of dish-dominant and tecomate-dominant assemblages within a single settlement system has proved elusive outside of the Mazatán area. The pervasiveness outside of the Mazatán area of the organizational patterns we describe therefore remains a topic for further work. Because of the strong diachronic component to tecomate percentages just at permanent villages and the possibility of changes of function of particular sites (Figure 4.2), classifying a site as specialpurpose or permanent is not a trivial matter, even if we assume the same basic adaptive patterns throughout the Soconusco. Figure 4.2 includes data from four sites outside the Mazatán area: La Blanca, Chiquiuitán, Cuauhtemoc, and El Mesak (see Chapter 1, Figures 1.1 and 1.2 for locations). The Middle Formative assemblage from the large, inland community of La Blanca (see also Chapter 8 herein) extends the trend toward reduction in tecomates seen among the Mazatán sites. Further, it appears that there were special-purpose estuary sites contemporary with La Blanca in the Naranjo region (Shook and Hatch 1979:147–148), suggesting that the organizational patterns discussed here persisted into the initial Middle Formative at least in that area. The data reported by Morgan for the more distant Chiquiuitán (Chapter 9) fit the patterns from Mazatán if we interpret the Huiscoyol-phase occupation as special-purpose and occupation in the Cangrejo and Tamarindo faces as permanent; however, no evidence of an inland village contemporary to the

74

archaic to formative

Huiscoyol occupation has been forthcoming. The only real divergence from the basic patterns in Figure 4.2 comes from the early deposits at Cuauhtémoc, much closer to the Mazatán area (Rosenswig 2006; see also Chapter 11 in this book). The low percentages of tecomates in those deposits remind us that we still have more to learn about variability in subsistence and social practices in Early Formative Soconusco. The fourth site outside of the Mazatán area, El Mesak, merits lengthier discussion (see also Chapter 10). The Jocotal occupation of that site, contemporaneous with the occupation of the El Varal, constitutes an altogether different pattern of special-purpose estuary use. The later Early Formative assemblage at El Mesak is dominated (75 percent to 90 percent) by excessively crude, deep, open-profile, vaguely conical jars (see Chapter 10, Figure 10.5). Pye (1995) believes those vessels were used in the production of salt by part-time specialists, and we find her arguments convincing. Extrapolating from the El Mesak case, we suggest that this second sort of special-purpose estuary site differs from the first in both spatial and temporal distribution: it is rare, and the patterns involved emerged progressively over the course of the Early Formative rather than rapidly near the beginning of the period. It also appears to be associated with permanent rather than part-time or seasonal occupation. To produce our Mesak data point for Figure 4.2, we have left out the special-purpose “Mesak jars” from Pye’s data in Table 10.4 (Chapter 10 herein) and included only tecomates and other vessels; the resulting “domestic” assemblage aligns clearly with regular Jocotal settlements. In our conclusions, we will attempt to make sense of this contrast between two sorts of special-purpose sites.

EARLY FORMATIVE, TECOMATE-DOMINANT SITES: INTERPRETATION Most observers have identified tecomatedominant sites as special-purpose locations for the production of estuary resources, and we

concur (for further discussion, see Lesure 2009c). Two questions follow. First, what resource or resources were produced? Second, how was production organized such that a divide between inland, dish- dominant and estuary, tecomate- dominant assemblages was its archaeological consequence? There has been considerable unanimity on the range of possible products but not much progress in identifying what was actually produced. Unless tecomates themselves were the products (Carballo et al. 2009; Ceja Tenorio1999), the extensive evidence for fire, the appropriateness of rounded-bottom tecomates as cooking pots, and patterns of burning on tecomate sherds prompt attention to what might have been cooked in these vessels (Blake et al. 1992a; Ceja Tenorio 1999; Clark 1991, 1994; Clark and Pye 2000; Lesure 1993; Shook and Hatch 1979; Smith 1997). Foods could have been cooked for consumption on-site. Alternatively, some product could have been processed and transferred to inland villages—perhaps salt, shrimp, or some sort of only vaguely defined fish product (“fish cheese” [Pailles 1980] or something similar to gefilte fish [Smith 1997]). If fish were prepared for storage, salting and drying would seem much more likely than boiling it, both in terms of its storability and its relation to practices with contemporary precedents on the Chiapas Coast (see Lindner 1944). Overall, despite numerous suggestions, there has been little detailed work attempting to weigh one possibility against another (although see Carballo 2001 and Smith 1997). While Archaic special-purpose sites are interpreted from the perspective of variability in hunter-gatherer settlement-subsistence systems, their Early Formative counterparts have most often been viewed through the lens of occupational specialization and societal complexity (Clark 1991, 1994; Clark and Pye 2000; Lowe 1977, 1978; Shook and Hatch 1979). That trend is certainly understandable given the evidence for hierarchy at sites like Paso de la Amada (Blake et al. 2006; Clark and Blake 1994), but inequality need not equate with social

heterogeneity and, indeed, may in many cases have emerged earlier (McGuire 1983). When we further consider that most proposed products are wild subsistence resources, a framework based on hunter-gatherer settlement systems would seem an equally viable route toward interpretation. Lesure (2009a) identifies four possible models for the organizational patterns that generated observed intersite assemblage differences. From greater to lesser overall complexity, the four are set forth as follows.

Part-Time Occupational Specialization With Year-Round Residence Lowe (1977:210) and Shook and Hatch (1979:147– 148) suggested that tecomate- dominant assemblages were produced by part-time specialists in residence year-round. The sites would thus have housed permanent, settled communities. Households would have produced beyond their own needs, and goods would have been transferred to consumers by some institutionalized mechanism, whether exchange relations or a tribute system. In this scenario, we would expect to find a domestic occupation with the same categories of features and artifacts, in comparable frequencies, that appear at regular habitation sites. As we have already intimated, El Varal does not fit this model; still, since most previous observers incline toward something along these lines, it seems helpful to explicitly rule it out.

Part-Time Occupational Specialization With Seasonal Residence A second possibility is that the occupants of El Varal were specialist producers, but they resided at the site only part of the year. In this scheme, we would expect the same sorts of relations of producers to consumers (exchange or tribute) but a different archaeological record. Seasonal residence would imply that the occupants of tecomate- dominant sites maintained other dwellings, elsewhere. At the location of production, we would expect domestic assemblages

adaptive, or gan i za tion al transformation

75

with reduced representation of artifacts and features not involved in the productive activity.

Logistical Collecting at Estuary Field Camps Binford’s (1983) collecting model applies to those hunting and gathering systems based on a general strategy of moving resources to people. Collectors establish relatively stable residential bases. Specially organized task groups leave those bases, sometimes for significant periods, to obtain specific resources. These task groups are “small and composed of skilled and knowledgeable individuals. They are not groups out ’searching’ for any resource encountered” (Binford 1983:344). One type of site generated in such a system is the field camp, a temporary center of operations for a task group. The group (a small subset of the entire community) sleeps and eats there while engaged in collecting and processing some specific resource. This third model is similar in many respects to that of occupational specialization with seasonal residence. In both cases, production sites would have been occupied for part of the year, by people who were producing for others and whose main residential location was elsewhere. There are important differences, however. In the collector model, producers owe food to others by virtue of their shared group membership. Further, the identity of “producer” is fluid: task groups are continually constituted and reconstituted. In the specialization model, producer-consumer relations are economic transactions between basic units of production and consumption (such as households) rather than relations of obligation between producers and members of a larger consumption unit. There is no simple way to distinguish the two models archaeologically, and we will do our best to build a case based on as many different sources of evidence as possible.

Generalized Foraging Binford (1983) contrasts collecting with foraging, in which entire groups shift frequently between residential bases. From each base, foragers move

76

archaic to formative

out daily to collect various resources depending on what they happen to encounter. Since one of the main archaeological consequences of a system like this is redundancy of structure and contents between sites, the generalized foraging model is of little direct help in explaining our inter-site assemblage differences. Still, this proves a useful model to have at hand. Two points of contrast with the collecting model are of particular interest. A generalized strategy in which visitors to field camps collected any of various resources they might encounter would constitute a notable departure from the strategy of collecting, as would anything approximating the basic foraging strategy of moving people to resources rather than the reverse.

RESULTS Our review of evidence from El Varal (along with comparable data from Archaic shell mounds and Early Formative dish- dominant sites) centers on three topics. We begin by rejecting the model of full-year resident specialists. We then review the applicability of the remaining organizational models to the production of those animal foods for which we have good direct evidence: mollusks, crabs, and vertebrates. Finally, we consider the possibility of specialization in other products, focusing particularly on salt. These topics provide the strands we weave together in two sets of conclusions— first, concerning the organizational arrangements that produced a division between tecomate- dominant and dish- dominant assemblages and, second, concerning ArchaicFormative adaptive and organizational transformations more generally.

The Case Against Full-Year Residence The case against full-year residence at El Varal from Early through Late Periods— and for such a pattern in the Terminal Period—rests on three broad sources of evidence. First, there is the stratigraphy exposed in the canal profi les and the distribution of features therein. From

the Early through the Late Period, there is only one case of a possible formal structure, and pits are rare and smaller than those observed at inland sites. Small residential platforms are Terminal in date, and burials are all either Late or Terminal. Thus, from Early through Late, the low level of investment in long-term facilities matches that expected of a field camp. Second, a study of seasonality in the collection of one mollusk species (Polymesoda radiata) by Kennett and Culleton (2009), modeled on studies of the same species in Archaic shell mounds, indicates dry season occupation (see Chapter 2 in this book for a brief description of the logic involved). Shells sampled from Early (N = 10) and Late (N = 8) Periods were all harvested in the dry season. The Middle Period sample was more mixed but still predominantly dry season (14 of 21 shells). The final source of evidence is the artifacts recovered. Lesure (2009a, 2009d) has considered both categories of artifacts and relative frequencies in comparisons between Early through Late versus Terminal Periods at El Varal and between Terminal Period El Varal and a sample of Cuadros-Jocotal domestic refuse from inland, dish-dominant Cantón Carralito (San Carlos Mound area). In all three cases, much the same range of artifact categories is present. Artifact frequency presents another story, however. While Terminal Period refuse compares very favorably with the sample from Cantón Carralito, the Varal Early-Late assemblage is impoverished in domestic vessel forms but also in obsidian, figurines, grinding stones, and worked sherds. Patterns hold either for standardization by deposit volume or by weight of associated sherds. Could the depressed frequencies of domestic artifacts be the result of examining an assemblage containing a mixture of “normal” domestic debris and broken tecomates deriving from specialized production? Even an elaborate analysis attempting to factor out such extradomestic debris fails to eliminate the depressed frequencies of domestic artifacts in Early-Late deposits (Lesure 2009a).

While an argument for part-year, predominantly dry season use of the site by people who maintained residences elsewhere seems well founded, the scale of activities at El Varal should not be underestimated. Lesure (2009a) calculates that the Vásquez Mound contained approximately 30,000 broken dishes and 116,000 broken tecomates, all deposited within 200 to 250 years. Even though several parameters can only be speculatively estimated, calculations using Schiffer’s (1975) discard equation (see also Varien and Mills 1997) suggest we should think in terms of yearly visits by a few hundred people. The presence of all basic categories of domestic artifact indicates that occupants were not singlemindedly devoted to whatever productive tasks brought them to the site, probably because occupations lasted for weeks and groups themselves were large and diverse.

Subsistence Remains: Mollusks, Crabs, and Vertebrates The problem of subsistence at El Varal includes the question of what people at the site ate as well as the possibility that those people produced food for others. We will take as a null hypothesis the idea that, at both dish-dominant and tecomatedominant sites, animals caught by occupants were processed and eaten on-site. By comparing the Varal remains (Early-Late Periods only) to those from Early Formative dish-dominant sites and to Archaic shell mounds, we look for any evidence that foods were instead removed from El Varal for consumption at inland sites. We are thus attempting to distinguish part-year residence by specialists or collectors (the second and third organizational models, indistinguishable based on the data under consideration here) from a null hypothesis positing lack of any specialized production or logistical collecting of animal foods. The overall amounts of faunal remains recovered from El Varal generally favor the null hypothesis. Conceivably, occupants could have boiled their catch, discarded the bones, and transferred the flesh inland; in such circumstances

adaptive, or gan i za tion al transformation

77

we would expect considerable accumulation of bones at El Varal. The Step Excavation yielded a total vertebrate NISP (number of identified specimens) of 1873 with a total MNI (minimum number of individuals) of 293, associated with a fullvessel equivalent of 93.3 tecomates (based on summing rim proportions). Either the occupants were implausibly inept with their pots, or they mainly cooked something else in tecomates. Another way of looking at the same issue is to note that vertebrate remains at an Archaic shell mound such as Cerro de las Conchas are probably the remains of meals of site occupants while they worked processing clams. The density of vertebrate remains at Cerro de las Conchas ranges from 200 through 1,300 identified specimens/m3 in three strata. The range for El Varal when individual lots are lumped by period is 100 to 300 specimens/m3. The density in our single analyzed sample from Paso de la Amada was approximately 1,000 specimens/m3 (recovered in the same 5 mm screens used at Varal). Still, that sort of density is clearly at the high end of the inland sites’ range, and there was variation lot-to-lot in the Varal case, with a high of 1,080 specimens/m3. Although crab remains are more common at Varal than Paso, the numbers do not seem so great as to suggest large-scale processing of crab (total Varal crab MNI is 555). The same goes for mollusk shells in the Early and Middle Periods at Varal. In the Late Period there are several dense dumps containing almost pure shells, but we suspect that even those mollusks were consumed on-site. Basically, the frequencies of animal remains at El Varal seem comparable to those at inland sites. The relevant point at the moment is that frequencies are not any higher than expected, a result that favors our null hypothesis. In the next section, we refer to this same evidence to make the point that frequencies at Varal are no lower than expected. Perhaps the bones of vertebrate fauna moved with the food from tecomate- dominant to dish- dominant sites. In other words, maybe Varal was a “field camp,” while dish- dominant sites inland from the estuary were “residential

78

archaic to formative

bases.” Our expectations for the distribution of faunal remains generated by such a system are that residential bases should have a richer and more diverse array of fauna than field camps. The remains at a field camp would represent fauna harvested from habitats in the immediate vicinity of the site, whereas those at the residential bases should reflect the results of pooling resources from multiple habitats. Unless processing at the field camp involves selective removal of the archaeologically visible part of a par ticular resource (such as clamshells), we would generally expect the fauna represented at a field camp to be a subset of that identified at the residential base. We begin examination of that possibility by means of a three-way comparison between Varal, two Archaic shell mounds (Tlacuachero, Cerro de las Conchas), and two dish-dominant Early Formative sites (Paso de la Amada and Aquiles Serdán). The Archaic cases provide a helpful third point of comparison, since, following Voorhies (2004), we are positing that those were indeed field camps. One question is thus whether the Varal assemblage might resemble those of the Archaic sites more than it does its Early Formative but dish-dominant counterparts, although we are also interested in any patterns that might collectively characterize the Early Formative assemblages. Table 4.1 provides relative percentages of basic classes of fauna, whereas Table 4.2 presents for each of those classes the number of taxa identified at the genus level (or higher levels if those were assigned MNI values by the original investigators).2 The high incidence of mollusks at El Varal (Table 4.1) constitutes the one point of similarity between that site and the Archaic estuary sites (although it should be emphasized that absolute numbers were orders of magnitude less and El Varal was in no sense a “shell mound”). In most other respects, the Varal assemblage falls in comfortably with its Early Formative counterparts. Previous observers including Blake et al. (1992a), Kennett, Voorhies, and Martorana (2006), and Neff et al. (2006) note a greater breadth of diet in the Early Formative Soconusco

TABLE 4.1 Estimates of Relative Percentages of Fauna at Late Archaic Acapetahua Shell Mounds (Tlacuachero/Campón) and Three Early Formative Sites

Site

Tlacuachero- Campóna

Mollusk

Crab

Fish

Reptile or amphibian

Bird

Mammal

99.4

0

0.44

0.06

< 0.01

0.10

b

El Varal

All classes

89.5

Without mollusks

6.9

2.7

0.2

0.3

0.4

65.5

25.7

2.1

3.2

3.4

74.6

6.2

9.2

9.9

8.4

58.7

9.0

9.0

10.3

8.8

61.5

9.4

9.4

10.8

67.4

10.4

10.4

11.8

83.4

4.6

4.4

7.6

Vertebrates only c

Paso de la Amada All classes

4.5

Without mollusks Vertebrates only Aquiles Serdán

a

unknown

unknown

note: Percentages are based on summed MNIs where available, except for values from Tlacuachero- Campón, which are based on estimates of meat biomass. a

From Voorhies (2004: fi gure 3.1 and pp. 125, 147). Only the four most common mollusks are included in the counts. Note that this means excluding tens of thousands of tiny Amphichaena kindermanni shells concentrated almost entirely in two Late Period lots. Even so, the par tic u lar abundance of mollusks in the Late Period has a substantial effect on the percentages. c Mound 1, lot 11, three units only (F9, H8, and H9). This is the only available sample from the site for which all classes have been studied. d From Blake et al. (1992a: tables 1 and 2). Mollusk shells and crab claws are not reported there. In the report, MNIs are not provided for all of the identified taxa. We have assumed an MNI of 2 for “several” and 1 for “rare.” b

than in the local Archaic. That pattern can be noted in both tables here, at both dish-dominant and tecomate-dominant sites. Among the shellfish, the number of taxa identified at El Varal is higher than at the Late Archaic sites (Tlacuachero in Table 4.2) and the Middle Archaic Cerro de las Conchas. The shells of Paso de la Amada were in terrible shape because of soil conditions; still, a range of taxa is present despite the small sample (158 grams of shell). Crabs, all but absent in the Archaic, were present in some diversity at both Varal and Paso, and in the Early Formative generally there is a greater representation of reptiles, amphibians, birds, and mammals than in the Archaic cases. Still, the Early Formative assemblages are not more diverse than those of the Archaic in every sense of the term. The most interesting pattern is a greater evenness of spread among

the fish taxa at the Archaic sites (see Lesure, Wake, and Steadman 2009). Early Formative vertebrate assemblages are more diverse in the sense that more taxa are actually represented, but among the top fish taxa, just a few were collected in large numbers. That pattern holds for both dish- and tecomate-dominant sites and distinguishes them collectively from the Archaic shell mounds. In other words, Varal is not identifiable as a field station in a three-way comparison between Archaic field stations and Early Formative villages, perhaps because of larger shifts in procurement strategies between Archaic and Formative. The next question becomes whether we can tease out any differences between the Early Formative sites that might indicate that tecomatedominant sites were field stations. The pattern that we are looking for is greater diversity of taxa

adaptive, or gan i za tion al transformation

79

TABLE 4.2 Number of Distinct Taxa (at the Genus Level or Above) of Mollusks, Crabs, Fish, Reptiles or Amphibians, Birds, and Mammals at Archaic and Early Formative Sites

Site

Cerro de las Conchas

Mollusk

Crab

Fish

Reptile or amphibian

Bird

Mammal

Vert. N f

~10b

~1

21

3

1

3

199

1 + ~4b

0

12

2

0

2

123

27c

5

23

5

15

7

292

~ 10d

3

14

7

13

9

135

Unknown

Unknown

5

8

13

9

327+

a

Tlacuacheroa El Varal Paso de la Amada e

Aquiles Serdán

note: Focus on the genus level helps to even out some of the differences of reporting among investigators. Taxa above the genus level were included only if the original investigators assigned an MNI value. a From Voorhies (2004: tables 3.2 and 3.3; also p. 147) and Wake (2004: tables 4.1 and 4.8). That last table has formatting errors in Lesure’s copy of Voorhies 2004. He has inferred actual MNIs based on column totals. b From Tlacuachero, the better value for comparison is 1: virtually all of a vast number of shells were from a single taxon (Polymesoda radiata). A few other taxa were identified in minuscule numbers (Voorhies 2004: table 2.2). At Cerro de las Conchas, the lowest stratum was very similar in content to Tlacuachero, and the upper two strata included a greater diversity of shells. c There were significant shifts over time in the relative percentages of the primary four species, Polymesoda radiata, Anadara grandis, Chione subrugosa, and Protothaca metodon. d Mollusk remains from all areas of the site were in terrible shape, apparently because of soil conditions. In the Lot 11 sample, the most common shells represented are the same four that are most consistently found across the sequence at Varal, in the following order from most to least common: Chione subrugosa, Anadara grandis, Protothaca metodon, and Polymesoda radiata. e From Blake et al. (1992a: tables 1 and 2). f Sample size (MNI) of vertebrate remains only.

at dish-dominant sites, indicating that these were base camps to which task groups brought resources collected from a variety of habitats. No such pattern is evident in Table 4.2. There are more taxa of reptiles, amphibians, and mammals at the dish- dominant sites, but for birds and fish the pattern is inverted, with more taxa represented at El Varal. Table 4.3 provides a more detailed look at diversity among fish, the most common vertebrate class at all three sites. The results are the same. Fish are more diverse at El Varal despite that site’s lower total NISP. In light of Table 4.3, we are prompted to ask to what extent the Early Formative faunal assemblages could be the result of generalized foraging in the vicinity of each site. A greater diversity of terrestrial fauna (reptiles, amphibians, and mammals) at inland sites could actually be consistent with our null hypothesis. To explore these issues further we consider two more topics: (1) habitat use in relation to

80

archaic to formative

site location and (2) the degree to which the El Varal fauna represents a subset of that at Paso. According to our null hypothesis, differences among the three assemblages should be explicable with reference to the habitats most readily accessible from each site. At El Varal we would expect an emphasis on fish species of the lower estuary, generally those oriented toward more saline habitats. Aquiles Serdán, by contrast, is near the freshwater Cantileña Swamp. The species recovered in a fishing trip we took to the swamp in 1997 were freshwater in focus (Table 4.4), and under our null hypothesis we would expect a similar pattern for Aquiles Serdán. Paso de la Amada lies inland from both swamp and estuary on the Coatán delta. Still, the site is located beside an abandoned river course that even today fi lls with water in the rainy season. Before twentieth- century drainage operations, it was possible to fish there during the rainy season. It is thus likely that the Early Formative residents of Paso de la Amada had direct access

TABLE 4.3 Two Measures of Diversity of Fish Genera at the Three Early Formative Sites

Diversity by MNI

Diversity by NISP

No. Taxa

Simpson’s D

Total MNI

No. Taxa

Simpson’s D

Total NISP

23

0.73

210

23

0.53

641

14

0.69

91

16

0.34

1,227

5

0.53

286

Tecomate- dominant El Varal Dish- dominant Paso de la Amada Aquiles Serdán

note: The first measure is a simple count of the number of taxa present at the genus level (higher levels were included only if Wake assigned an MNI value to the taxon in question). The second is Simpson’s diversity measure, described by Pielou (1969:223–224). D can vary from 0 to 1. The minimum possible diversity of 0 corresponds to the case in which all specimens are assigned to the same taxon, whereas the maximum diversity of 1 would be the case in which each specimen is from a different taxon.

TABLE 4.4 Most Important Fish Recovered on Wake and Lesure’s 1997 Fishing Trip to the Hueyate Swamp WITH DATA ON HABITAT

Environment comments

N

x

Freshwater

63

x

Freshwater

56

x

Mainly freshwater, tolerates marine

30

Atractosteus tropicus

x

Freshwater, rivers and lakes

1

Cichlasoma sp.

x

Freshwater

1

Rank

Taxon

1

Amphilophus macracanthus

2

Cichlasoma trimaculatum

3

Dormitator latifrons

4/5 4/5

M

x

Br

x

FW

note: M = marine habitats; Br = brackish water habitats; FW = freshwater habitats. Habitat information (including comments on environment) are from www.fishbase.com (Froese and Pauly 2000, consulted May 2008).

to the upper estuary system within the immediate vicinity of the site, at least during the rainy season. We would expect an aquatic fauna with salinity preferences intermediate between the lower- estuary focus of El Varal and the freshwater-swamp focus of Aquiles Serdán. The alternative to all this is the collecting model, which would postulate transfer of fish

from estuary sites to the two dish- dominant sites. Relevant information is provided in Table 4.5. The top fish genera at each site are ranked by MNI. The table also includes assessments of the appearance of the different taxa in marine, brackish, and freshwater habitats as well as further notes on environmental preferences of the

adaptive, or gan i za tion al transformation

81

TABLE 4.5 Most Important Fish at the Three Sites, Ranked by MNI WITH DATA ON HABITAT

Rank

Taxon

M

Br

FW

Environment comments

MNI

NISP

102

447

El Varal 1

Ariidae

x

x

x

Mainly marine-rackish

2

Centropomus sp.

x

x

x

Estuaries, lagoons

33

57

3

Lutjanus sp.

x

x

Marine, tolerates brackish

17

51

4

Batrachoides waltersi

x

x

x

Common in bays, tolerates freshwater

13

17

5

Eleotris sp.

x

x

Mainly freshwater

9

17

Paso de la Amada 1

Ariidae

x

x

x

Mainly marine-brackish

47

994

2

Dormitator latifrons

x

x

x

Mainly freshwater, tolerates marine

19

57

3

Cichlasoma sp.

x

Freshwater

7

44

4

Lutjanus sp.

x

x

Marine, tolerates brackish

4

32

5

Centropomus sp.

x

x

Estuaries, lagoons

4

35

x

Aquiles Serdán 1

Cichlasoma sp.

2

Ariidae (Sciades sp.)

3

Atractosteus tropicus

4

Lutjanus sp.

x

x

x

x

x

Freshwater

x

Mainly marine-brackish

97

x

Freshwater, rivers, and lakes

15

Marine, tolerates brackish

171

3

note: M = marine habitats; Br = brackish water habitats; FW = freshwater habitats. Habitat information (including comments on environment) from fishbase.com (Froese and Pauly 2000, consulted May 2008).

different fish. That information was obtained online at fishbase.com (Froese and Pauly 2000, consulted May 2008). Given that many of our identifications are to genus rather than species, we have tried to summarize habitat information generally characteristic of the probable

82

archaic to formative

range of species that may have been present at El Varal. Given, first, that there is no simple way to associate a par ticular habitat or even a general level of salinity with a par ticular fish species because many shift among habitats on a regular basis or over the course of their lives,

and, second, that the archaeological identifications are often to genus level but not species, there would appear to be some support for both generalized foraging and collecting as contributing to the fish of dish- dominant sites. Most fish identified could have been caught at a variety of places within the estuary-lagoonswamp system. However, the top four fish at El Varal are marine and brackish-water in focus, whereas the first- and third-ranked fish at Aquiles Serdán prefer freshwater. There would appear to be something of the gradient in salinity preferences expected under our null hypothesis as we move from Varal (high salinity) to Paso (intermediate) to Aquiles Serdán (low salinity). While we see a component of generalized foraging, there are some hints of transport of species to dish-dominant sites from spots beyond an arbitrarily defined 5 km catchment. Although we cannot prove that animals deriving from distances of greater than 5 km were harvested with a collecting as opposed to foraging strategy, such a suggestion at least starts to become plausible. The observation of most interest is the high rank of marine catfish at all sites. We believe marine catfish would have been unavailable in the Hueyate Swamp and thus anywhere in the vicinity of Aquiles Serdán. A mixture of the two strategies may thus have been involved at the sites, but what we find most striking is that the faunal assemblages at dishdominant sites seem to bear a strong imprint of foraging undertaken a short distance from the residential base. For a final look at the problem we examine the nature of the overlap in fauna represented at tecomate- dominant El Varal and dishdominant Paso de la Amada. In the collecting model, the fauna at residential bases (Paso de la Amada) are expected to be diverse because foods are transported there from a variety of different locations. The taxa identified at a field camp (El Varal) should constitute a subset of those recovered from the residential base. Of the Varal MNIs, 75 percent fall into taxa shared with the inland site and 25 percent are from

taxa identified only at El Varal. Among the Varal MNIs assigned to taxa shared with Paso, 96 percent were in taxa that would have been available within 5 km of the inland site. In contrast, of the total MNI of fifty-three for taxa identified only at Varal, 61 percent were in taxa with a lower estuary, lagoon, or marine focus not available within 5 km of Paso de la Amada. These last include one of the top five fish genera from El Varal (Batrachoides waltersi) as well as three other reasonably important Varal taxa (Eleotris picta, Diapterus sp., and Caranx sp.). Alternatively, the shared taxa that were from outside the immediate catchment of Paso (Sciaenidae and Cheloniidae) seem to have been of minor importance to the diet at either site. Again, the overlap in faunal assemblages seems explicable with reference to overlap in the habitats easily accessible from the two sites, while divergences between them seem mainly related to differences in catchment. Both points bolster the null hypothesis. In sum, comparisons of faunal remains at  three Early Formative sites suggest that the  majority of wild animals consumed at each site— irrespective of whether its vessel assemblage was dish- dominant or tecomatedominant—was procured through practices more akin to generalized foraging than collecting. While residing at any of these locations, fisher-hunters probably foraged for a variety of locally available foods, returning to the site each day to consume their catch and share with others. As a result, terrestrial animals are most common at inland sites, freshwater fish most common at sites near the swamp, and estuary or lagoon fish most common at sites beside estuaries and lagoons. In sum, faunal remains by themselves provide no obvious basis for explaining the divide between tecomate- dominant and dish- dominant assemblages. On the basis of these results, we would appear to have three options for pushing the investigation further. First, if we are not yet ready to reject logistical collecting of foods as the source of intersite differences, then we need to posit a narrow focus on a few fish species

adaptive, or gan i za tion al transformation

83

(especially sea catfish) or else on the collecting of foods that would have left no macroremains at all. Second, we could interpret the faunal evidence from Varal as the remains of meals eaten by task group members while they produced some nonsubsistence resource. Finally, we could revisit the (foraging) idea of moving people to resources. The concentration of aquatic resources would have been at a peak in the dry season of the lower estuary and correspondingly scarce at inland sites. Could the tecomatedominant sites have been seasonal outposts for whole communities rather than for specialized task groups? We examine the first two of these options in the next section, then move toward conclusions that draw on all three.

Salt Production and the Issue of Specialization Of the possible products for tecomate-dominant sites beyond those considered in the last section, salt seems by far the most likely. After brief mention of other potential products, we present the case for salt production as a significant activity at El Varal. We then review various strands of argument suggesting that this case should not be understood as one of occupational specialization. Carballo et al. (2009) examine the possibility of specialized production of tecomates based on chemical characterization of sherds from tecomate- dominant El Varal and dishdominant Cantón Carralito. Results indicate that both dishes and tecomates were mainly made at the sites in which they were recovered. A minimally plausible alternative is that Varal could have produced some sort of boiled fish product that would have involved destruction of the bones through cooking (Smith’s [1997] gefilte fish hypothesis). There is no space to fully review this issue here, but we now consider it unlikely (see Lesure 2009c). A third possible product is shrimp, championed by Voorhies (2004:147– 157) for the Archaic. Shrimp swarm in local lagoons during the dry season (Voorhies 2004:150– 151). They are

84

archaic to formative

produced in bulk in Chiapas today (1,000,000 pounds annually in the early 1940s [Lindner 1944]). Further, the local method of production involves brief boiling– a positive point given our tecomates—followed by sun drying (Lindner 1944:79; Voorhies 2004:152). Harvesting and drying of shrimp for transfer to inland sites is a possibility for El Varal, but we have no actual evidence of such a practice. In contrast to products for which justification requires grasping at straws, a case for the manufacture of salt at El Varal appears strong. Late sixteenth-century aboriginal salt making along the Guatemala coast involved scraping up salt-laden earth and placing it in wooden troughs (Coe and Flannery 1967:92). Water poured over the earth would trickle out below as brine, which was then boiled and reduced to salt. Andrews (1983:62– 63) noted mounds of earth along the Chiapas coast, which he took to be relatively recent waste heaps from this sal cocida process. Nance’s (1992) excavation of the Guzmán Mound in the Naranjo region of nearby Guatemala appears to document similar practices in the Late Formative. The mound lies beside a salt flat, and the stratigraphy consisted of layers of “grey loam” alternating with refuse layers containing sherds, burnt earth, and charcoal (Nance 1992:29), strikingly similar to the pattern observed at El Varal. Pottery consisted overwhelmingly of plain, crudely finished jars with either slight necks and relatively open profiles or distinct, vertical necks and closed profiles. Santley (2004) identified Early Formative salt production (coeval with El Varal) beside a salt spring in the Tuxtla Mountains of Veracruz. The patterns he noted included densities of broken pottery among the highest known from the region; a vessel assemblage dominated by coarse ceramics, among which tecomates are prominent; lack of architectural features, storage pits, or burials; and evidence of burning on the exteriors of pots. Something of a contrast is provided by a third case, El Mesak, at the southeastern extreme of the Soconusco. As already mentioned, the sal cocida process in that case seems to have involved boiling brine in deep,

conical, open-profile vessels, which dominate the Jocotal assemblage. The positive indications for salt production at El Varal include the evidence of burning, the quantities of utilitarian pots of the same form, and the overall stratigraphy of the mound. Salt production of the sort described for the region in the sixteenth century and seemingly documented for the Late Formative is particularly attractive in that it provides an explanation for the layers of silt, with few artifacts, registered in the Vásquez Mound profile, features for which we otherwise have no satisfying explanation. Finally, given the overabundance of tecomates in relation to animal remains, the boiling of brine provides an attractive explanation for the contents of tecomates. We would argue, then, that salt was a significant focus of production at El Varal. How was production organized? One obvious possibility is part-time, seasonal occupational specialization. In such a scenario, social units of production would have constituted a subset of households from inland communities. Those people would have produced salt at an estuary camp for some part of the year and transferred their product to nonspecialist consumers through exchange or tribute. Four sets of observations lead us to reject that line of argument in favor of an organizationally simpler proposal, to be developed in the next section. First, while comparison with the Late Formative Guzmán Mound (Nance 1992) helps clinch the case for salt production, there are significant contrasts. At El Varal, there were notably fewer formal hearth features, more artifacts not related to salt production, and a much greater frequency of faunal remains. The extraordinary paucity of animal bones at Guzmán is particularly striking (total vertebrate NISP of 4, or about 0.2 specimens/m3, whereas there were 100 to 300 specimens/m3 at Varal). It would appear that activities at Guzmán were particularly concentrated on salt production, while the occupants of El Varal engaged in more of a spectrum of activities.

Second, the Varal tecomates seem unlikely as tools of specialized salt production. For a long time we resisted identifying salt as a product at El Varal because the restricted-mouthed tecomate seemed an inappropriate vessel form when the goal was evaporation rather than prolonged simmering. The vessel assemblages apparently associated with salt manufacture at Guzmán (Nance 1992) and El Salado (Santley 2004) have helped convince us to revise that position. While open vessels are preferable for evaporating brine, variability among convincing cases of pre-Hispanic sal cocida production (McKillop 2002; Nance 1992; Pye 1995; Santley 2004) suggests lack of any strong selection for a par ticular vessel form. Still, where occupational specialization seems supported by other sorts of indicators, two vessel characteristics recur: there is some kind of appropriately openmouthed evaporation technology, and the coarse ware involved in production is extremely crude. Neither of these characterize El Varal. We would expect specialists to have come up with something better than the tecomate, and for utilitarian ware, the tecomates themselves were embarrassingly well decorated with multiple contrasting bands involving slipping, burnishing, brushing, plastic modification, and incised motifs (well illustrated for Salinas la Blanca in Coe and Flannery 1967). The Varal tecomates do not seem like strictly expedient tools of salt manufacture and we are led to wonder if they might have been used for other things besides reducing brine. Third, assuming we are correct to claim that the adaptation of late Early Formative inhabitants of El Varal was part of a pattern of estuary use that first appeared in the Locona phase, a series of more general observations cast further doubt on Varal as being involved in the specialized production of salt. A striking feature of the Early Formative pattern is the rapidity of its emergence in the Locona phase. We believe that the Pampa Cabildo contains multiple tecomate- dominant sites in use as seasonal camps from this early period. We are proposing a rapid proliferation of such sites, and

adaptive, or gan i za tion al transformation

85

we have trouble envisioning occupational specialization in the production of salt being formulated so widely, so rapidly. A shift to a focus on maize agriculture could occasion a much elevated demand for salt (Pye 1995:312–313), but such a shift would not fit with the isotope and settlement evidence—together with all the evidence of estuary fauna— suggesting a broad subsistence base in the Early Formative. There should not have been demand at a level to support specialized production of salt of the scale implied by the frequency of tecomate- dominant sites. The situation would have been different in the Late Formative, when settlement had withdrawn inland and maize agriculture was clearly the focus of subsistence. The Guzmán Mound is one of numerous Late Formative mounds along salt flats of the lower Naranjo area (Coe and Flannery 1967; Nance 1992). The producers in that case were probably part-time occupational specialists responding to significant regional demand. If the tecomate- dominant sites of the Early Formative supported salt production but not specialization, what sort of trajectory led from there to the very different Late Formative pattern? El Mesak hints at an answer. The evaporation technology matching that expected of specialized production in that case grew in importance over the Early Formative until it dominated the assemblage by the Jocotal phase (Pye 1995). If that technology can be taken as a signal of specialized production of salt for the Jocotal phase, then such productive organization was emergent but still rare at the end of the Early Formative, much as we might expect in the context of a regional subsistence system still on the verge of significant further steps in the growing importance of maize. We would propose that, in the Jocotal phase, the sites of El Varal and El Mesak were organizationally distinct, the former associated with adaptive patterns that were quite ancient, the latter presaging what was to come. Our fourth point is one we will elaborate in the next section. El Varal seems to have been occupied primarily in the dry season (Kennett

86

archaic to formative

and Culleton 2009). That should have been a good time to produce salt, but it is also the season in which aquatic resources would have been most scarce inland and at their peak in the estuary-lagoon system. We propose that the Early Formative occupants of tecomatedominant sites pursued a variety of estuary resources beyond salt. Acquisition of resources was generalized, and organizational patterns involved a combination of foraging and collecting strategies—both moving people to resources and resources to people, depending on the product—while products themselves flowed through generalized social obligations such as relations of kinship and alliance. Carballo et al. (2009) conducted an analysis of lipid residues in sherds from Varal and Cantón Carralito. The work was done while analytical practices and standards were still being worked out in this emerging field. Until the study can be repeated with a larger sample and updated methods, we propose to take the results seriously: that the Varal tecomates were probably used to cook a variety of contents that overlap with the contents of tecomates at the dish-dominant, inland site of Cantón Carralito.

EARLY FORMATIVE SETTLEMENT AND SUBSISTENCE IN MAZATÁN The Early Formative settlement system of the Mazatán area of the Soconusco— and possibly other areas of the Pacific coast as well— comprised permanent habitation sites (with dish-dominant vessel assemblages) located just inland of the estuary and seasonal camps (with tecomate-dominant assemblages) located in the estuary. The parties occupying seasonal camps were large and diverse, and at least some people spent weeks or even months there. We have pursued two questions concerning these camps: what was produced, and how was production and consumption organized? Evidence from El Varal provides the following answers. Salt was an important focus of activity. Dry season visitors to the estuary scraped up saltladen earth from areas exposed by seasonally

receding lagoon waters, extracted a brine, and boiled that down to salt in tecomates. Salt may indeed have been the fundamental raison d’être of tecomate- dominant sites in that, given the technology involved, processing had to take place on location rather than at inland villages. In light of the evidence against this as a case of occupational specialization, we would propose that production of salt was organized in a fashion approximating the collecting model. Task groups drew membership broadly, from numerous social units of consumption, and products flowed to consumers through very basic relations of social obligation such as kinship or marriage. While tecomate- dominant sites were salt production locales, they were also much more. The dry season would have been an ideal time of the year for harvesting a variety of estuary resources, including salt, shrimp, and fish. Further, by the later dry season, aquatic resources would not be generally available inland on the Coatán delta (except for the river itself). The occupants of El Varal seem to have foraged for a whole variety of animals in the vicinity of the site. Differences in the resulting faunal assemblage from those recovered at inland sites are largely explicable with reference to variation in the habitats readily assessable from each location. Practices of resource acquisition at El Varal were not different from those at inland sites, and much of what was collected around the site was probably eaten there—but by whom? The full spectrum of evidence—including the strong contribution of generalized foraging to the faunal assemblage and densities of animal remains both several orders of magnitude higher than observed at the Guzmán Mound and comparable to those observed at inland Early Formative sites— suggests a settlement system characterized by considerable dry season mobility. Visitors to El Varal were numerous, and we suspect that they included people well beyond those directly involved in salt production. In other words, collection and acquisition of wild animal resources in Early Formative Mazatán resembled forag-

ing in the organizational sense: people acquired much of their animal foods from the immediate vicinity of wherever they happened to be, and when, in the dry season, resources became scarce around their primary residences, they spent significant amounts of time camping out in the estuary. Still, we do not think Early Formative wild resource acquisition was restricted to generalized foraging. Although in terms of evidence the ice is rather thin here, processing of certain estuary foods for storage and transfer to inland sites would help explain the interest in salt. We propose shrimp and sea catfish as possible products. There is no par ticular evidence either for or against shrimp. As for sea catfish, an examination of Ariidae bones by element suggests that cranial bones are represented in similar proportions at both Varal and Paso (Lesure, Wake, and Steadman 2009), so unfortunately heads seem not to have been cut off at estuary camps. Voorhies (2004:408), however, notes that fish dried in the Acapetahua area today are transferred inland with all bones still attached.

THE ARCHAIC-FORMATIVE TRANSITION The Archaic-Formative adaptive transformation in Soconusco touched off a trajectory of demographic expansion similar to those associated with other, simultaneous Formative transitions in different settings across Mesoamerica. Neff et al. (2006) propose that technological innovation in a situation of climate stress was the key to the local Early Formative readaptation and, ultimately, its demographic success. Although they mention agricultural intensification, their emphasis is on technology for the recovery of aquatic animals. We began by expressing interest in that general scheme but skepticism about the role of recovery technology; we were inclined, instead, to account for demographic success in relation to the traditional Formative triumvirate of agriculture, sedentism, and pottery. The insights into Early Formative adaptations presented here support our emphasis on processing technology

adaptive, or gan i za tion al transformation

87

(pottery) over recovery technology (nets and fish hooks) and provide a basis for refining understandings of the interrelations between agriculture, sedentism, and pottery in the local Early Formative readaptation. Archaic visitors to the estuary foraged for fish as they concentrated on the bulk processing of marsh clams for transport to inland base camps. Early Formative peoples abandoned clams as a bulk crop. They included an expanded range of shellfish in their diet and added crab. They ate a spectrum of vertebrates similar to that of their Archaic predecessors, although they were more apt to concentrate on a few species of fish. This last shift downward in evenness may mean that Early Formative peoples focused their fishing efforts in a few specific habitats of the estuary-lagoon system. That sort of change might indeed have been related to innovations in recovery technology, but given the concentration on fewer species and lack of any par ticular abundance of fish remains in comparison to Archaic sites, we have trouble envisioning the innovations involved as key to Early Formative demographic success. Instead, rounded-bottom ceramic tecomates placed over direct heat would appear to have much greater revolutionary potential in the specific historical and technological context of final Archaic lifeways in Soconusco. This technological innovation shifted the adaptive landscape by opening up for exploitation a range of estuary resources hitherto of only minor significance. Most dramatically, tecomates could be used to reduce brine to salt, which in turn opened up new techniques for food preservation, potentially making other foods more attractive than clams as objects of bulk harvesting. Still, given the array of evidence reviewed here, we are proposing that Early Formative visitors to the estuary used the tecomate as a generalized rather than specialized tool, applying it to the processing of a whole variety of resources. Tecomates could be used to boil shrimp before drying. In the case of fish, salt produced in tecomates could have been used in combination with sun drying rather than boil-

88

archaic to formative

ing. But tecomates were also used to boil an expanded variety of foods for immediate consumption. Those might have included fish, crab, or shellfish— even, as Coe and Flannery (1967:102) suggested, tamales. The tecomate was a key innovation in intensified production of estuary resources. Realizing the potential of this new tool required a shift in settlement strategy. If a wider spectrum of foods was subject to processing for storage and transfer to permanent residential bases, then transport costs would have risen, conceivably favoring relocation toward the vicinity of the resources. However, reliance on tecomates for immediate consumption of estuary food would have been another factor promoting settlement reorganization: when you put the crab on to boil you need to have your consumers at hand. Use of direct heating in ceramic vessels would thus have favored a shift toward the strategy of moving people to resources as compared to the more strongly logistic Archaic strategy of estuary exploitation. Yet the Early Formative peoples also did not choose full foraging in which entire groups shifted continually around among residential bases. Instead, they moved close to, but not simply to, the estuary and established sedentary villages. In the Mazatán area, significant investment in earthen platforms and substantial residential architecture dates to the same phase as the appearance of the rounded-bottom cooking pot. It would appear that consumption of estuary resources was not the only factor governing settlement patterns. It seems most likely that the availability of arable land and the need to reside in proximity to gardens or fields in which significant labor had been invested were important concerns. The Early Formative readaptation would thus have involved intensification on two fronts. Intensification of cultivation focused on greater inputs of labor in the tending of gardens and fields, thus favoring sedentism. At the same time, the tecomate was used to diversify and intensify the exploitation of wild foods in the estuary. Realization of the two goals together

required reciprocal compromises. The estuary strategy favored relocation toward the source of the resources, but there was insufficient arable land in the estuary itself. So Early Formative villagers congregated particularly in areas like the Coatán delta, where they could simultaneously achieve acceptable returns from both their terrestrial and aquatic intensification strategies at the shortest possible distance from each other. They established permanent settlements on arable land, planting crops and elaborating residential architecture. Still, the strategy of locating the settlements as close to the estuary as possible allowed for considerable mobility of individuals (in estuary-focused task groups) and probably even of larger groups—namely, entire units of consumption. This sort of flexibility in the relative positioning of people and resources—the ability to shift back and forth between collecting and foraging— allowed the full potential of the tecomate to be realized. Acquisition of estuary resources was thus organized as a combination of what Binford (1983) refers to as collecting and foraging strategies. What we are proposing for Early Formative adaptations is a flexible set of strategies for the simultaneous intensification of several sorts of resources. They involved both increased sedentism and, in a way, increased mobility (since in the dry season whole consumption units camped at estuary field stations for days at a time). This general set of settlement and subsistence strategies was probably both widespread in the Soconusco and durable, persisting from 1700 b.c. until perhaps 700 b.c. The par ticular pattern of intersite assemblage variation described here is one archaeological manifestation of these practices. Still, given variation in habitats throughout the Soconusco, we would not be surprised if the dish- dominant versus tecomate- dominant scheme proves too simplistic for characterizing the full range of assemblage variation between Early Formative sites. The rapid emergence of the pattern of intersite assemblage variation suggests that

the basic Early Formative adaptation did not involve specialization in the organizational sense. Instead, the associated adaptive patterns formed the organizational ground out of which true occupational specialization arose. Extrapolating from the cases of El Mesak and Guzmán, we would suggest that specialization arose fi rst as limited local experiments, which could themselves be ephemeral rather than evolutionary. (Interestingly, the technological apparatus of specialization of El Mesak appears to have completely died out, with Guzmán involving a reinvention of the organizational scheme with an unrelated technological complex.) It seems possible that in a situation of very low overall social heterogeneity it would have taken considerable time to generate consumer demand and exchange mechanisms sophisticated enough that demand could be reliably met without overproduction. We have suggested that the transition to the Early Formative in the Soconusco involved four elements observed in simultaneous shifts in other parts of Mesoamerica: greater sedentism, the adoption of pottery, some sort of agricultural intensification, and the initiation of a trajectory of demographic expansion. According to traditional understandings, the first three elements account for the fourth. However, recent work on the distinctiveness of Early Formative subsistence in the Soconusco leaves the local applicability of this traditional understanding unclear. Our efforts to clarify the Early Formative settlement and subsistence system of the Mazatán area end up bolstering the traditional view. In a broad sense, sedentism, agriculture, and pottery account for the population expansion initiated locally in the Early Formative. Still, there was plenty of local specificity in the manifestation of and interrelations between these elements. Roundedbottom tecomates that could be used to heat contents directly over a fire were of par ticular significance in Soconusco, but apparently not in highland settings. Sedentism in Soconusco involved serious investments of labor in platforms and architecture, but it was accompanied

adaptive, or gan i za tion al transformation

89

by flexibility in the positioning of people across the landscape and even by seasonal mobility between villages and estuary outposts. NOTES 1. For similar gradual changes in the fire- cracked rock and grinding stones, see Clark, Pye, and Gosser 2007. 2. One methodological challenge is the confusing and fluctuating taxonomy of marine catfi shes (Ariidae; Cook et al. 2004:266–268). The most common Ariidae genus in our collection (Sciades) changed names several times during Wake’s work on the Varal fauna and now includes at least two species identified at El Varal that had previously been classified as separate genera (S. guatemalensis [formerly Hexanematichthys guatemalensis] and S.  troscheli [formerly Notarius troscheli]). In our analyses, we have decided to treat all identified Ariidae as a single taxon. Note also that we have calculated MNI by stratum (generally, lot) at El Varal, a decision supported by the well- defi ned stratigraphy and the stratigraphic precision that proved possible because of the excavation context, in which we were digging into a profi le exposed by the bulldozer.

REFERENCES Ambrose, Stanley H., and Lynette Norr 1992 On stable isotopic data and prehistoric subsistence in the Soconusco region. Current Anthropology 33: 401–404. Andrews, Anthony P. 1983 Maya Salt Production and Trade. University of Arizona Press, Tucson. Arnold, Philip J. III 1999 Tecomates, Residential Mobility, and Early Formative Occupation in Coastal Lowland Mesoamerica. In Pottery and People: A Dynamic Interaction, edited by James M. Skibo and Gary M. Feinman, pp. 157– 170. University of Utah Press, Salt Lake City. Arroyo, Bárbara 2004 Of Salt and Water: Ancient Commoners on the Pacific Coast of Guatemala. In Ancient Maya Commoners, edited by J. C. Lohse and F. Valdez, pp. 73– 94. University of Texas Press, Austin. Binford, Lewis R. 1983 Willow Smoke and Dog’s Tails: Hunter[1980] Gatherer Settlement Systems and Archaeological Site Formation. In Working at Archaeology, pp. 337–356. Academic Press,

90

archaic to formative

New York. [Original: American Antiquity 45(1): 4–20.] Blake, Michael, Brian Chisholm, John E. Clark, and Karen Mudar 1992a Non-agricultural Staples and Agricultural Supplements: Early Formative Subsistence in the Soconusco Region, Mexico. In Transitions to Agriculture, edited by T. D. Price and A. B. Gebauer, pp. 133– 151. Prehistory Press, Madison, WI. Blake, Michael, Brian Chisholm, Barbara Voorhies, George Michaels, Michael W. Love, and John E. Clark 1992b Prehistoric subsistence in the Soconusco region. Current Anthropology 33: 83– 94. Blake, Michael, Richard G. Lesure, John E. Clark, Warren D. Hill, and Luís Barba 2006 The Residence of Power at Paso de la Amada, Mexico. In Palaces and Power in the Americas: From Peru to the Northwest Coast, edited by J. J. Christie and P. J. Sarro, pp. 191–210. University of Texas Press, Austin. Carballo, David M. 2001 Geochemical Approaches to Explaining the Ceramic Patterning of Early Formative Villages in Chiapas, Mexico. Unpublished Masters thesis, Department of Anthropology, University of California, Los Angeles. Carballo, David M., Richard G. Lesure, Jelmer Eerkins, Douglas J. Kennett, Stuart Tyson Smith, Hector Neff, and Michael D. Glascock 2009 The Manufacture and Content of Pottery Vessels in Early Formative Mazatán. In Settlement and Subsistence in Early Formative Soconusco: El Varal and the Problem of Inter-Site Assemblage Variation, edited by Richard G. Lesure, pp. 223–237. Cotsen Institute of Archaeology Press, University of California, Los Angeles. Ceja Tenorio, Jorge Fausto 1999 Los Alvarez. Un campamento prehispánico de pesca y recolección de mariscos en el Soconusco, Chiapas, México. In Taller Arqueología de la Región de la Costa Sur de Guatemala, edited by Christa Schieber de Lavarreda, pp. 109– 125. Ministerio de Cultura y Deportes, Guatemala. Chisholm, Brian and Michael Blake 2006 Diet in Prehistoric Soconusco. In Histories of Maize: Multidisciplinary Approaches to the Prehistory, Linguistics, Biogeography, Domestication, and Evolution of Maize, edited by John E. Staller, Robert H. Tykot,

Bruce F. Benz, pp.161– 172. Elsevier Academic Press, Boston. Chisholm, Brian, Michael Blake, and Michael W. Love 1993 More on prehistoric subsistence in the Soconusco region: Response to Ambrose and Norr. Current Anthropology 34: 432–434. Clark, John E. 1991 The beginnings of Mesoamerica: Apologia for the Soconusco Early Formative. In The Formation of Complex Society in Southeastern Mesoamerica, ed. by W. R. Fowler, Jr., pp. 13– 26. CRC Press, Boca Raton, FL. Clark, John E. 1994 The Development of Early Formative Rank Societies in the Soconusco, Chiapas, Mexico. Ph.D. dissertation, Department of Anthropology, University of Michigan, Ann Arbor. Clark, John E., and Michael Blake 1994 The Power of Prestige: Competitive Generosity and the Emergence of Rank Societies in Lowland Mesoamerica. In Factional Competition and Political Development in the New World, edited by E. M. Brumfiel J. Fox, pp. 17 –30. Cambridge University Press, England. Clark, John E., and Dennis Gosser 1995 Reinventing Mesoamerica’s First Pottery. In The Emergence of Pottery: Technology and Innovation in Ancient Societies, edited by W. Barnett and J. Hoopes, pp. 209–221. Smithsonian Institution Press, Washington, DC. Clark, John E., and Mary E. Pye 2000 The Pacific Coast and the Olmec Question. In Olmec Art and Archaeology in Mesoamerica, edited by J. E. Clark and M. E. Pye, pp. 217–251. Studies in the History of Art No. 58, National Gallery of Art, Washington, DC. Clark, John E., Mary E. Pye, and Dennis C. Gosser 2007 Thermolithics and Corn Dependency in Mesoamerica. In Archaeology, Art, and Ethnogenesis in Mesoamerican Prehistory: Papers in Honor of Gareth W. Lowe, edited by Lynneth S. Lowe and Mary E. Pye, pp. 23–42. Papers of the New World Archaeological Foundation No. 68, Brigham Young University, Provo, UT. Coe, Michael D., and Kent V. Flannery 1967 Early Cultures and Human Ecology in South Coastal Guatemala. Smithsonian Contribu-

tions to Archaeology No. 3, Smithsonian Institution Press, Washington, DC. Cooke, Richard G., Máximo Jiménez, Conrado Tapia, and Barbara Voorhies 2004 A Closer Look at the Late Archaic Fish Fauna. In Coastal Collectors in the Holocene: The Chantuto People of Southwest Mexico, edited by Barbara Voorhies, pp. 207–299. University Press of Florida, Gainesville. Dietler, John, and Thomas A. Wake 2009 Crab Exploitation in Early Formative Soconusco. In Settlement and Subsistence in Early Formative Soconusco: El Varal and the Problem of Inter-Site Assemblage Variation, edited by Richard G. Lesure, pp. 89– 98. Cotsen Institute of Archaeology Press, University of California, Los Angeles. Feddema, Vicki 1993 Early Formative Subsistence and Agriculture in Southeastern Mesoamerica. Masters thesis, University of British Columbia. Froese, R, and D. Pauly (eds.) 2000 Fishbase 2000: Concepts, Design and Data Sources. ICLARM, Los Baños, Laguna, Philippines. Electronic document, available at http://fishbase.com. Kennett, Douglas J., and Brendan J. Culleton 2009 Shellfish Harvesting Strategies at El Varal. In Settlement and Subsistence in Early Formative Soconusco: El Varal and the Problem of Inter-Site Assemblage Variation, edited by Richard G. Lesure, pp. 173– 178. Cotsen Institute of Archaeology Press, University of California, Los Angeles. Kennett, Douglas J., Barbara Voorhies, and Dean Martorana 2006 An Ecological Model for the Origins of Maize-Based Food Production on the Pacific Coast of Southern Mexico. In Behavioral Ecology and the Transition to Agriculture, edited by D. J. Kennett and B. Winterhalder, pp.103– 136. University of California Press, Berkeley. Lesure, Richard 1993 Salvamento arqueológico en El Varal. Una perspectiva sobre la organización sociopolítica olmeca de la costa de Chiapas. Segundo y Tercer Foro de Arqueología de Chiapas, pp. 211–227. Instituto Chiapaneco de Cultura, Mexico. Lesure, Richard G. 2009a Artifact Synthesis and Intra-Site Assemblage Variability. In Settlement and Subsistence in Early Formative Soconusco: El Varal and the Problem of Inter-Site Assem-

adaptive, or gan i za tion al transformation

91

blage Variation, edited by Richard G. Lesure, pp. 179– 199. Cotsen Institute of Archaeology Press, University of California, Los Angeles. Lesure, Richard G. (ed.) 2009b Settlement and Subsistence in Early Formative Soconusco: El Varal and the Problem of Inter-Site Assemblage Variation. Cotsen Institute of Archaeology Press, University of California, Los Angeles. Lesure, Richard G. 2009c Site Assemblage Variation in Early Formative Soconusco. In Settlement and Subsistence in Early Formative Soconusco: El Varal and the Problem of Inter-Site Assemblage Variation, edited by Richard G. Lesure, pp. 3–21. Cotsen Institute of Archaeology Press, University of California, Los Angeles. Lesure, Richard G. 2009d The Organization of Salt Production: Specialization or Collection? In Settlement and Subsistence in Early Formative Soconusco: El Varal and the Problem of Inter-Site Assemblage Variation, edited by Richard G. Lesure, pp. 239–250. Cotsen Institute of Archaeology Press, University of California, Los Angeles. Lesure, Richard G. 2011 Comment on “Backward Bottlenecks: Ancient Teosinte/Maize Selection,” by David Webster. Current Anthropology 52(1): 95– 96. Lesure, Richard G., and Isabel Rodríguez López 2009 Pottery. In Settlement and Subsistence in Early Formative Soconusco: El Varal and the Problem of Inter-Site Assemblage Variation, edited by Richard G. Lesure, pp. 115– 146. Cotsen Institute of Archaeology Press, University of California, Los Angeles. Lesure, Richard G., Thomas A. Wake, and David W. Steadman 2009 Subsistence in the Estuary: Surplus Production, Expedient Meals, or Something Between? In Settlement and Subsistence in Early Formative Soconusco: El Varal and the Problem of Inter-Site Assemblage Variation, edited by Richard G. Lesure, pp. 203–222. Cotsen Institute of Archaeology Press, University of California, Los Angeles. Lindner, Milton J. 1944 The commercial marine fishes, crustaceans and molluscs of the West Coast of Mexico. Transactions of the American Fisheries Society 74: 71– 80.

92

archaic to formative

Love, Michael W. 1999 Ideology, Material Culture and Daily Social Practice in Pre- Classic Mesoamerica: A Pacific Coast Example. In Social Patterns in Pre- Classic Mesomerica, edited by D. C. Grove and R. A. Joyce, pp. 127– 153. Dumbarton Oaks Research Library and Collection, Washington, DC. Love, Michael W. 2002 Early Complex Society in Pacific Guatemala: Settlements and Chronology of the Río Naranjo, Guatemala. Papers of the New World Archaeological Foundation No. 66, Brigham Young University, Provo, UT. Lowe, Gareth W. 1977 The Mixe-Zoque as Competing Neighbors of the Early Lowland Maya. In The Origins of Maya Civilization, edited by R. E. W. Adams, pp. 197–248. University of New Mexico Press, Albuquerque. Lowe, Gareth W. 1978 Eastern Mesoamerica. In Chronologies in New World Archaeology, edited by R. E. Taylor and C. W. Meighan, pp. 331–393. Academic Press, New York. McGuire, Randall H. 1983 Breaking down cultural complexity: Inequality and heterogeneity. Advances in Archaeological Method in Theory 6: 91– 142. McKillop, Heather 2002 Salt: White Gold of the Ancient Maya. University Press of Florida, Gainesville. Michaels, George H., and Barbara Voorhies 1999 Late Archaic Period Coastal Collectors in Southern Mesoamerica: The Chantuto People Revisited. In Pacifi c Latin America in Prehistory: The Evolution of Archaic and Formative Cultures, edited by Michael Blake, pp. 39– 54. Washington State University Press, Pullman. Nance, C. Roger 1992 Guzmán Mound: A Late Preclassic salt works on the south coast of Guatemala. Ancient Mesoamerica 3(1): 27–46. Neff, Hector, Deborah M. Pearsall, John G. Jones, Barbara Arroyo, Shawn K. Collins, and Dorothy E. Freidel 2006 Early Maya adaptive patterns: Mid-Late Holocene paleoenvironmental evidence from Pacific Guatemala. Latin American Antiquity 17 (3): 287–315. Pailles, Maricruz 1980 Pampa El Pajón, an Early Middle Preclassic Estuarine Site, Chiapas, Mexico. Papers of the New World Archaeological Foundation

No. 44, Brigham Young University, Provo, UT. Pielou, E. C. 1969 Introduction to Mathematical Ecology. Wiley-Interscience, New York. Pye, Mary E. 1995 Settlement, Specialization, and Adaptation in the Río Jesus Drainage, Retalhuleu, Guatemala. Ph.D. dissertation, Department of Anthropology, Vanderbilt University, Nashville, TN. Rosenswig, Robert M. 2006 Sedentism and food production in early complex societies of the Soconusco, Mexico. World Archaeology 38: 329–354. Rosenswig, Robert M. 2007 Beyond identifying elites: Feasting as a means to understand early Middle Formative society on the Pacific coast of Mexico. Journal of Anthropological Archaeology 26: 1–27. Santley, Robert S. 2004 Prehistoric salt production at El Salado, Veracruz, Mexico. Latin American Antiquity 15(2): 199–221. Schiffer, Michael B. 1975 Archaeology as behavioral science. American Antiquity 77: 836– 848. Shook, Edwin M., and Marion Hatch 1979 The Early Preclassic sequence in the Ocos/ Salinas La Blanca Area, south coast Guatemala. Contributions of the University of California Archaeological Research Facility 41: 143– 195. Smalley, John, and Michael Blake 2003 Sweet beginnings: Stalk sugar and the domestication of maize. Current Anthropology 44: 675– 703. Smith, Natalie Jill 1997 Economic Specialization and the Emergence of Complexity in Early Formative

Chiapas, Mexico. Unpublished Masters thesis, Department of Anthropology, University of California, Los Angeles. Steadman, D. W., M. P. Tellkamp, and T. A. Wake 2003 Prehistoric exploitation of birds on the Pacific coast of Chiapas, Mexico. Condor 105(3): 572– 579. Varien, Mark D., and Barbara J. Mills 1997 Accumulations research: Problems and prospects for estimating site occupation span. Journal of Archaeological Method and Theory 4(2): 141– 191. Voorhies, Barbara 1976 The Chantuto People: An Archaic Period Society of the Chiapas Littoral, Mexico. Papers of the New World Archaeological Foundation No. 41, Brigham Young University, Provo, UT. Voorhies, Barbara 2004 Coastal Collectors in the Holocene: The Chantuto People of Southwest Mexico. University Press of Florida, Gainesville. Voorhies, Barbara, Douglas J. Kennett, J. G. Jones, and Thomas A. Wake 2002 A Middle Archaic archaeological site on the west coast of Mexico. Latin American Antiquity 13: 179–200. Wake, Thomas A. 2004 A Vertebrate Archaeofauna From the Early Formative Period Site of Paso de la Amada, Chiapas, Mexico: Preliminary Results. In Maya Zooarchaeology: New Directions in Method and Theory, edited by Kitty F. Emery, pp. 209–222. Cotsen Institute of Archaeology Press, University of California, Los Angeles. Webster, David 2011 Backward bottlenecks: Ancient teosinte/ maize selection. Current Anthropology 52(1): 77– 104.

adaptive, or gan i za tion al transformation

93

part two

Emergent Complexity The Archaeological Records of Early Political Centers

FIVE

Building History in Domestic and Public Space at Paso de la Amada an examination of mounds 6 and 7 Michael Blake

B

y about 1900 cal b.c. , the people living along the coast of Chiapas, Mexico were beginning to settle in permanent communities. Their settlements, with substantial dwellings, some public structures, and numerous smaller features, were occupied year-round and were not simply temporary seasonal stops on a sequence of trips through the swamps and estuaries of the coastal plain.1 From our archaeological point of view—that is, looking back into the distant past with only a few scattered and fragmentary remains of these ancient settlements—this new lifestyle looks revolutionary. But, seen from the perspective of the entire Pacific coast of what is now Latin America, from Mexico south to Chile, the Early Formative coastal inhabitants of Chiapas were somewhat tardy in this settling down process. Farther to the south, on the coastal plain of Ecuador, for example, Valdivia period villages with clusters of oval-shaped houses began to appear by about 3300 b.c. (for example, Raymond 1999). Similarly, by about 2700 b.c. along the coast of Peru, there were dozens of settlements whose remains boast giant complexes of pyramids and temples, some with elab-

orate stone architecture. The complexes in the Supe River valley and Norte Chico come readily to mind (Moseley 2001; Shady Solis and Leyva 2003). Whether during the fourth millennium b.c. in Ecuador or the third millennium in Peru, or in early second-millennium Chiapas Mexico, we can discern a common thread—namely that the scale, permanence, and widespread impact of these new types of settlements transformed the way people created and used the spaces in which they lived (Bandy and Fox 2010). It is likely that earlier Archaic peoples in all these regions also transformed their landscapes by creating both residential and public spaces, each containing clearings, large shell mounds, cemeteries, buildings, and countless other features of varying sizes and types of constructions. But, almost by definition, these were generally on a smaller scale and of considerably less permanence than the kinds of features and structures constructed within sedentary villages and towns. In Chiapas this is attested to by the frustrating difficulty in locating previllage sites. The kinds of sites that stand out as the candidates for such settlements

97

are primarily shell mounds. In some cases, such as those described by Barbara Voorhies, these were built on a grand scale—partly accumulations of food debris and partly intentionally constructed surfaces rising above the surrounding estuary flood zone (Voorhies 2004:53). But it appears that although people lived on these features they did not do so permanently. Instead they used them seasonally, and these sites were but one stop on a migratory round—the other stops of which are still maddeningly obscure (Voorhies 2004:414). Recent research by John Hodgson and John Clark (2007) has begun to document an array of Archaic Period mounds in the estuaries adjacent to the plain on which Paso de la Amada sits. Some of these sites are both earlier and grander than expected and may provide evidence of a much larger and denser pre-Formative occupation than that previously documented in the Mazatán locality. It is only with the appearance of constructions in permanently occupied settlements that we begin to see new forms of social space, both residential and public, that were intended to function during the course of the whole year, year after year, for generations. The details of the constructions in these settlements are some of the most valuable and productive sources of information we have about the nature of these ancient societies. In the Mazatán zone of Chiapas, archaeological investigations during the past several decades—carried out under the umbrella of the New World Archaeological Foundation—give us a reasonably clear idea of the range of Early Formative-period communities that existed and, to some degree, the nature of their organization (Clark 2004a). One such settlement, Paso de la Amada, has been the object of investigation by teams of researchers beginning in the 1970s. The first of these was carried out by Jorge F. Ceja Tenorio who mapped and test excavated several parts of the site, outlined the ceramic sequence, and discovered several burials (Ceja Tenorio 1985). In 1985, John Clark and I embarked on a large-scale project to study Early Formative settlements in the Maza-

98

emergent complexity

tán zone—and this became the focus of his doctoral research (Clark 1994)—work which continued intermittently for a decade. During the fall of 1985, as part of this larger project, I began excavations in Mound 6 at Paso de la Amada, and then returned again in 1990, 1993, and 1995. These four seasons of work expanded to include other projects at Paso, and many colleagues including Richard Lesure, Warren Hill, Vicki Feddema, and Dennis Gosser joined us to carry out their own graduate research projects. Collectively, we kept coming back to the site because it continued to yield new information about a wide range of community activities that spanned at least five centuries of occupation (Blake et al. 1995; Clark 1994; Feddema 1993; Hill 1999; Hill, Blake, and Clark 1998; Hill and Clark 2001; Lesure 1997, 1999). The entire network of Early Formative settlements that existed in the Soconusco region, of which the site of Paso de la Amada was but a part, has not been, nor will ever be, completely documented. Even so, the partial record we do have presents a remarkable view of the complex and changing Early Formative-period settlement patterns that must have existed in the region (Clark 1994; 2004a:55). Even in its obscurity, the settlement record of this period is much better known than that of the preceding Archaic. When describing Paso de la Amada, it will be important to keep in mind that the site is just one of hundreds of sites in the region and that, because it was one of the largest communities (if not the largest one) at the time, people from the whole region likely visited it, interacting with the permanent residents who must have been friends, kin, acquaintances, trading partners, strangers, and perhaps at times even enemies. This suggests that the construction of the community, its public spaces and buildings, its residences, and so forth were visible to and referents for many more people than the primary occupants of the community alone. The social places and spaces of communities like Paso de la Amada served a wide audience and their impacts extended down through the generations (see Chapter 6 in this book). It was, in part, their

physical permanence and constant use that made this so. Mesoamerican archaeologists have themselves been active agents in developing new theoretical perspectives for helping us interpret ancient settlements and their inextricably intertwined landscapes (for example, Ashmore 2004; Clark 2004a, 2004b; Joyce 2004). These perspectives are part of a global surge in the conceptualization of ancient built environments— and the extent to which action, ranging from the quotidian to the monumental, structures social, economic, and political space and history (Moore 1996; Pearson and Richards 1994; Rodning 2010). In this chapter, I will present data from the two largest constructions at Paso de la Amada, Mounds 6 and 7, drawing on the general approach already sketched here. I will look at the transgenerational nature of their constructions, uses, and abandonments—showing how they shaped both residential and public space—both within them and between them. My primary goal in doing so is to show that in picking apart the carcasses of these buildings and by keeping in mind that much of their physical existence is owed to the actions of hundreds of community members spanning many generations, we can perhaps decipher some of their social “sense.” By trying to highlight the idea that the physical remains of Mounds 6 and 7 (as indeed all material remains from the past) are the products of human actions intended for multiple purposes, I hope to show that we might penetrate the misty history of ancient Paso de la Amada. Richard Lesure’s Chapter 6 extends beyond these two buildings to examine the broader implications of “formal” and “informal” activity at the site, tracking subtle shifts in building sequences, site planning, and the use of space in Paso de la Amada’s history. I have organized this chapter as a narrative of the detailed discoveries and interpretations as they were made for a number of reasons.2 First, I want to give a sense of the spaces, features, and structures as they appeared to us during the excavation process. Because the structures were all made of clay derived from the surrounding

terrain, they required us to pay close and constant attention to subtle shapes, textures, and colors. These variables, therefore, become the raw data for interpretation and inference building and cannot be easily tabulated, measured, and analyzed in a standardized archaeological form. Second, I want to show that our interpretations started off as hypotheses that could be tested with each new season of excavation as we uncovered more information from earlier features. The structures and features that emerged from the mounds, in ghostly fashion, were, nevertheless, solid enough for photography, mapping, and physical data collection. As these features—floors, walls, post molds, trash pits, burials, ovens, refuse concentrations on floors, buried offerings, erosion episodes, and so forth— came into view, they connected up in a chronological sequence that we were in the process of discovering and creating. Third, in telling the story beginning with the most recent features in the sequence, I can give the impression that we are moving back in time, generation by generation, exploring the decisions and actions of the builders and residents that grew out of a tapestry of preceding decisions and actions. In other words, the archaeological sequence represents a history of practice, each material instance of which—a floor shape, a post alignment, a human burial—had some discernable impact on the actions and choices of the subsequent generation. In a way, it is like reading one’s ancestors’ letters backward—you have some knowledge of the future outcomes long before the writer did at the time she or he wrote the letter. It is in this sense the archaeological sequences at Mounds 6 and 7 at Paso provide this generation-by-generation historical narrative, thereby giving us a glimpse of the nature of ancient Paso society.

BUILDING MOUND 6 I start with Mound 6 because it has received the most attention since our fi rst work there in early November 1985. When excavations of Mound 6 first began, John Clark and I did not

history in space at paso de la amada

99

15

low ground

12 13

7 11 10

Trench 1

Trench 2 low ground

6

FIGURE 5.1 Map of Paso de la Amada, showing locations of Mounds 6 and 7 and the major excavation units

N 0

100 m

emergent complexity

low ground

contour interval 50 cm

know what to expect other than that we might find stratified sequences of ceramics and other artifacts similar to those discovered in test-pitstyle excavations at Paso and other sites in the region during the preceding two decades (for instance, Ceja Tenorio 1985; Green and Lowe 1967). For the most part these excavations suggested that Early Formative peoples were living permanently in the region and that they were building large structures, but both the nature of their buildings and the range of features present in their settlements were not well known. Basic questions such as “What were the styles of houses used?” were unanswered, if not unasked (for example, Flannery 1976). We decided to excavate Mound 6 because it was one of the largest mounds still remaining at Paso (see Figure 5.1) and our inspection of the ceramics visible on the surface and around the skirt of the mound indicated that it was

100

2

probably entirely Early Formative in age— all the ceramics initially looked Ocós in style, as previously described by Coe (1961) and Ceja Tenorio (1985).3 Ceramic density on and around the mound was not particularly high, unlike the case with some of the later sites we had visited in the estuary closer to the ocean—for example, the later-period Los Alvarez site that had been tested in the 1970s. Even so, the mound looked to us to be relatively intact and so was a promising candidate for testing. I began excavating the plowed surface of the mound and within the first 15 cm came upon what looked like a compact clay floor with fragments of cane-impressed burned daub. Excavation followed the floor remnants exposing them using a 2 × 2 m grid hoping that there was enough material preserved to tell us something about this uppermost floor, and by extension the structure it represented.

This remnant of the uppermost floor (Floor 1) on the top of the mound was well preserved near its center but had been badly disturbed by plowing toward its outer edges, where the floor had been scoured away and pushed down the slope.4 Even so, the following characteristics could be discerned: it was heavily burned in patches, either accidentally or in hearth locations; the hard-packed clay was up to 4 cm thick in spots and looked to be a specially prepared surface different from the surrounding earth; before being covered with construction fi ll, the clay floor had cracked, as is often seen in earth floors in modern houses today; and there were artifacts embedded in the floor material, including small fragments of ceramics, obsidian flakes, and charcoal fragments. The total area of the remaining floor surface was only about 6 × 8 m. Where we excavated through Floor 1 to explore its depth, especially in areas where it was the most disturbed, we came across a layer of looser clay and silt soil about 25 to 30 cm thick that overlaid an earlier floor. Excavation proceeded through Floor 1 and the layer of underlying construction fi ll in order to trace out Floor 2 below it. This layer of construction fill between the two floors contained melonsize lumps of clay mixed with mostly sterile mottled yellowish-brown sands and clay. Excavating Floor 2 proved to be a relatively easy and straightforward task because it was well preserved over its whole extent, protected as it was from plowing by the thick construction layer and floors above it. Using the same 2 × 2 m grid (labeled with letters A through K on its S–N axis and 21to 28 on its E-W axis) we dug down to the top of Floor 2 and recorded all artifacts on its surface and screened the layer of construction fill covering it (beneath Floor 1). As we exposed Floor 2, a crew member noticed a thin, straight, charred orange line along the outer edge of the floor. This pencilthin line of charcoal or organic material extended to a depth of several centimeters and could be traced for almost 7 m. In scraping the floor to

expose this line, which marked the edge of Floor 2, I noticed a dark-stained post mold set back 5 cm (that is, the post mold was visible inside the floor). This discovery created a good deal of excitement because as we followed the line along the edge of the floor five additional large post molds appeared at floor level, just inside the line. A similar line marking the edge of the floor was found on the opposite side of the structure. It too was linked to postholes and could be traced out for several meters. Although Floor 2 was made of compacted clay, we found that it was still hardened and relatively easy to trace over most of its surface. In some patches it was not as well preserved as others, possibly because of rodent disturbance and ancient tree roots. But for the most part these disturbances were rare and confined to fairly small patches. This suggested that both Floors 1 and 2 had not been very near the original surface of the mound and that before recent agricultural activities there may have been several later floors and construction fill layers that had been plowed away in the intervening years. We will look at the implications of this later. We fully exposed Floor 2 and all of the postholes associated with it, labeling it Structure 2. It was apsidal in outline, 17.5 m long and 9 m wide, with a total floor area of 122 m2 (Blake 1991:34). One important discovery on the floor was the center line of three post molds that must have held large posts supporting the main roof beams. This information about the shape and the layout of the structure associated with Floor 2 provides some important clues about the creation of social space at Paso. The structure was made of the same types of materials (clay floors, wood posts and beams, presumably thatched roof, and possibly wattle-and- daub walls but more likely canes walls) as many domestic buildings in the Soconusco today. It was bigger than most houses we see today, and this may have been a function of the size of the social group it held. The northeast side of the structure was facing onto what we think may be a

history in space at paso de la amada

101

plaza area, or at least a large open space flanked on the northwest side by Mound 7 and more or less open toward the northeast side (Lesure, Chapter 6 in this book; Clark 2004a). The structure faced toward the mountain range in the near distance, the most prominent member of which is the Tacaná volcano. The back side of the mound faced toward the Pacific coast, 7.6 km away, and opened onto an estuary channel that may have been an open waterway during the site’s occupation—perhaps giving canoe access to the sloughs and estuary channels that snake along the coast. Perhaps just as important as its size was the location of Structure 2 atop one of the highest mounds at the site. Furthermore, this mound was the site of a specific history of occupation: the people living in and using Structure 2 were doing so on a spot with a known heritage of earlier constructions and it is likely they remembered specific earlier buildings. The memory of these structures is represented by the particular details of the floor plan and layout. As we will see, Structure 2 had a floor plan that was derived from its immediate predecessor: Structure 3. After exposing the floor and other features of Structure 2, we continued digging down through the floor material, encountering another layer of construction fill that had been brought in to provide the foundation for Structure 2’s floor. This layer of construction fill, approximately 15 cm thick, completely covered Floor 3. Although not as well prepared and preserved as Floor 2, Floor 3 was also made of clay. Perhaps it had been exposed to the elements for a prolonged period, or possibly it had been destroyed during multiple rebuilding episodes. But the most important thing about Floor 3, and by extension Structure 3, was that it provided the template—both in shape and location—for Structure 2. It was 12.6 m long, and its first building stage was 6.3 m wide. At some point it was expanded in length by about 1.5 m and then in width by another 0.5 m. It increased from 71 m2 to 87 m2 in floor area. Later it was widened by another 1.0 m, giving

102

emergent complexity

a  final floor area of 98 m2 (Blake 1991:36). Because it was rebuilt without raising the platform, it was somewhat more difficult to trace out the floor than was the case for Structure 2. Also, because the construction fi ll that covered Floor 3 was only about 15 cm thick, many of the postholes from Structure 2 penetrated down through Floor 3. Even so, we could easily identify and note all of the posts from Structure 2 and were left with a reasonably complete record of Floor 3’s construction sequence. One notable aspect of Floor 3 was that as it was expanded, the central line of roof beam supports was shifted to correspond to the new center of the widened structure and more closely aligned with the locations of the central posts in Structure 2 that followed. After exposing Structure 3 and photographing it, we excavated a 2 m × 2 m test pit through the central floor area (H25) and three similar test pits on the edge of the mound (A25, E28, and K21) in the short time remaining in the season. The stratigraphy in the central test pit showed that there was still about 2.5 m of deposit left to plumb and gave us a window into the structure’s past history and our future work (Blake 1991). We returned to Paso in 1990 to continue excavating Mound 6 and testing other parts of the site. The central test pit was a key to our strategy for opening up more of the mound because in 1985 we had found a series of welldefined floors visible in the profi le of test pit H25. The uppermost of these floors was a clearly visible horizontal layer of clay burned to a brick orange color. It sat approximately 90 cm below Floor 3, so our strategy for the 1990 season was to open up the mound and trace out this floor (labeled Floor 4) to see if it was similar to Floors 2 and 3. We also intended to open up the two or three visible floors below it, including the lowermost floor that sat on the basal layer of sterile sand underlying all the features at the site. In excavating test pit H25 that revealed Floor 4, we observed a thick layer of mottled

clay consisting of unstratified material and containing very few artifacts, especially pottery. This deposit was completely undisturbed because it had been protected by the later floors and fi ll layers above it. We were surprised that it did not look like the construction fi ll deposits that had been used to cover Floor 3 before Structure 2 was built. Instead, it looked like lumpy, basket-sized chunks of clay and earth, irregularly shaped and varying widely in color from dark brown to light yellowish. Some parts were fairly compacted and others less so. We then began to dig down below Floor 3 in some of the units adjacent to the central test pit (H25), which we cleared out upon our return to the site. Just as the profile showed, there was no visible stratigraphy as we excavated the adjoining units—yet we screened the deposits to collect a representative sample of artifacts contained within. The floor layer, burned in patches, was very easy to excavate. Our trowels quickly found the interface between it and the overlying construction fill. Of course, we expected that this new Floor (labeled Floor 4) was going to be similar to Floors 2 and 3 above, and we were fairly certain that the material we were digging through to uncover Floor 4 was a mass of platform construction fi ll used to raise the mound on which Structure 3 was built. We had screened a sample from several units beside the central test pit and concluded that this was sufficient to give us an indication of the contents of the construction fi ll. And since it was construction fi ll and not primary deposits, we decided to excavate it more rapidly by foregoing screening in order to expose Floor 4. As we were excavating through the units adjacent to H25, we came across a linear “block” of compacted homogeneous clay. The top of this block was about 50 cm above Floor 4 and completely unexpected. As we traced out the top of the block within the excavation units where it first appeared, we saw that it extended horizontally into the adjacent units. We also saw that it extended vertically downward, merging with Floor 4. We

soon discovered that it was a low clay wall that ran along the southwestern side of the structure. Teams of excavators followed the wall— exposing both it and Floor 4 as they went. The wall was about 10 m long and came to an abrupt end on the floor in the northwestern end of the structure. On the southeastern end, it made a right-angle turn jogging out to the edge of the platform, and then curved along the edge of the platform, making a near perfect half circle forming the southeast end of the structure. The wall extended partway across the floor and then stopped. At that point it was interrupted, forming a doorway or entrance to the side of the structure facing the “plaza”—that is, the side facing the mountains. Excavating the floor to the northeast, we soon reconnected with the low wall, another perfect half circle defining that end of the house (see Figure 5.2). Structure 4 had the same orientation and shape as the structures that followed but was much larger, better constructed, and more elaborate (see Figure 5.3). It was 22.5 m long and 10 m wide. Flanking each side of the structure was a long step made out of the same clay as the floor, the platform, and the wall, in effect making the width of the structure even greater than that defined by the platform alone (12.5 m). The platform was evenly made and about 0.7 m thick over its entire surface. When the floor was cleared of the construction fill that had covered and preserved it so well, we could easily see two lines of clearly delineated post molds spaced at 2.5 to 3 m intervals running parallel to one another 4 m apart. Two more parallel lines of post molds were discovered at the tops of the steps, spaced 3 m apart from the center lines. This was a very different pattern from that observed in Structures 2 and 3 and, as we later discovered, from the post pattern used for the structures that preceded Structure 4. One of the most interesting differences was that there were none of the end posts circling the apsidal ends of the structure. There were other details of construction and features that distinguished Structure 4 from

history in space at paso de la amada

103

Hearth

Clay walls Postholes

Step Porch

Burned patch

Porch

Test pit Test pit

N FIGURE 5.2 Plan of Structure 4 showing the layout of the clay walls, postholes, hearths, porches, and steps

0

Hearth

2

Burned patches

4 m 7E

FIGURE 5.3 Photo of Structure 4 fully exposed and the 1993 excavations beginning to uncover the construction fi ll overlying Floor 5

12N

its successors. There were two clearly defined hearth areas at each end of the building—and these were in addition to several other burned patches of floor. Also, there were two circular arrangements of small stake holes about 1.5 to 2.5 m in diameter and located on the inside floor adjacent to each end of the long straight clay walls. A feature similar to these was discovered on the clay floor of an oval structure found deeply buried in the Late Archaic period shell mound at Tlacuachero (Voorhies 2004:53– 60). This structure at Tlacuachero dates to about 3350 cal. b.c., so it pre-dates Structure 4 by about 1,700 years. Two other features were discovered in the Floor 4 excavations. One was an infant burial near or under the hearth in the northwest end of the building. It was not well preserved, partly because the infant appears to have been a newborn and partly because it was perhaps damaged by the ongoing use of the hearth. The practice of interring infant burials in the floors of houses was common in Mesoamerica through to the Postclassic Period. I found several infant burials (placed in used ollas) in Postclassic Maya houses at the Canjasté site on the Lagartero River 140 km northeast of Paso de la Amada (Blake 2010). Other features include possible offerings of partial or complete birds in the floor. The birds, identified by David Steadman (Steadman, Tellkamp, and Wake 2003), include a complete green heron (Butorides virescens) from Level 9 Lot 2; a cluster of bones from a crested caracara (Polyborus plancus) from Level 9 Lots 26, and multiple bones of an Inca dove (Scardafella inca) from Level 9 Lot 1. Even though Structure 4 was the largest floor in the house sequence, it was completely covered by the construction fi ll for Structure 3. The actual building that constituted Structure 3 (as indicated by its posthole pattern) was relatively small, but its platform was huge, entirely enveloping Structure 4. The size of the structure more or less overwhelmed our resources for the 1990 season, and we were not able to do much more than expose and map it and draw all the profi les. More excavation had to wait for the next season.

We returned to Mound 6 in 1993 to excavate beneath Structure 4 and try to uncover the earlier floors initially seen in the profile of the original test pit H25. We knew that there were at least one and possibly two floors some 0.7 m beneath the top of Floor 4. So in 1993 we uncovered the previous season’s backfi ll once more and prepared to excavate beneath Structure 4. But before doing so we collected a 100 percent sample of the entire floor to a depth of about 2 cm. We collected the floor samples in a 0.5 m grid so that we could recover and plot the microartifacts (ceramics, lithics, bone, charcoal, and other debris) embedded in the earthen floor surface. We also analyzed the chemistry of the soil in each grid unit and used this information to plot changing uses of the structure (Blake, et al. 2006). Once this sampling was complete, we began the process of excavating out from the central test pit just as we did to expose Floor 4. This meant that we had to excavate the clay platform that comprised the main structural surface. It turned out to be made of the same irregular lumps of different colored clays as the platform underlying Structure 3. We excavated down through it until we found the well-preserved clay floor (Floor 5) that demarcated Structure 5, and then extended the excavation out in all directions until we reached the edge of the floor. In this case, however, we decided to leave the clay walls in place, thus every part of the floor except that which underlies Structure 4’s clay wall was exposed (Figure 5.3). The Structure 4 platform exactly covered the southeast end of Floor 5, neatly overlaying a series of post molds that inscribed the rounded end of the structure. Not so at the opposite end of Floor 5. There the Structure 4 platform extended an additional 2.5 m beyond the limit of Floor 5. In other words, while Structure 5 formed a template for the platform that followed it, the builders did not follow the plan exactly since they decided to make the new platform covering Floor 5 an additional 2.5 m longer toward the northwest. This construction plan had the effect of completely preserving the

history in space at paso de la amada

105

Limit of clay floor

Burned patches

Clay wall

12N Test pit Test pit

Postholes

N FIGURE 5.4 Plan of Structure 5 showing edge of the clay floor, postholes, burned patches, and the small clay wall

0

Burned patches

2

floor and posthole pattern of Structure 5. We were able to trace and excavate all of the postholes around the periphery of the floor and a line of five posts running down the center line of the building (in exactly the same style as Structures 2 and 3, but differing from the two parallel lines of central posts in Structure 4 (See Figure 5.4). Two additional construction details linked the history of Structures 5 and 4. One was the discovery of a doorway in the exact location as the presumed doorway for Structure 4. The clay floor had no postholes in this 6 m interval along the northwest side of the structure facing the plaza, and the clay floor deposit, 3 to 6 cm thick, extended out about 0.5 m beyond the expected floor line suggested by the line of postholes on either side. The second major detail was a 2.8 m

106

emergent complexity

4 m 7E

long section of clay wall, 0.16 m high, aligned almost exactly beneath where a clay “step” was found in alignment with the clay walls in Structure 4. This hints at the concept of making clay walls in the earlier construction, but for some reason was not carried out to any greater extent. As we had done with Structure 4, we gridded the entire floor using a 50 cm grid and scraped off the top 1 to 2 cm of clay floor to carry out a microartifact and soil chemistry analysis. The analyses showed that the occupants of the building performed the same array of general activities on the floor but in somewhat different locations (Blake et al. 2006). There were, then, both activity and architectural similarities between the two structures—but differences as well. One of the key differences was that Structure 5 was built on the level of the sur-

rounding ground surface while Structure 4 was raised 0.7 m on a platform of imported clay fill. In terms of architectural history, the two structures were symbolically connected by shape, orientation, and even by size (although Structure 4 was longer than Structure 5). When we excavated the thirty-five postholes that we discovered on Floor 5, we found that almost all of them had been filled with the same lumpy clay fill that was used to construct the Structure 4 platform. This suggests that the builders carefully salvaged all the old house posts and may even have reused some of them for the new structure. But the builders of Structure 4 innovated by creating a new distinction: the investment of more labor to build an even more prominent structure than before. This new prominence would be apparent both to people who remembered the earlier building and those who were familiar with its internal and external features. But also, the change would have been apparent from a distance. It would have been possible to see Structure 4 from much farther away since it was built on a slightly higher platform. People would have recognized that it had the same general shape (especially if it preserved the same roof shape and materials), but they would have also noted that it had a completely new architectural innovation—low clay walls surrounding its ends. When Floor 5 was completely exposed, we could see that its well-formed and compacted clay was built on the surrounding sandy ground surface. It was mainly visible because the sand surface outside the structure was exposed in our excavation around the entire floor surface. Test pit H25 suggested, however, that there might be an ephemeral floor beneath Floor 5. We therefore excavated through Floor 5 to see what kinds of features might be preserved under it. Was there yet an earlier structure to be found? This was answered quickly by the discovery of many post molds on the sandy floor immediately below Floor 5. No prepared clay surface existed, nor were there clear-cut features as in the later structures. But we could see lines of post molds that were similar to, and almost in the exact

positions of, the two parallel lines of posts that supported the roof of Structure 4 (not Structure 5 directly on top of it). There were only a few (four) peripheral posts that may have marked the outer edge of Structure 6’s floor, and these were actually beneath the clay of Floor 5, so we know they were not associated with that later building. In addition to the two parallel lines of larger posts about 5 m apart, there was a central line of five or six smaller posts between these and spanning about 15 m of the structure’s length. Altogether, there was not really enough of a clearly defined or prepared floor to calculate the floor area of the structure, but at a very minimum it would have been large enough to encompass all the post molds at this level. We estimate it to have been somewhat smaller than the 165 m2 of Floor 5. When excavating to the level of Floor 5 in test pit H25, we discovered two potential offerings that could have been associated either with Structure 5 or 6 (because both floors were so close together, it is difficult to determine which of the two structures the objects are associated with)(Blake 1991). One was a deer scapula pendant, carved into an A-shape and covered with red hematite. The other was a greenstone celt. Both of these objects, while they may have been offerings, may also have simply been stored in a cache pit in the floor either for safekeeping or for later use. During this season of excavation we also probed the aprons of Mound 6 by digging two trenches and some test pits both to the northeast and the southwest—that is, perpendicular to the long axis of the structure. Our goal in doing this was to determine if there were outbuildings and refuse deposits associated with the different structures that might give us more clues about their functions. We knew the deposits were rich in materials because in 1985 we had excavated three 2 × 2 m test pits in these parts of the mound and had discovered refuse layers, pits fi lled with trash, and an oven pit. The expansion of these periphery explorations was valuable because they showed that there were surfaces sloping down and away from the

history in space at paso de la amada

107

Estimated original mound height

Str 1 Floor

Clay wall Step

0

5m

Str 2 Floor Str 3 Floor

Str 4 Floor Porch

Porch

Str 5 Floor

Str 6 Floor

Step

Str 4 Platform

Sterile sand

FIGURE 5.5 Mound 6 in profi le (across its short axis) showing the superimposed layers of fi ll that built up the mound and supported the successive structures

various superimposed structures, indicating that there were later structures on top of Floor 1 that had been swept away by annual plowing, beginning perhaps as early as the late 1950s. They also showed that larger artifacts and debris associated with the activities inside the structures had been regularly cleaned off the floors and tossed onto the edge of the mound within a few meters of the structure. In one of these peripheral trenches we discovered a pit containing an adult female and an infant.5 In 1995 we returned to the Mound for one last season to excavate more of its periphery. We extended the 1993 trenches perpendicularly to the mound’s long axis and also excavated two long trenches extending out to the southeast and northwest (aligned with the mound’s long axis). John Clark excavated the largest of these trenches extending 215 m north from Mound 6 toward Mound 7, almost reaching it (Figure 5.1). We opened up a large quadrant of the northeastern apron of the mound, exposing the layers of fill and their surfaces, helping us estimate the total size and volume of the platforms that were built to support the successive structures. Some of these data have been presented previously (Blake et al. 2006; Clark 2004a:56, figure 2.4), so here I will focus on characteristics that help us understand the way these builders were creating social spaces at the site. These peripheral excavations also helped us get a glimpse of the later occupations of the mound— the structures of which were destroyed by plowing. The mound appears to have been occupied during the Ocós phase, and much of the midden debris off the side of the mound was deposited during this phase or spread out-

108

emergent complexity

ward as the top layers were progressively lowered by plowing.

THE MOUND 6 SEQUENCE— BOTTOM TO TOP The successive building and expansion of each platform and construction of new structures atop them can be viewed as the ongoing renewal of or investment in an idea about the social group that commissioned and used the structure (see Figure 5.5). John Clark and I have argued elsewhere that the structures on  this spot most likely served as houses— domestic spaces—for the community’s most prominent household or sequence of lineage heads (Blake et al. 2006; Clark 1994; Lesure and Blake 2002). Others have suggested that the Mound 6 structures may have functioned primarily as men’s houses (Marcus and Flannery 1996:90– 91). As Lesure points out in Chapter 6, even if these structures were primarily residential and people performed a range of informal daily activities in them, residents could also have hosted public and formal ritual events. As domestic spaces the structures were also clearly designed to have a prominent public role. We might expect that the smaller houses of less prominent kin groups at the site would have played a less public role— or that their public roles would have been socially restricted to a much smaller segment of the community. This was the case at Mound 12 and some of the smaller mounds, also located within the site’s ceremonial core, where the sequence of residences shows less concern with public or formal functions (Lesure 1995, 1997).

In the case of the structures represented at Mound 6, the public roles of the household members may have been regularly displayed in their residential settings. Frequent gatherings may have taken place there, and at times perhaps events affecting the whole community could have been staged in and around the residence. At such times the social functions of the buildings would have been highly public—hence the need for increased visibility (Clark 2004a). One of the central implications of the historical continuity of the sequence of structures creating Mound 6 is that they represent a social memory of the structure’s inhabitants across many generations. We estimate that the Mound occupation spans some 200 to 300 years from the early Locona phase around 1700 cal. b.c. to the late Ocós phase around 1400 cal. b.c. If the buildings and their platforms were refurbished, rebuilt, and expanded at least every generation, then the sequence could represent ten generations of lineage members (with an average generation span being from twenty to thirty years). Although we were able to find the remains of six structure floors, the layers of platform construction on the apron of the mound suggest that there were at least three others that sat on top of Floor 1 (Figure 5.5). There is a reasonable fit between the estimated number of structures in the sequence and the estimated number of generations that spanned Mound 6’s occupation. There are some features, such as earth and clay brick ovens, and trash pits (Blake 1991) that were constructed during the last occupations of the mound, but unfortunately these cannot be connected with specific house floors or building structures such as we are able to see in the sequence of Structures 1 to 6. The correspondence between significant domestic functions and large-scale construction over many generations, as well as the public prominence of the Mound 6 structures, provides some basis for the interpretation that the social group living there held high social and political status. If this interpretation is correct, then the Mound 6 sequence represents a crossgenerational transfer of legitimate rights to

place and symbols of prestige embodied in the building that housed the residents. They were distinguished by the large size of the structures, their shape, building materials, and prominent location in a central part of the community where many other large mounds were located. This set of distinctions differentiates the social group from the much larger number of smaller, less prominent, and lower-status households at the site. An alternate interpretation is that the social group that built and occupied the successive structures was not a high-status kin group or lineage, but instead a cross- cutting association or sodality such as a men’s society (Marcus and Flannery 1996). If so, then the group participated in a range of domestic activities remarkably similar to those of the members of smaller households, the remains of which we have also excavated at the site. At some point more evidence may be found to support this alternate hypothesis, but for the time being, I think the interpretation that the structures belonged to a kin group of preeminent rank in the settlement remains the best-supported one. Furthermore, if this interpretation continues to hold, then we may also be able to say that because Paso de la Amada appears to be the largest and most complex site in the region for this time period, then the household represented at Mound 6 could also have had a preeminent social standing, and possibly exerted some measure of political influence, extending over many neighboring settlements as well. In other words, Mound 6 is a reasonable candidate for the residence of hereditary regional political leaders—leaders who may have traced their descent to the founding builders of the first house at the base of the mound, Structure 6. We might call Mound 6 the ancestral home of the Paso de la Amada chiefs.

BUILDING MOUND 7 In 1990, after we had almost completely exposed Structure 4, we decided to test Mound 7 with three test pits to see if it was a candidate for another sequence of superimposed structure

history in space at paso de la amada

109

floors and associated features and refuse deposits. As such, it would have been ideal to compare with Mound 6. These test pits were placed across one-half of the short axis of the mound: one in the center, one halfway toward the edge, and one just off the mound. This plan was based on the strategy that had worked so well in Mound 6. In effect, we were trying to duplicate the success of Test Pit H25 that we dug in 1985 and that showed the entire sequence of superimposed floors. We had no such luck in Mound 7. The central test pit showed that there was a layer of dark gray soil 1 m thick interspersed with irregular, gently undulating darker surfaces that were neither clear-cut nor well defined. Then, just over 1 m below the surface, we came upon a deposit of finely stratified sandy layers that were interspersed with fine clay layers. None of these deposits contained the density of artifacts that were found in and around Mound 6 or many of the other mounds at the site. These finely laminated sands and clays continued down another meter and then ended rather abruptly on a hardened floor surface. But, there was no burning on the floor, almost no artifacts, and no other features. Below this was the same soft sandy deposit that we had discovered at the bottom of all our test pits both on and off the mounds. The second test pit, halfway between the center and the edge of the mound, was completely different in that we immediately contacted the type of mixed clay deposit that characterized the construction fill that that we had just finished excavating by the wheelbarrow load under Structure 3 in Mound 6. Again, very few artifacts were present in the fi ll. At the time we did not really know what to make of this pattern of deposition in the test pits. But there were enough indications that we could find level floor-like surfaces in the center of the mound and that with more horizontal exposure we might be able to find the structural remains of houses inside Mound 7. During our 1993 season we were so preoccupied with uncovering Structures 4, 5, and 6 in Mound 6 that we did not have time to return

110

emergent complexity

to Mound 7. Finally, during the 1995 season, we planned to excavate larger sections of Mound 7 in order to open up possible floor surfaces and look for posthole patterns and other features typically found in the house mounds at the site. This excavation was carried out as Warren Hill’s doctoral dissertation research project (Hill 1999).6 We had hypothesized that because Barra-phase sherds had been found at the base of Mound 7 in 1990, we might be able to uncover the well-preserved remains of a Barraphase house or two. We started the excavation by locating the 1990 test pits and expanding them to create a trench across the midsection of the mound, reasoning that this would allow us to spot the floors, if they existed, and then follow them out in all directions (Hill 1999:80). We ran into problems with this strategy almost right away. As we excavated across the mound we discovered that the dark earth in the middle of the mound was unevenly layered and somewhat mixed. It seemed to have a high humus content. To either side of the mound this humus layer was very thin and we found light and dark- colored lumps of clay that were deposited in a jumbled fashion—reminiscent of the basket loads of construction fi ll that underlay Structure 3 in Mound 6 (see Figure 5.6). The pattern that emerged as we bisected the mound was that there was some kind of sunken internal feature that had been weathered and that possibly had led to the pooling of water. Underneath the dark organic surface deposits in the center of the mound were more of the very fine sandy- clay laminations that appeared to have washed down into the center of the structure—the same layers that we had seen five years earlier. These laminated layers of sand and clay, each barely a millimeter in thickness, were identical to the layers we had seen in deposits on the apron of Mound 6. They were layers that had been exposed to rain and, as in Mound 6, had washed downslope. It slowly dawned on us that in order for there to be slope wash forming in the middle of Mound 7, it must have consisted of two parallel mounds that were unlikely to have been part of

0

2m

light gray clay

yellow-gray clay humic layers

laminated silt layers

yellow-brown clay

humic layers

Ballcourt Alley

Bench Bench brown-gray wood-grain silt

0

1m

yellow-brown silt

FIGURE 5.6 Mound 7 in profi le (northwest face of Trench 1) showing the construction sequence, erosion layers, mound features, benches, and alley

a single house platform. The horizontal layers of sandy clay in the center of the mound near the level of its base suggested a large flat openair surface between two adjacent clay mounds. We were left with only one realistic hypothesis that made sense of what we were observing in our cross-section of Mound 7: John Clark’s hypothesis that Mound 7 was a ballcourt (Hill 1999:80). We continued excavating the 50 m long cross-sectional trench and preparing the trench walls for close inspection and profi ling. After carefully scraping the trench walls we could clearly see the outlines of the two mounds that flanked the central alleyway (Figure 5.6). From that point on, our excavation strategy was to dig a series of shorter trenches to define the extent of the ballcourt and attempt to unravel its construction history. Briefly, we excavated twenty-two short trenches (ranging from 1 × 2 m to 1 × 4 m) and four longer trenches (ranging from 1 × 25 to 1 × 35 m) to define the width of the alley, the dimensions of the two low clay benches that sloped downward from the parallel mounds toward the alley floor, the nature of the alley floor, and the shapes of the ends of the long mounds. We also excavated

thirteen “control units” to obtain screened samples of sherds and other artifacts in each of the known architectural components of the mound. In all, we excavated almost 10 percent of the ballcourt area, and we are confident that the basic size, shape, and construction sequence is well understood (Hill 1999). (See Figure 5.7.) One important discovery is that the ballcourt was built in at least two main stages. The initial parallel mounds were made of clay that was likely brought in as basket loads of fill to build up structures that could contain the rubber ball in play much like the later-period ballcourts found throughout Mesoamerica. The dimensions of these initial mounds are estimated to be 74.3 × 4.3 m (south mound) and 77.5 × 5.4 m (north mound) and would have stood at a height of about 3 m above the surrounding natural ground surface (before the same kind of plowing that destroyed the top layers of Mound 6 and most other mounds at the site). These two mounds were separated by a 6.8 m wide playing field or alley that was flanked by hard-packed clay benches, each 2.5 m wide and averaging 0.36 m high. Together, the

history in space at paso de la amada

111

End Zone

Test Pits and Trenches

N

Trench 1 (see fig. 5.6)

Alley Bench Initial Mound Expansion Mound End Zone

FIGURE 5.7 Plan view of Mound 7 showing the excavation units, benches, alley, and lateral mounds

0

alley and its flanking benches provide about 12 m of separation between the two mounds. The total width of this initial ballcourt, including the two mounds, benches, and alley, was 21.5 m—the same as the length of Structure 4 in Mound 6. One of the striking aspects of this initial court is that the architectural elements— parallel mounds and low sloping benches flanking a long narrow alleyway— are the essential characteristics of all later Mesoamerican ballcourts (Figure 5.7). In later times, ballcourts were often faced with stone but, even so, the overall design, if not the exact measures and proportions, remained the same. This means that by the beginning of the Early Formative period this key architectural concept in ballcourt construction was already in place. There is evidence that the ballcourt was refurbished at least once. It appears that similar loads of clay construction fill were added on

112

emergent complexity

25

50 m

to the exteriors of the parallel mounds, widening and lengthening each one. The southeast mound was widened an additional 6.4 m from its initial width of 4.3 m, giving it a total width of 10.7 m. The northwest mound was widened an additional 3.2 m, giving it a new width of 8.6 m. In addition, both mounds were lengthened with additional fill on each end: the southeast mound by a total of 8.3 m and the northwest mound by a more modest 2.1 m. The longer of the two mounds, the southeast mound, was now 82.5 m long, while the overall width of the ballcourt had expanded to 31.1 m. The tops of the mounds may have been raised as well, but there is no remaining evidence of this, probably because of the same plowing that took place over the whole site. We did not find any evidence of specialized end zone construction, although if low earthen markers or perishable materials were used, there is little chance that such evidence will ever be found.7

It can be inferred that the ballcourt was first built during the early Locona phase because the construction fi ll from the initial set of parallel mounds contains only sherds that date to either the Early Locona or Barra phases. There were none of the later Locona- or Ocós-phase ceramics commonly found in association with other mounds close by Mound 7. We found a postballcourt deposit of later ceramics at the foot of one of the ballcourt mounds (on the outer slope of the mound), as though later people may have built residential structures on top of the mound after the ballcourt had been fi lled in and it had been reduced to a raised single mound surface that could serve as an occupation. At any rate, our ceramic analysis and radiocarbon dating of a tiny carbon sample from within one of the stratified sand- clay laminated layers above the alley floor indicates that the ballcourt began to fill in sometime between 1659 cal. b.c. and 1388 cal. b.c. (Beta 82234, 95.4 percent probability, OxCal 4.0), and was probably badly eroded by the end of the Locona phase or early in the Ocós phase.8 The ballcourt’s long axis is aligned toward the distant mountains and therefore is parallel to the short axis of the structures that were built at Mound 6. It is clear that the ballcourt flanks a large open plaza space on the northwestern edge of the site while Mound 6, with its long axis aligned perpendicularly to the ballcourt, forms the southwestern edge of the plaza. In between the two are few mounds, although there are some floors of what appear to be residential structures buried deeply within the plaza; these may have been present when the early structures in Mound 6 and the ballcourt were in use. John Clark (2004a) has pointed out that Gareth Lowe first recognized this arrangement in the early 1970s when Jorge F. Ceja Tenorio was working at the site— and that this was an interpretation that we readily dismissed. Clark’s subsequent analyses of the spatial arrangements of the site, and his estimation of its enormous scale, have helped us appreciate Gareth Lowe’s insights from more than thirty years ago with renewed respect.

As a public structure, the ballcourt is unequivocally restricted in functions to nondomestic activities. There are few features or artifacts associated with the ballcourt that indicate that any other activity but ball games took place there. Although we have an admittedly small excavated sample from Mound 7, we did not find any ritual features or offerings, which are common in later-period ballcourts in Mesoamerica (Fox 1996). There was no evidence of Locona- or Ocós-phase feasting or other public activities associated with the use of the court, although these may have been cleaned up afterward (Hill and Clark 2001). Associated feasting activities—including both cooking and eating—must have taken place in residences or other spaces nearby, and of course Mound 6 is a prime candidate because of its close spatial and chronological association with the ballcourt. The construction history of the ballcourt is rather simple compared with the history of Mound 6. The site of the ballcourt may have been a location for the playing of the game on an open field long before the formal court was built. I suggest this only because there was a well- compacted surface underlying the alleyway floor and it had a very small number of what appear to be Barra-phase pottery sherds associated with it. A radiocarbon sample from that deposit yielded a good Barra- or Early Locona-phase association: 1918 cal. b.c. to 1611 cal. b.c. (Beta- 82233, 95.4 percent probability, OxCal 4.0). I have no other evidence for this hypothesis, and it cannot be tested without a  much more extensive excavation of the ballcourt area. However, I do think that the ballcourt was built very early in the Locona phase, at about the same time as Structure 4, both because of the ceramics and some construction similarities between the two. The first ballcourt required a substantial investment in labor. The people who built it had a clear idea of what was required to make a ballcourt because it was built all at once, not piecemeal nor haphazardly. Ironically, even though this is, as far as I know, still the earliest Mesoamerican ballcourt, it must have had

history in space at paso de la amada

113

precursors in neighboring sites and there are very likely remnants of these earlier courts still to be discovered. That it was built with the two alley-flanking benches is another indication that the template was in place for such a construction. In a sense, it is as though league rules required a specific court form and this allowed players to play the same form of ballgame over a wide network of interacting communities. What is surprising is that the essential characteristics of this form continued for another 3,000 years until the Spanish Conquest. The reason this is surprising is that the Paso ballcourt was abandoned before the site was abandoned— and as far as we know no new court was built at the site. But obviously the game and the notion of this par ticular court style continued throughout Mesoamerican history. It is unlikely that the Paso ballcourt designers and players were the ones who passed this on to posterity, since the court was in decline or even abandoned by the Ocós phase and completely unused during the Cherla phase. Other much more prominent courts, in settlements that had longer and more significant influence on the future than did Paso, must instead have seeded the widespread and continued knowledge of the game and its court styles. The one rebuilding of the court required a substantial reinvestment of labor. This effort must have been considered worthwhile by the political movers and shakers of ancient Paso. They made the structure the largest single building in the community’s history and thereby transformed the settlement into one that could host large-scale regional and perhaps interregional gatherings that demonstrated the social and political competence of the community’s leaders.9 But the lack of continued rebuilding and refurbishing of the ballcourt indicates that it may have had a somewhat tenuous hold on the social life and political aspirations of the community leaders. The decline of the court may have corresponded with the decline in the community’s fortunes, or they may simply have lost their franchise. I will compare the history of the Mound 7 ballcourt with the his-

114

emergent complexity

tory of the Mound 6 structures to suggest there are some important distinctions between the two realms and that the nature of political stability might have been tenuous when taken out of established kin-based levels or structures.

BUILDING HISTORY WITH THE HISTORY OF BUILDINGS The earliest constructions in Mounds 6 and 7 were roughly contemporaneous. But as we have seen, they were not the same. When the first two of many subsequent structures were built on the spot where Mound 6 eventually rose, there was not much to distinguish them from other such buildings (whether dwellings or public buildings) at the site. We do not know the exact size of Structure 6, but we can say that Structure 5—built on the flat ground surface, covering Structure 6’s modest floor—was well crafted. The thick, smooth layer of clay that formed the floor was applied right up to the posts that defi ned the wall of the building. The front doorway, facing the mountains, had a bit of extra clay floor extending another 0.5 m toward the patio. There is a hint of a similar but smaller doorway on the back side of the structure. These two doorways suggest that the building was walled—probably with cane and not daub, since no wall daub was found in association with this building. The wall posts and the central line of five posts running down the long axis of the building were all about the same size. The burned patches of floor found at both ends of the structure are interpreted as hearths, and the microrefuse in and around the hearths, especially in the northwest end, indicates that domestic activities took place there (Blake et al. 2006:202). The low clay wall or bench on the structure’s floor has an interesting link with both the earlier building and the building that followed it, thus providing a view of architectural history. The clay wall sits directly over one of the two parallel lines of posts that would have supported the roof of Structure 6. Even more telling is the observation that it directly under-

lies a section of the lowest “bench” in the overlying Structure 4. Even though Structures 6 and 4 had parallel lines of center posts and intervening Structure 5 did not (it had a single center line of posts), the benchlike feature links all three structures. This, combined with the size, shape, orientation, location of hearths and microrefuse on the floor, demonstrates that these three structures were part of a continuous architectural idea. I suggest that part of this idea was the connection to the ballcourt. Even though these structures, buried deeply within the base of Mound 6, were roofed areas that we have interpreted to be residences, they do show a symbolic “reflection” of the ballcourt. This can be seen most clearly in the low clay walls in Structure 4. Structure 4’s floor plan—which, as we have seen, draws on the historical connections in the two earlier structures— also mimics the layout of the ballcourt. The two parallel clay walls suggest the “alley” of the ballcourt and the two circular clay walls suggest the “end zones.” The width of this “alley” is 4.5 m between the inside faces of the clay walls and somewhat less (4 m) if measured from the two lines of center posts. It is not likely that this series of structures was thought of as a ballcourt, but it is possible that there were resonances between the two locations. For example, as mentioned earlier, we do not have any idea what the end zones of the ballcourt at Mound 7 looked like. But on the basis of this hypothesized architectural linkage between the internal layout of Structure 4 (and the two earlier structures), it is quite possible that the ballcourt was conceived of as having apsidal end zones just as all these structures in Mound 6 had apsidal end zones. As Warren Hill and John Clark have argued, the social group living in and responsible for the earliest structures in Mound 6 may have been the primary sponsors of the ballcourt’s initial construction (Hill and Clark 2001). They may have represented that sponsorship in the layout of their house—the “ballcourt” house. Clark (2004a:61) now thinks it unlikely that

the residents of Mound 6— and in par tic u lar the “aggrandizer” who may have organized the building of Structure 4—was also responsible for organizing the building of the ballcourt. I agree, yet think it likely that the social group who lived at Mound 6 must have been among the backers and organizers of the ballcourt project. Their interest (and perhaps presumption) in making the symbolic architectural connection between the two structures may have derived from the desire to draw attention to their role in organizing and possibly financing the labor necessary for its construction. I suggest that the architectural similarities among the early structures at Mound 6 were the result of intentional efforts to build a history into the structure of the buildings and then materially show that it extended across the space between the two mounds. The resonance that we can see in these two sets of structures, dim though it may be, is possibly the linking of public and private space as well as the linking of private and public time (history). If the logic of this argument, founded on the particularistic construction details of two separate mounds built and used more than three-and-a-half thousand years ago has any validity to it, then it is likely that larger-scale linkages between time and space, architecture and history, and notions of public and private practice may also be found. In this respect, John Clark’s probing of the sacred geometry of Paso de la Amada and other ancient settlements in the Americas is likely to provide fertile ground for hypothesis building and testing (Clark 2004a, 2004b; Clark, Gibson, and Zeidler 2010). Much more work will be needed to test the implications of these ideas, but his efforts do show that we will need to expand the dimensions of our engagement with the archaeological record of these settlements by at least an order or two of magnitude if we are to imagine new ways of  perceiving these relationships. The Early Formative archaeological record in the Mazatán zone of the Soconusco cannot

history in space at paso de la amada

115

be characterized by Paso de la Amada and certainly not by two mounds, albeit relatively large ones, at this one site. However, the main reason they capture our interest has to do precisely with their precocity and uniqueness. We cannot yet compare the ballcourt at Paso with other contemporary ballcourts because there are none (as yet discovered). Likewise the remains of the Mound 6 buildings, and especially the sequence of structures spanning several generations, are rare both at Paso and beyond. But as impressive as these structures were, these initial Early Formative builders were about to be eclipsed by the creations of their descendants toward the end of the second and beginning of the first millennium b.c. One wonders what, if any, social memory remained of the great ballcourt and the big houses at Paso de la Amada in the decades and centuries following its abandonment. New centers sprang up in the Soconusco— such as La Blanca to the south and then Izapa nearer by— all with enormous mound structures and monumental stone sculptures supported by ever-increasing populations. The social and political gaze shifted elsewhere. Perhaps only now, three-and-a-half thousand years later, can we appreciate the legacy and contributions of these earliest villages and towns along the rivers and among the mangroves of the Soconusco.

ACKNOWLEDGMENTS This research was generously supported by the New World Archaeological Foundation, (Brigham Young University), the Social Sciences and Humanities Research Council of Canada, and the Laboratory of Archaeology at the University of British Columbia. John Clark, Richard Lesure, Warren Hill, Vicki Feddema, Mike Ryan, Doug Brown, and Rob Rosenswig helped direct and carry out the excavations at Paso de la Amada, and I thank them for their generosity in sharing the results of their labors. I am grateful for the support of Consejo de Arqueología of the Mexican Instituto Nacional de Antropología e Historia for permission to conduct our

116

emergent complexity

work between 1985 and 1995. The authorities and land owners in Buenos Aires graciously granted us permission to work at the site over many seasons. None of the work could have been done without the skill and patience of several dozen men from that community who helped in all aspects of our fieldwork. I thank Richard Lesure, both for organizing the symposium at UCLA and— along with the other conference participants and two anonymous reviewers—for commenting on a previous draft of this chapter. NOTES 1. All dates listed are calibrated radiocarbon estimations, unless otherwise mentioned. 2. The discussion of Mound 6 builds on the description published much earlier (Blake 1991) before the three earliest and best preserved structures in the Mound 6 sequence were excavated. Some of the information has been presented in unpublished reports and dissertations. 3. During this and subsequent seasons, John Clark and I divided up the labor. While I focused on Mound 6, John carried out excavations at other sites and conducted an intensive survey of the Mazatan zone. Later, Vicki Feddema, Richard Lesure, Dennis Gosser, and Warren Hill joined the Paso excavations for their thesis or dissertation projects. 4. Mound 6, showing the light- colored fi ll making up the mound, is clearly visible on the Google Earth satellite image of August 19, 2003, at 14° 52.592’ N and 92° 29.414’ W (DigitalGlobe/Cnes Spot Image, accessed December 5, 2010). 5. At the time of our excavation we had only the artifacts (ceramic sherds and fi gurine fragments) in the pit to indicate that the burial’s age was approximately contemporary with the later structures, perhaps dating to the Ocós phase. Later we were able to obtain an accelerator mass spectometry (AMS) radiocarbon date on bone collagen from the infant that gave a calibrated median date of 877 b.c. with a 95.4 percent probability range from 972 to 816 b.c. (NUTA 2-395; 2740 ± 30 uncal BP— calibrated using OxCal 4.1 [Bronk Ramsey 2009]) considerably younger than the associated structures, and so undoubtedly intrusive. 6. The details of the ballcourt excavation strategy, its features, and its dimensions are presented in Warren Hill’s Ph.D. dissertation (1999). 7. Mound 7 (like Mound 6— see Note 4) is clearly visible on the Google Earth satellite image of August

19, 2003, at 14° 52.711’ N and 92° 29.428’ W (DigitalGlobe/Cnes Spot Image, accessed December 5, 2010). 8. Note that the two sigma time range on this one date coincides almost exactly with the estimated 300 years of occupation of Mound 6, mentioned earlier. 9. The scale of the construction of both Mound 7 and Mound 6 is dwarfed by the scale recently proposed by John Clark in his analysis of the layout of Paso de la Amada. He proposes enormous earthmoving activities that created platforms, plazas, and sunken courts (Clark 2004a, 2004b). In evaluating the scale of Mound 7 construction, I am comparing it to the other mound structures at the site, not Clark’s proposed larger-scale landscaping.

REFERENCES Ashmore, Wendy 2004 Classic Maya Landscapes and Settlement. In Mesoamerican Archaeology, edited by J. A. Hendon and R. A. Joyce. Blackwell, Malden, MA. Bandy, Matthew S. and Jake R. Fox (eds.) 2010 Becoming Villagers: Comparing Early Village Societies. University of Arizona Press, Tucson. Blake, Michael 1991 An Emerging Formative Chiefdom at Paso de La Amada, Chiapas, Mexico. In The Formation of Complex Society in Southeastern Mesoamerica, edited by W. L. Fowler, pp. 27–46. CRC Press, Boca Raton, FL. Blake, Michael 2010 Colonization, Warfare, and Exchange at the Postclassic Maya Site of Canajasté, Chiapas, Mexico. Papers of the New World Archaeological Foundation No. 70, Brigham Young University, Provo, UT. Blake, Michael, John E. Clark, Barbara Voorhies, George Michaels, Michael W. Love, Mary E. Pye, Arthur A. Demarest, and Barbara Arroyo 1995 Radiocarbon chronology for the Late Archaic and Formative Periods on the Pacific Coast of Southeastern Mesoamerica. Ancient Mesoamerica 6: 161– 183. Blake, Michael, Richard G. Lesure, Warren D. Hill, Luis Barba, and John E. Clark 2006 The Residence of Power at Paso de La Amada, Mexico. In Elite Residences in the New World: From Peru to the Northwest Coast, edited by J. J. Christie and P. J. Sarro, pp. 191–210. University of Texas Press, Austin. Bronk Ramsey, C. 2009 Bayesian analysis of radiocarbon dates. Radiocarbon 51(1): 337–360.

Ceja Tenorio, Jorge F. 1985 Paso de La Amada: An Early Preclassic Site in the Soconusco, Chiapas. Papers of the New World Archaeological Foundation No. 49, Brigham Young University, Provo, UT. Clark, John E. 1994 The Development of Early Formative Rank Societies in the Soconusco, Chiapas, Mexico. Ph.D. dissertation, Department of Anthropology, University of Michigan, Ann Arbor. Clark, John E. 2004a Mesoamerica Goes Public: Early Ceremonial Centers, Leaders, and Communities. In Mesoamerican Archaeology, edited by J. A. Hendon and R. A. Joyce, pp. 43–72. Blackwell, Malden, MA. Clark, John E. 2004b Surrounding the Sacred: Geometry and Design of Early Mound Groups as Meaning and Function. In Signs of Power: The Rise of Cultural Complexity in the Southeast, edited by J. L. Gibson and P. J. Carr, pp. 162–213. University of Alabama Press, Tuscaloosa. Clark, John E., Jon L. Gibson and James Zeidler 2010 First Towns in the Americas: Searching for Agriculture, Population Growth, and Other Enabling Conditions. In Becoming Villagers: Comparing Early Village Societies, edited by M. S. Bandy and J. R. Fox, pp. 205–245. University of Arizona Press, Tucson. Coe, Michael D. 1961 La Victoria, an Early Site on the Pacific Coast of Guatemala. Papers of the Peabody Museum of Archaeology and Ethnology, Harvard University No. 53, Cambridge, MA. Feddema, Vicki L. 1993 Early Formative Subsistence and Agriculture in Southeastern Mesoamerica. Unpublished Masters thesis, Department of Anthropology and Sociology, University of British Columbia, Vancouver. Available at https://circle.ubc.ca/handle/2429/2254 . Flannery, Kent V. 1976 The Early Mesoamerican House. In The Early Mesoamerican Village, edited by K. V. Flannery, pp. 16–24. Academic Press, New York. Fox, John G. 1996 Playing with power: Ballcourts and political ritual in southern Mesoamerica. Current Anthropology 37(3): 483– 509. Green, Dee F., and Gareth W. Lowe 1967 Altamira and Padre Piedra, Early Preclassic Sites in Chiapas, Mexico. Papers of the New

history in space at paso de la amada

117

World Archaeological Foundation No. 20, Brigham Young University, Provo, UT. Hill, Warren D. 1999 Ballcourts, Competitive Games, and the Emergence of Complex Society. Unpublished Ph.D. dissertation, Department of Anthropology, University of British Columbia, Vancouver. Available at https:// circle.ubc.ca/handle/2429/9880. Hill, Warren D., Michael Blake, and John E. Clark 1998 Ball court design dates back 3,400 years. Nature 392: 878– 879. Hill, Warren D., and John. E. Clark 2001 Sports, gambling, and government: America’s first social compact? American Anthropologist 103(2): 331–345. Hodgson, John G., and John E. Clark 2007 New Evidence for Archaic Period Occupations in the Soconusco, Chiapas, Mexico. Paper presented at the Seventy-Second Annual Meetings of the Society for American Archaeology, Austin, TX. Joyce, Rosemary A. 2004 Unintended consequences? Monumentality as a novel experience in formative Mesoamerica. Journal of Archaeological Method and Theory 11(1): 5–29. Lesure, Richard G. 1995 Paso de La Amada: Sociopolitical Dynamics in an Early Formative Community. Unpublished Ph.D. dissertation, Department of Anthropology, University of Michigan, Ann Arbor. Lesure, Richard G. 1997 Early Formative Platforms at Paso de La Amada, Chiapas, Mexico. Latin American Antiquity 8: 217–235. Lesure, Richard G. 1999 Platform architecture and activity patterns in an early Mesoamerican village in Chiapas, Mexico. Journal of Field Archaeology 26(4): 391–406. Lesure, Richard G., and Michael Blake

118

emergent complexity

2002

Interpretive challenges in the study of early complexity: Economy, ritual, and architecture at Paso de La Amada, Mexico. Journal of Anthropological Archaeology 21(1): 1–24. Marcus, Joyce, and Kent V. Flannery 1996 Zapotec Civilization: How Urban Society Evolved in Mexico’s Oaxaca Valley. Thames and Hudson, New York. Moore, Jerry D. 1996 Architecture and Power in the Ancient Andes: The Archaeology of Public Buildings. Cambridge University Press, England. Moseley, Michael 2001 The Incas and Their Ancestors: The Archaeology of Peru. Thames and Hudson, New York. Pearson, Michael Parker, and Colin Richards (eds.) 1994 Architecture and Order: Approaches to Social Space. Routledge, London. Raymond, J. Scott 1999 Early Formative Societies in the Tropical Lowlands of Western Ec uador: A View From the Valdivia Valley. In Pacific Latin America in Prehistory: The Evolution of Archaic and Formative Cultures, edited by M. Blake, pp. 149– 159. Washington State University Press, Pullman. Rodning, Christopher 2010 Place, landscape, and environment: Anthropological archaeology in 2009. American Anthropologist 112(2): 180– 190. Shady Solis, Ruth, and Carlos Leyva (eds.) 2003 La Ciudad Sagrada del Caral-Supe: Los Orígenes de La Civilización Andina y La Formación del Estado Prístino en el Antiguo Perú. Instituto Nacional de Cultura, Lima. Steadman, D. W., M. P. Tellkamp, and T. A. Wake 2003 Prehistoric exploitation of birds on the Pacific coast of Chiapas, Mexico. Condor 105(3): 572– 579. Voorhies, Barbara 2004 Coastal Collectors in the Holocene: The Chantuto People of Southwest Mexico. University Press of Florida, Gainesville.

SIX

Paso de la Amada as a Ceremonial Center Richard G. Lesure

P

aso de la Amada , with its dozens of

low mounds and an occupation from the Barra through Cherla phases, has been known for some time as a significant Early Formative site of the Soconusco region (Blake et al. 1995; Ceja Tenorio 1985; Lowe 1977). Though discoveries of the sequence of high-status residences in Mound 6 and the ballcourt in Mound 7 irrevocably established the site’s larger importance (Chapter 5 in this book), in a sense they pull interpretation in divergent directions. The ballcourt links Paso de la Amada firmly to larger Mesoamerican traditions. The large buildings in Mound 6 are rather different. Outside the Soconusco, there is nothing comparable to the spectacularly preserved Structure 4 of Mound 6 (henceforth, Structure 6-4), with its low clay walls and inset porches. Whereas the Mound 7 finds make Paso de la Amada “Mesoamerican,” those at Mound 6 raise the possibility of precocity, even uniqueness, for Early Formative developments in the Soconusco. Buildings 22 m long are not something we would have expected in the seventeenth century b.c., and they prompt larger

questions of how we should understand the organization and development of the Paso de la Amada community. John Clark’s (2004a, 2004b) recent proposals on the site are thus of great interest. He suggests that those of us who have worked there have approached it with a “naturalistic bias” and a preconceived vision of the “Early Mesoamerican Village” as an unorganized scatter of domestic structures (2004a:53, 57). His bold counterproposal is that the residents of Paso de la Amada constructed their landscape on a massive scale. The site thus deserves the status of “earliest known ceremonial center in Mesoamerica” (Clark 2004a:45). This chapter is a response to that suggestion from the perspective of the excavations in small mounds and off-mound areas that I conducted at the site from 1990 through 1997. I take a “ceremonial center” to be a community deliberately configured for rituals involving substantial numbers of people and embedding, in its large-scale design, a consistent vision of sacredness or the cosmos. This is not a stringent defi nition. From 900 b.c. or so, most

119

substantial Mesoamerican communities were ceremonial centers in this sense. Our question here is the emergence of ceremonial centers in the Early Formative period. My results support some aspects of Clark’s proposal for Paso de la Amada, cast doubt on other aspects, and do not bear at all on still others. The bottom line, however, is that the site was indeed constructed and arranged on impressive scales; it deserves the “ceremonial center” label. The next question, though, is whether we can unveil its par ticular qualities as a ceremonial center, especially if one grants that patterns of spatial organization at Paso de la Amada might be unprecedented given received wisdom on the construction and organization of space at Early Formative sites. A case in point is the question of whether the Mound 6 buildings were residences or temples. The evidence points both ways, suggesting that “temple” versus “residence” is simply not an appropriate categorical distinction in this case. This chapter, in my initial conceptualization, was to have been an extended argument for an alternative categorical distinction— formal versus informal— as a more promising framework for exploring the cultural categorization of space at the site. Indeed, the plan was to champion that distinction not just over temple versus residence, but also over public versus private and even elite versus nonelite. In its final version the chapter is still centrally concerned with differential formalization as a framework for understanding use of space, but changes over the course of the occupation loom large and seem, indeed, to invite application of a series of analytical categories. Differential formalism that cannot be reduced to “ritual” or “residential” was a central theme at the moment of the site’s emergence as a ceremonial center. Yet it was ephemeral, an initial stage in  the development of ceremonialism and community. Subsequent stages are susceptible to analysis with more familiar categorical distinctions.

120

emergent complexity

ANALYTICAL STRATEGY My twin goals are thus to assess Clark’s “ceremonial center” suggestion and to push toward a deeper understanding of the site as constructed, lived, ceremonial space. Clark (2004a, 2004b) founds his proposals on a claim that Paso de la Amada was laid out with the aid of two units of measure, a Standard Unit (SU) of 1666 mm and a Standard Macro-Unit of 52 SUs, or 86.63 m. I am skeptical of the claim, though in all fairness I should point out that my estimate of 30 m for the length of the Locona platform in Mound 32 (Lesure 1999) corresponds to 18 SUs (18 × 1666 mm = 29.988 m). Basically, I am at a loss as to how we could satisfactorily evaluate the proposed units of measure. The fact that the same unit can form the basis for two radically different versions of the large-scale organization of the site (in Clark 2004a and 2004b) is, in my view, a mark against it rather than for it. It is possible, though, to bracket and set aside the question of units of measure as we investigate the construction and use of space at Paso de la Amada, and that is my plan for this paper. Beyond the units- of-measure issue, Clark’s (2004a) ceremonial center designation is based on a claim that the site was massively constructed, laid out according to a grand vision, with contexts for ritual action at scales of 200 m to 300 m (Figure 6.1, square units) and larger patterning at scales of 500 m or more (alignment of squares in Figure 6.1). It is possible to view his claims for the site as falling at one extreme of a continuum of conceivable degrees of “constructedness” and arrangement. At the opposite end of the continuum would be a minimally constructed “early village,” characterized by little intervention into the natural landscape and a lack of patterning beyond the scale of the house yard, perhaps 40 m in diameter. My procedure is to build toward a synthetic assessment of the site from smaller scales. The operating question becomes this: at what scales

FIGURE 6.1 John Clark’s interpretation of the organization of Paso de la Amada. Redrawn and adapted from Clark 2004a:fi gure 2.5.

do we find patterning? As the inhabitants of Paso de la Amada went about their lives, did they act as if the landscape they inhabited was constructed, segregated, and arranged— and, if so, on what scale? Answers to these questions will rely on basic assessments of the nature of constructed forms. This is not, by any means, a simple task given complex underlying sediments and a lack of rock at the site, as well as the effects of erosion, soil formation, and modern plowing. To architectural platforms built to support a perishable structure (Lesure 1997), I would now add lateral or ad hoc extensions to preexisting platforms or natural landforms. Clark (2004a) prompts consideration as well of the possibility of cleared, flattened spaces, either patios or plazas; extensive platforms that could have supported multiple buildings, served as settings for public activities, or partitioned space at a large scale; and subsurface constructions such as ditches, wells, pools, or other waterworks. Beyond the identification of such works, there

is their scale, orientation, and arrangement with respect to each other and to natural landforms at the site. Such issues lead, in turn, to a consideration of the contexts for activity and interaction created by the arrangements of constructed forms, and thus to a consideration of how people used constructed spaces. There is a considerable literature on the interpretation of the Mound 6 buildings and other large or platform-top structures at Paso de la Amada (Blake 1991; Blake and Clark 1999; Blake et al. 2006; Clark 2004a; Lesure and Blake 2002; Love 2007:285–286; Marcus and Flannery 1996:90– 91). There are also grounds for puzzlement, particularly if we attempt to apply to them the dichotomy of temple versus residence (Lesure 1999; Lesure and Blake 2002:6– 10). Still, there is good evidence for the full array of domestic tasks that we find elsewhere in the site: people lived in the Mound 6 buildings. Blake and Clark interpret these as successive houses of the village headman or chief. In the terms of the the aggrandizer model

paso de la amada as a ceremonial center

121

(Blake and Clark 1999; Clark and Blake 1994), contrasts between the construction sequence of Mound 6 and those of other mounds at the site are understandable as the outcome of political competition between households. Our interpretive focus has thus been on the residential qualities of constructed forms. Still, discovery of the ballcourt prompted greater attention to public, ritual, and collective dimensions of constructed spaces. Hill and Clark (2001:343) judged the original aggrandizer model to be “too materialistic and political because it shortchanges the phenomenological side of cultural practices,” even as they emphasized the importance of sponsorship by an aggrandizer—housed at Mound 6— of ballgame events and the court itself. In his proposals concerning Paso de la Amada as a ceremonial center, without giving up any of the basic tenets of the aggrandizer model, Clark moves further toward an emphasis on the ritual, the sacred, and the communal as opposed to the residential, the political, and the individual. The idea that an aggrandizer living at Mound 6 sponsored the building of Paso de la Amada as a ceremonial center he now judges “simplistic.” Instead, “a project this size . . . may have been a ’communal’ endeavor directed and promoted by a coalition of leaders” that led ultimately to the “forging of community identity” (Clark 2004a:61). The suggestion that Paso de la Amada be thought of as a ceremonial center, in other words, raises once again the issues at play in the temple-versus-residence debate concerning Mound 6. It is definitely not my intention to reopen that frustrating discussion. Instead, as a framework for exploring the par ticular qualities of Paso de la Amada as a ceremonial center, I will develop my previous suggestion (Lesure 1999) that a distinction between “formalized” (or simply “formal”) and “informal” activities might help get at the use of constructed space at the site. Formalization is an adherence to restricted modes of activity, often viewed by participants as tradition laden and invariant (Bell 1992,

122

emergent complexity

1997; Connerton 1989:59– 61). Bell (1997:139) notes that formality is one of the most frequently cited characteristics of ritual. . . . Formal activities set up an explicit contrast with informal or casual ones; and activities can be formalized to different extents. In general, the more formal a series of movements and activities, the more ritual-like they are apt to seem to us.

Formalized activity can be important in the reproduction of social power. Individual acts are fleeting, but a formalized per for mance creates a sort of permanence by calling forth memories of a series of repeated actions. In this manner, formalization supports a crucial claim of traditional authority: that the present is irrevocably determined by the past (see Bell 1992:88– 93, 120– 121; Bloch 1974). Formality, then, is a dimension of human action that is of interest to investigations of the emergence of social inequality and political complexity. How is it to be identified and investigated? My orientation here takes inspiration from two sources. First, I attempt to assimilate Bell’s (1992) point that ritual and its dimensions are not inherent categories. They are created contextually, as people recurrently distinguish one activity from another. Formalism becomes a category only as it is repeatedly contrasted with informal activity. The content of each category is created in specific cultural contexts. Such a position poses challenges for actualization in archaeology, but Hodder’s (1992, 1999:30– 65) hermeneutic approach provides a methodology. The questions under investigation concern larger-scale patterning of practices at the site. We enter the hermeneutic circle at small spatial scales, bringing with us certain preconceived notions of what should constitute formal and informal activity. We refine those preconceptions in confrontation with the smallscale evidence. At that point we will have moved some way toward a context-specific understanding of contrasting uses of space. We then take those refined criteria and map them across the site, looking for larger-scale patterns.

POINT OF DEPARTURE: FORMALIZATION AT PASO DE LA AMADA With that methodological program in place, the formal-informal distinction I pursue here can be identified as the product of a previous hermeneutic loop. In pondering my finds in relation to those at Mound 6, I became convinced that more obvious pairs of contrasting categories for the investigation of early complex societies— ritual versus residential, public versus domestic, even elite versus nonelite—were a poor fit for the archaeological record of the site. I suggested formal-versus-informal as an alternative. Formalism is a quality of ritual, but all formal activity is not ritual, and formalized spaces are not necessarily temples. Likewise, while the buildings we interpret as high-status residences were important loci of formalized action, they were also locations for more informal activities. Thus, formal-versus-informal cannot be reduced to elite-versus- nonelite—the distinction concerns modes of activity that may be performed by the same person. The idea is not to do away with the common sets of contrastive terms, but to grapple with their unsatisfactory fit with the archaeological remains from the site. Formalization may appear in speech, gestures, bodily movements, and complicated sequences of actions. It can be suggested archaeologically by spatial contrivances that would have separated and constrained different sorts of activities or by evidence of efforts to promote continuity in repeated actions. Indeed, I am more interested in the spatial dimensions of distinctions between formal and informal activity than in classifying particular activities one way or another. The architectural platforms of Paso de la Amada would have separated interior, raised spaces from exterior, low spaces, a distinction absent at ground-level structures and therefore suggesting some degree of differential formality. In the archaeological record of Mound 6, this theme becomes impossible to ignore. The overall sequence of large buildings, directly superimposed and preserving the same alignment

(Blake and Clark 1999: figure 5) betokens efforts to promote continuity, a sense dramatically reinforced in the transition between Structure 6-5 and Structure 6-4: in the latter case, the line of the front entrance step from the porch to the interior was located precisely above the bizarre, diminutive “wall” segment of the former structure (compare Figures 5.2 and 5.4, Chapter 5). Further inspection of the Mound 6 record reveals other activity traces that fit comfortably with the sense of formalism perceptible in the sequence as a whole. First are two central, subfloor offerings, a small celt and a deer scapula (Blake 1991: figure 11). Second is formal termination of a building, involving dismantling of the perishable structure, filling of postholes with colored sediments, and deposition of fill atop the abandoned structure. Finally, certain rare ritual objects— ceramic statuettes 60 to 70 cm tall, ceramic masks, effigy censers, and greenstone punches (bloodletters?) are present only at platforms (Lesure and Blake 2002). The patterning of such objects helps flesh out the category of formal activity as manifested at Paso de la Amada, but only if we bear in mind that other likely ritual objects (smaller figurines, censers without effigies) appear widely, in both platform and nonplatform contexts. The following, then, are the attributes that, appearing in clusters, would suggest formalization: architectural platforms 50 cm or more in height; traces in the form of post molds or clay walls (or both) of large structures more than 10 m long; careful structure termination; refurbishment of structures or platforms with continuity of location, orientation, and function; subfloor offerings; and the presence of certain rare ritual objects. Nonresidential function for a building would also be of great interest, though in such cases a term such as temple or public building might be fully appropriate and rather more straightforward than abstract talk of formalization. Following Bell’s (1992) logic on ritual more generally, formalized action simultaneously references and distinguishes itself from an alternative: informal action. It makes sense to

paso de la amada as a ceremonial center

123

50

21 32 15

7

12

13

1

11 10 14

4

6

FIGURE 6.2 Map of Paso de la Amada, with excavated mounds shaded

argue for the above as material traces of formal use of space only if it is possible to identify an alternative set of attributes for a contrasting mode of activity, an informal use of space. I will work through the logic in the next section, but the attributes I will end up with are lateral or ad hoc extensions to platforms; traces in the form of post molds of structures 8 m or less in  length; isolated human burials; clusters of burials; domestic garbage on the floor of a structure; pits filled with domestic trash; toss middens representing accumulations of domestic trash on an occupation surface; ditches (often filled with trash); very deep pits that were probably wells; and occupation of an unstable surface of loose sand subject to substantial alterations in rainy season downpours.

EVIDENCE FROM THE SMALL-MOUND AND OFF-MOUND EXCAVATIONS In this section, I consider most of my excavations at Paso de la Amada, with several goals simultaneously in mind. I will, first, tease out

124

emergent complexity

the logic for selecting the attributes just listed as indicating an informal use of space. At the same time, I consider the nature and scale of constructed forms and assemble evidence of the sediments underlying the cultural layers in each location in preparation for an assessment of the large-scale balance between natural and constructed landscapes at the site. The cases are considered more or less in an order of decreasing formalism from Mound 32 to Mound 12, with some more tentative insights into the crucial southwestern part of the site added at the end. It seems useful to note, though, how my excavation strategy developed over five seasons. At the end of the 1990 season, under the benevolent tutelage of Blake and Clark, I excavated small soundings in Mounds 10, 11, 12, and 13 and two test pits (29 and 30) on the low ridge between Mounds 6 and 14 (Figure 6.2). The goal was to prepare for a dissertation on residential differentiation at the site. In the spring of 1992, under the general supervision of Clark, I excavated Mound 1 and, immediately to the south, Test Pits 31, 32, and 33; Test Pit 32 was

expanded into the Area A excavation. Based on Ceja Tenorio’s (1985) test pits and Clark’s reexamination of Ceja Tenorio’s finds, Mound 1 appeared (and indeed turned out) to be an important Cherla-phase construction. I opened a massive extensive excavation, with unexpected results. In the fall of 1992, sobered by the time and resources I had sunk into Mound 1, I began more conservatively at Mound 12, digging a trench across the mound—again with unexpected results. By the time I excavated a series of aligned test pits in Mounds 21 and 32, I was both proficient at recognizing platform fill and well aware of the considerable differences of depositional sequence between mounds at the site. I returned in 1993 (with Blake) to explore a series of hardened surfaces at Mound 12 identified at the end of the 1992 season. A second test in Mound 13 helped clarify the sequence there. All this provided materials for my dissertation (Lesure 1995). I returned in 1997 to explore the Locona platform in Mound 32 (Lesure 1999) and to test Mound 15. Six trenches at Mound 32 successfully defi ned both the size and orientation of an architectural platform. The work on which I draw here thus includes extensive excavations in two mounds (1, 12) and one off-mound location (Area A), substantial trenching at one mound (32), multiple soundings at three mounds (13, 15, 21), and single soundings at two mounds (10, 11) and four off-mound locations (Test Pits 29, 30, 31, 33).

Mound 32 At Mound 32, 89 m2 were exposed in trenches. Excavations revealed a Locona-phase platform (Structure 32-1) approximately contemporary with and similar in size to Structure 4 in the Mound 6 sequence. Several points of similarity help link this platform to the constructions at Mounds 6 and 7, solidifying the suite of attributes I associate with formalism and suggesting certain dimensions of a contrasting, informal use of space.

The platform was approximately 30 x 12 m (Figure 6.3) and probably architectural, based on similarities with Structure 6-4. The fi ll layer was 60 to 70 cm thick, but it had been decapitated through plowing. Originally, its surface must have stood 90 to 100 cm above the surrounding ground. The long axis of the platform was oriented approximately 30° east of north, parallel to the orientation of the nearby Mound 7 ballcourt, and more or less perpendicular to Structure 6-4. Trench 4, extending off the platform to the northwest, revealed a toss midden composed of refuse dumped off the surface of the platform, its earliest layer Locona in date. The scarcity of artifacts in the Locona layer on other sides of the platform (Figure 6.3) suggests that these areas were swept clean, thus maintaining the (formal) spatial division between platform surface and surrounding area. The Locona midden area probably marks the “back” of the building atop the platform. A division between an open, public “front” and a more closed off “back” also seems evident in the layout of Structure 6-4 (Clark 2004a: figure 2.4). The midden itself contained a full array of domestic artifacts, including the usual variety of pottery vessel forms, as well as obsidian flakes, a few personal ornaments, fragments of grinding stones, fire-cracked rock, pumice, pieces of burnt daub, and animal bones. Also present, however, were more than seventy fragments of a hollow ceramic statuette (see Figure 6.3). The figure is anthropomorphic, with female breasts. Its size is unprecedented: the face from chin to top of head is 17 cm, and the original piece must have been 60 to 70 cm in height. Trenches radiating from Mound 6 indicated that Structure 6-4 was surrounded by a prepared clay floor extending somewhat to the sides and back but particularly to the front, forming a patio 45 m wide and extending 13 m out from the building (Clark 2004a:58, figure 2.4). Something similar seems likely for Structure 32-1, though conditions of preservation were in this case poor. The Locona ground

paso de la amada as a ceremonial center

125

FIGURE 6.3 The Locona platform at Mound 32. Inset: large fragments of a hollow ceramic statuette that originally stood 60– 70 cm tall, with location of the fragments indicated.

surface was essentially flat for about 7 m to the front of the platform. Thereafter it began to descend gradually. Occupation of Mound 32 continued apparently without interruption from the Locona to the Ocós phase, but the character of the occupation gradually changed. The platform became a “mound,” with gently sloping sides. The deposition of domestic garbage continued in the back but also crept around to the front of the mound (Lesure 1999: figures 3 to 5). Lateral extensions

126

emergent complexity

to the back (but not the front) increased the surface area of the mound. Associated middens contain all the usual domestic artifacts. In general, the Ocós occupation of Mound 32 displays less order, arrangement, and segregation than does the preceding Locona occupation. There was no longer a clear spatial distinction between the platform and its immediate surroundings, and the distribution of artifacts indicated that the occupants no longer made an effort to maintain such a distinction.

We have here a network of contrasts suggesting an informal mode of spatial practice against which formalism would have been constituted. There is, first, the contrast between the front and back of the Locona platform. The front was formal: the swept- clean patio before a large platform-top building suggesting a public face in which actions were self- conscious, even decorous. The back of the structure, in contrast to the front, was not swept clean. This was an area for the digging of pits and the discarding of garbage. These spatial distinctions maintained by the Locona inhabitants of Mound 32 suggest an important material trace of an informal use of space: the deposition of garbage in trash pits and toss middens. In the Ocós phase, the gradual conversion of the platform into a “mound” and the resort to lateral (even ad hoc) extensions to the mound surface are associated with the digging of pits and generalized discard around the perimeter of the mound. Collectively, these observations suggest an occupation that had altered in character to become more informal than the Locona occupation. Cherla occupation of the mound is indicated by a trash-fi lled pit and a poorly preserved burial. Such features are rare atop platforms, but we will find that these two are commonly associated with Cherla occupation. They seem to constitute a further step toward an informal as opposed to formal use of space. The idea is not that burial rituals were informal, but that the locations chosen for interment were not those otherwise bearing traces of formal action. Adult burials, for instance, do not appear under the floors of large buildings or in frontal, patio areas. I have dwelt at some length on Mound 32 because the occupation of the mound displays a gradually altering character in which we can glimpse, on the one hand, elements of formalism that resonate with the record at Mound 6 and, on the other, a suite of elements that suggest a contrasting use of space. Material indicators of the latter include toss middens, trashfilled pits, burials, lateral extensions to platforms,

and lack of effort to spatially segregate discarded material. Still, while the trajectory of decreasing formalism at Mound 32 is helpful in an initial effort to identify contrasting uses of space, because general changes over time will loom large by the end of my analysis, it will be important to make the case also for synchronic distinctions between formal and informal modes.

Mound 21 In Mound 21, some 130 m northeast of Mound 32, five soundings exposed a total of 10 m2. Locona occupation consisted of a series of surfaces and an associated toss midden on what was probably a low natural promontory of yellow-brown sandy silt. Where it was preserved under platform fill, the uppermost Locona surface had a dark, organic appearance suggesting some degree of landscape stability. The original platform was constructed, like that at Mound 32, of masses of sediments of different textures, both yellow and gray. Artifacts in the platform fi ll were Locona with some Ocós, and construction in the Ocós phase seems likely. The platform was probably architectural, providing the base of a perishable structure. The platform layer at the time of excavation was about 60 cm thick. Its original surface had been removed by plowing. A well-preserved surface to the east suggests that this platform, like that at Mound 32, faced a small patio. Later deposition of brown sandy silt extended the platform to the east and also probably to the west. That appears to be the result of lateral, ad hoc extensions to the platform. There are some similarities here with Mound 32: an architectural platform, raised above an associated patio, the platform subsequently extended in a less formalized fashion. The platform here, however, is later in date than that at Mound 32.

Mound 13 Explorations in Mound 13 consisted of two adjacent soundings in the summit of the mound,

paso de la amada as a ceremonial center

127

exposing a total of 6 m2. On a natural surface of fine yellowish-gray sand with no evidence of soil formation, inhabitants built a platform 50 to 60 cm high, most likely during the Locona phase, judging from its scant contents: Barra with some Locona, both in the fi ll layer and underneath it. The platform was likely architectural, judging from the modest size of the mound and the similarities in color and composition of fi ll with that used for Locona architectural platforms in Mounds 6 and 32. A series of resurfacings may have gradually raised the upper surface by 20 cm or so from the Locona into the Ocós phase, when the previous platform was capped with an additional 50 cm of fill to form the base of a new structure. Occupation continued into the Cherla phase, with the mound acquiring another 50 cm in height. The size and degree of formality of what I suspect were multiple resurfacings is unknown. During the Cherla phase, an intrusive pit was dug into the mound surface and filled with domestic refuse. We have already seen such a mound-top Cherla pit at Mound 32. As in that case, its presence suggests an altered character to the conduct of activities atop the mound, a more informal use of space.

Mound 1 At Mound 1, 182 m2 was exposed in an extensive excavation and trenches. Finds include fragmentary evidence of perishable structures and a large, Cherla-phase platform. The sequence here contrasts with that at Mound 6. While a large, ground-level building was dismantled immediately before platform construction, the platform was not directly superimposed on the preexisting structure but placed off to one side. The substratum in this area consisted of a fine yellow sand overlying coarse gray sand. The Locona through early Cherla occupation layer was a yellowish-brown silty sand 20 to 40 cm thick derived from gradual accumulation on a living surface as well as relatively small episodes of purposeful fi lling. A few post

128

emergent complexity

molds suggest a poorly preserved structure, Structure 1-5, associated with Locona trash pits, and a better preserved successor, Structure 1-4, of which the postholes and patches of hardened floor remained. Structure 1-4 measured 7.5 by 3 m. A few meters away were a trash-fi lled pit and a hearth. These details fall among attributes I am calling informal. The careful termination of Structure 1-4 is thus a surprise. The building was dismantled and the postholes and floor area fi lled in with gray clay, in a mode of formalized practice similar to that observed at Mound 6. Structure 1-4, abandoned sometime in the Ocós phase, lay directly below the summit of the Cherla platform, but that superposition seems likely to have been fortuitous, since there was an intervening layer with a construction positioned differently. Occupation continued into the Cherla phase, with one adult interment (Burial 8) and then, thereafter, construction of a large groundlevel building, Structure 1-2, preserved only where it underlay the Cherla platform (Figure 6.4). What remained were postholes spaced 2 to 3 m apart indicating a wall line running 40° east of north and two postholes suggesting a wall perpendicular to that. A patchy ridge of hardened clay composed of mixed chunks of different colors (including bits of earth burnt red), similar to that observed in the composition of the low walls of Structure 6-4, ran along the exterior of the building. The possibility that these are the dismantled remains of a low wall similar to that at Mound 6 is intriguing, but must remain speculative. What was left was amorphous in shape and only 5 cm thick. I suspect a residential occupation for Structure 1-2, but the large adjacent patio area almost devoid of features deserves note (Figure 6.4). The only two other features on the surface were a small concentration of domestic refuse, Feature 4, and fragmentary evidence of what may have been another structure 8 m to the north of Structure 1-2. The single surviving posthole of this Structure 1-3 was smaller than those of Structure 1-2, raising the possibility that this

FIGURE 6.4 Structure 1-2 at Mound 1, with the overlying platform indicated in dotted lines.

was a smaller residence associated with the large building. One of the main reasons for suspecting residential occupation at Structure 1-2 or in its immediate vicinity is the content of the platform fi ll deposited atop the dismantled structure. The fi ll consisted of a redeposited elite domestic midden. Several indications of formalization at Structure 1-2 can be noted. The postholes seem to be those of a substantial building, perhaps 8 × 16 m. There is the large, swept-clean patio area beside at least part of the structure. The absence of a

trash-filled pit in this extensively cleared Cherla surface is striking given the frequency of such features in comparatively small exposures elsewhere at the site. The strip of prepared floor— or dismantled wall—surrounding the building suggests a symbolic effort to emphasize the separation between exterior and interior spaces. The orientation of the building itself is approximately perpendicular to the orientation of the later structures in Mound 6. Finally, a formalized termination sequence is suggested for Structure 1-2 but not for the (admittedly meager)

paso de la amada as a ceremonial center

129

remains of Structure 1-3. Structure 1-2 was dismantled, its posts removed, and the holes each filled partway with a sandy silt mixed with hardened chunks of clay and burnt earth (potentially scraped from the floor or wall deposit). The upper part of each hole was then fi lled with the same redeposited Cherla refuse that forms the lower layer of platform fill. The platform marks a significant break in the occupation of this area. It was offset from the dismantled Structure 1-2, indicating no effort to cultivate a sense of continuity of practice. Further, it was of a shape different from the architectural platforms of Structures 32-1 or 6-4. The precise shape is unknown, but its width must have been similar to its length, suggesting a nearly round (or perhaps square?) rather than oblong construction, with perhaps one episode of expansion (Figure 6.4). A single line of possible postholes— only a few centimeters were preserved of each—were the only features preserved atop the mound. The absence of any Cherla pit in this large exposure is noteworthy, though the mound has been shaved down by the plow. No Cherla domestic features appeared in the limited off-platform exposures of the trenches. The absence of domestic refuse deriving from use of the platform raises the possibility that this construction was nonresidential in function. If so, it would mark the appearance of a new kind of construction at the site, a point to which I return in my discussion of a similar feature in Mound 12.

Area A The three off-mound test pits to the south of Mound 1 indicated general Locona and Ocós occupation of this area, but only one, Test Pit 32, hit an intact feature. We expanded that test into the Area A excavation with an exposure of 36.5 m2. Features included three trash-filled pits, a trash-fi lled ditch, and three poorly preserved adult burials, dated to the Locona and Ocós phases.

130

emergent complexity

The pits appear to have been dug from a surface no more than 30 to 40 cm below the modern ground surface, though the ancient occupation surface itself was not identifiable because of bioturbation, soil formation, and rodent burrowing. There was never any platform construction in this location, but correlation of the stratigraphy in Area A with that of Test Pit 31, 15 m north toward Mound 1, suggests that in the Locona phase this area was a low, natural, sandy hillock. The finds in Area A reinforce some of the associations I am proposing for an informal use of space (no platform, a location for burials, presence of trash-fi lled pits or trash-fi lled ditches). Even more important, they extend this clustering of attributes back to the Locona phase, helping to make plausible the contemporaneity of the contrast between formal and informal. The one note of formalization here is that two of the burials (and possibly the third, which was quite damaged) were oriented in very roughly the same direction as the long axis of the large buildings in Mound 6.

Mound 15 Two soundings at Mound 15 with an exposure of 4 m2 yielded rather little cultural debris and, on the mound itself, surprisingly thin cultural deposits. The sounding in the summit admits two possible interpretations. The first, based on the distribution of cultural debris by level, would identify this as a low, natural, topographical rise with intermittent settlement on the surface but never any platform construction. The other possibility is that there was a modest platform here (less than 50 cm thick), probably relatively late in date (Cherla phase). The startling fi nding here is that, under either of the suggested scenarios, this low natural rise very close to the Mound 7 ballcourt (see Chapter 5, Figure 5.1) was of little importance and probably unoccupied during the Locona phase at the time the ballcourt was used.

Mound 12 At Mound 12, 132 m2 were exposed in extensive excavations, a trench, and several test pits. Finds include a Cherla platform similar in many ways to that at Mound 1 as well as very complex preplatform deposits in which attribute associations of informal uses of space are further reinforced. Immediately before construction of the platform, this area was characterized by a stable, generally flat, Ocós occupation surface. A oncedeep pit, Feature 11, was by this time only a slight depression on the landscape, a stand of weeds or even perhaps a kitchen garden. Nearby were four trash-filled ditches, a trash-filled pit and other surface concentrations of refuse, a small hearth, and a cluster of three burials. The burials included an infant, a dog (disturbed), and two adults buried in a single grave. The adults were oriented in more or less the same direction as the burials in Area A—and thus roughly in the same direction as the long axis of the large buildings at Mound 6. Apart from that one element of formalism, the use of space seems characteristic of what I am referring to as informal. This area was likely the yard of a house or a use area shared by multiple households. Any associated houses would have been in ground-level dwellings, not on platforms. The complications at Mound 12 lie in the stratigraphy underlying this stable late- Ocós surface. The Barra-phase inhabitants settled on a surface of loose, unconsolidated, mediumgrain gray sand that provided a distinctly unstable ground surface. Shallow ditches, some with artifacts apparently damaged by water transport but all eventually filled with sediment and trash, were identified in Locona and early Ocós levels. Further, the inhabitants appear to have excavated a deep pit, Feature 11, probably to serve as a dry season cistern or well. Feature 11 was approximately 9 m across and over 3 m deep when it sat open late in the Locona phase. It was fi lled eventually with clay, laminated sand layers, and Locona and Ocós

trash. Though the lowest excavated levels of Feature 11, at the modern water table, contained late Locona or early Ocós sherds, suggesting that the feature was an intrusive pit from that era, evidence from the 1993 extensive excavations indicates that there had been a previous pit in the same location earlier in the Locona phase. The extensive excavations revealed a series of hardened Locona floors or occupation surfaces, the upper ones bearing the marks of numerous posts (see Blake et al. 2006). The postholes seem to be a palimpsest, apparently from a series of small houses (less than 8 m in length) or outbuildings. No one complete structure can be sorted from the chaos, but there are several instances of postholes in lines approximately parallel to the orientation of the ballcourt. Successive floors are separated by a few centimeters of fi ll, and except for the hints of a recurring general orientation, the confusion of the posthole record suggests no concerted attempt to maintain continuity of activity in successive buildings. The implications of all this are a distinct lack of stability to the living surface of this area, a degree of instability that would seem the polar opposite of, but contemporary with, the sweptclean formality and carefully contrived continuities in the use of space at the Mound 6 and Mound 32 platforms. Features associated with this general instability of living surface include palimpsests of postholes from structures 8 m or less in length and exhibiting no concerted effort to maintain continuity of practice; the cistern or well; ditches that served as places to dump trash; a stack of large sherds beside the cistern, apparently left in provisional discard; and an isolated (disturbed) burial early in the sequence. Though the surface instability decreased over time, informal use of this area continued through the very end of the Ocós phase or the early Cherla phase, when, in a dramatic break with the previous mode of occupation, a large platform was constructed. The status of Mound 12 as a “mound” dates particularly from this Cherla-phase construction event. The platform

paso de la amada as a ceremonial center

131

layer was about 80 cm thick, but, like other platforms discussed here, it had been damaged by plowing. It was probably originally over 110 cm high, but not much more than that, judging from three features that penetrated the platform from above (and appeared immediately below the plow zone): a burial almost entirely plowed away, a horseshoe-shaped hearth filled with sherds dating to some epoch after the Early Formative, and a probable offering consisting of three shaped-stone spheres 5.4 to 6.4 cm in diameter, two of white andesite and one of granite. The dating of the burial and the offering is uncertain. Neither was, for instance, centrally located on the platform. The Mound 12 platform appears to have been constructed in a single episode in a shape closer to its near contemporary at Mound 1 than to the earlier platforms at Mound 6 or 32. Dimensions were 20 to 22 m north-south and perhaps 26 m east-west. As at Mound 1, no Cherla pits penetrated the surface of the mound, whereas such pits were frequent elsewhere at the site. Limited off-platform exposures did not reveal toss middens or other features associated with occupation of the platform surface. At the time of excavation, I thought that Cherla constructions at Mounds 1 and 12 were probably architectural platforms for elite residences. However, there is actually no evidence of such occupation in the excavated sample from these platforms. I am not concerned with the paucity of postholes, since it is clear that the surfaces of both have been damaged by plowing. It is the absence of domestic garbage that I consider noteworthy. The hypothesis I would now propose is that these platforms represent public buildings with nonresidential functions. They would thus mark the appearance of a new form of building and a new mode of use of space at the site.

Mounds 10 and 11 and the Mound 14–Mound 6 Ridge So far, I have identified a proliferation of what were probably architectural platforms (built as

132

emergent complexity

the base for a single building) and lateral or ad hoc extensions to platforms, but not extensive platforms built to support multiple buildings or to delimit large public spaces. The southwestern corner of the site appears to be more promising. The later Locona and Ocós expansions at Mound 6 can be deemed extensive platforms (Blake et al. 2006). They are not lateral or ad hoc, because the preexisting platform was entirely encased in a new, higher platform, with attention to continuity of alignment and expanded platform-top space outside of the structure. Outdoor features atop the successive platforms included trash-fi lled pits, human burials, and a fire pit or oven (Blake 1991). When I propose a history of ceremonialism at Paso de la Amada, I will make much of the implications of informality raised by the presence of these platform-top features. They are of interest at the moment because they suggest platform-top, outdoor activities of the sort that might appear on extensive (rather than more narrowly architectural) platforms. Clark (2004a), elaborating on suggestions by Lowe (1977:211), proposes that a Southern Plaza dominated the southwestern part of the site. It would have been delimited by Mound 6 to the southwest, Mound 7 to the northwest, and Mound 14 to the southeast (see Figure 6.1). Clark’s long trench from Mound 6 crossed what would be the southwestern corner of the proposed plaza (see Chapter 5, Figure 5.1). Clark (2004a: 57– 58) found the Locona ground surface there to be flat and devoid of features. The orientations of the Mound 6 buildings and the Mound 7 ballcourt, approximately perpendicular to each other, seem consistent with a larger organization around a plaza. In other words, Mounds 7, 6, and 14 might have been built as a single complex. That observation raises the possibility of earth movement at a massive scale. There is the plaza area itself. Was this constructed in any significant way? Clark (2004a:59) speculates that it might have been hollowed out, with the earth removed from that area transported more than 300 m to the northeast for fi ll. Then there is

the larger topographical prominence topped by Mound 6. How much of that is artificial? Finally, there is the ridge between Mounds 6 and Mound 14. Could that be one long, linear platform? My excavations in this part of the site consist of isolated 1 × 2 m soundings in Mounds 10 and 11 and along the ridge between Mounds 6 and 14. No final conclusions can be drawn from such small tests, but while we await a published report on Clark’s trenches, these provide valuable glimpses into the stratigraphy of this part of the site, especially when they are considered in comparison to my other excavations. A recurring pattern in units downslope from mounds is a gray, organic, often clayey layer immediately beneath the plow zone. This seems to be an indicator of long-term stability and, generally, sediment accumulation (due to slope wash, plowing, or seasonal inundation) since the Early Formative. Any such layer is absent atop mounds decapitated by plowing. I consider the proposed Southern Plaza, the larger Mound 6 landform, and the linear Mound 14 ridge. Mounds 10 and 11 lie toward the middle of Clark’s proposed Southern Plaza (see Figures 6.1 and 6.2 herein; also Clark 2004a:57– 58 and his figure 2.5). One immediate problem with the plaza interpretation is that there is some low modern relief to this area. Mound 11 appears as a finger extending into a lower-lying area, including an old river channel curving around the northern edge of Mound 7. The lower stratigraphy of the Mound 11 sounding reinforces that observation: a pre- occupation ground surface of fine yellow-brown sand slopes off to the north at this location. A small test between Mounds 12 and 11 descended over a meter in a recent deposit of organic-rich clay, again suggesting that modern relief in this part of the site is less marked now than during the Early Formative occupation. The low-lying areas to the north and east of Mound 11 were probably not part of any Southern Plaza. Perhaps the most interesting layer in the Mound 11 sounding lies directly above that sterile sand. It is composed of a sandy clay with

pockets of differently colored sediments. The cultural contents are Locona and Barra in date. I am inclined to see this as fi ll that extended this low natural ridge to the north. That fi ll was deposited atop an earlier Locona pit. The new, raised Locona occupation surface at Mound 11  would have corresponded in elevation to the  likely pre- occupation ground surface of yellowish-gray clayey sand at Mound 10, some 40 m away. Locona construction activity at Mound 11 could thus have involved the creation of an extended public space along the lines suggested by Clark, but at least partly as the result of deposition rather than removal of earth. Still, consideration of surface topography along with excavation stratigraphy indicates a more modest space than originally suggested by Clark. Test Pit 30 was located on the larger topographical eminence crowned by Mound 6. A nearly sterile yellow-gray sandy clay (overlying gray sand) is probably the pre-occupation ground surface. Atop this are 60 cm or more of LoconaOcós occupational accumulation likely including extensive platform construction associated with Mound 6. The gray layer beneath the plow zone suggests sediment accumulation since the Early Formative, probably as a result of the transfer of sediments from slope wash and plowing. The best case for an extensive platform in the excavations reported here was revealed in Test Pit 29. Though the modern ground surface is about 30 cm higher than at Test Pit 30, the likely pre- occupation ground surface is some 40 cm lower, leaving 150 cm of accumulation to be accounted for. The test pit itself was located toward the southwestern end of the low, linear elevation of which Mound 14 forms the highest point. It seems likely that the location of Test Pit 29 was a topographical low point early in the occupation of Paso de la Amada. The light gray clayey layer above sand—apparently the pre-occupation ground surface—is 50 to 60 cm thick here, suggesting accumulation due to rainy season inundations at this location. The uppermost layers of the sounding derive from slope wash and deposition from plowing, with the Cherla ground

paso de la amada as a ceremonial center

133

surface at around 60 cm depth. It was in the 60 to 80 cm level that I first noticed a concentration of artifacts that would eventually be more clearly identifiable as fill of an intrusive Cherla pit. Thus, by the time of residential occupation in the Cherla phase, this location had experienced 90 cm of artificial accumulation of sediments, probably as platform fill. In 1993, Warren Hill excavated a sounding in Mound 14 itself; I examined sherds from that unit for my dissertation (sample 1401 in Lesure 1995). As at Test Pit 29, there was significant platform construction, superimposed in this case on a Locona domestic midden. As much as 60 cm of accumulation is a likely Cherla construction. Given the topography between Mound 14 and Test Pit 29, it seems likely that the platform registered in the latter pit extended all the way to Mound 14— and that much of this low, linear ridge is of artificial construction. My general assessment of the Mound 14 ridge is thus in close agreement with that of Clark (2004a:57). Still, I would propose a construction date distinctly later than that of Mound 7 and the earliest buildings in Mound 6. No midden such as that under platform fill in the Mound 14 test underlies early construction at Mound 6 or 7. That point suggests a later Locona or Ocós date for the extensive platform at Mound 14. The significant Cherla presence in both test units also contrasts with the records at Mounds 6 and 7.

THE SITE AS A CONSTRUCTED LANDSCAPE Moving toward synthesis, I first consider Paso de la Amada as a site constructed and arranged on impressive scales. As polar opposite points of reference, we have, on the one hand, the image of an “early village” consisting of a hodgepodge of houses perched precariously on natural landforms. On the other hand, we have the suggestion that the topography of Paso de la Amada is the result of human intervention at horizontal scales of 200, 300, even 500 m.

134

emergent complexity

The pre- occupation landscape was formed of relatively recent (Holocene era), waterworked deposits. Surface relief was distinctly greater than it is today. Wind may have played some role in creating an undulating ground surface (with dunes), but the interlayering of sediments of different textures (including clay in Mound 32 trench 3) indicates that water was the primary factor. The complex relief probably resulted when old levies were cross- cut by later channels. The Coatán River no longer flowed through the site at the time of earliest settlements (see Chapter 7 in this book). Though it is possible that some minor channels were activated seasonally, by the earliest Formative period, lower-lying areas were locations for the deposition of fine-grained sediments in low-energy environments—a condition that has persisted to the present. The landscape was vegetated. At least, that seems the most likely explanation for the gray clayey layer underlying Early Formative constructions in several locations of varying elevation. In some areas (Mounds 1, 12, and 13), the earliest settlement was directly on sand, either because a surface layer of incipient soil formation was scraped off or because peaks of old over-bank river deposits (or else dunes) remained free of vegetation. The Early Formative inhabitants, then, settled on a naturally undulating surface with differences of elevation generally greater than the current topography of the site. Still, at least some of what are today topographical high points were low points at the time of initial settlement. The inhabitants built numerous earthen platforms. How big were such constructions? What was the scale of individual interventions in the natural landscape? I do not find support for the most ambitious of Clark’s (2004a:58– 59) suggestions, that vast quantities of earth were removed from a Southern Plaza area and used in the creation of a massive artificial elevation to the northeast—in other words, that there were interventions at horizontal scales greater than 300 m. Where I

have a good understanding of the composition of platforms (particularly in Mounds 1 and 32), it is possible to explain the contents with reference to deposits that would have existed within 40 m of the platform itself. In those cases, at least, there is no need to appeal to transport of sediments from distances beyond that. Most earthen constructions were architectural platforms with horizontal dimensions of 10 to 30 m and vertical dimensions of 0.5 to 1.5 m. Larger constructed forms would be considered extensive platforms. They are much rarer and confined (I would propose) to the southwestern corner of the site, the ceremonial core. There is the Mound 7 ballcourt, a significant portion of the landform on which Mound 6 sits, and the Mound 14 ridge. Horizontal dimensions in these cases seem to be in the range of 40 to 150 m (the twin mounds of the  ballcourt are each 80 m long and 30 m wide [Hill, Blake, and Clark 1998]). We would expect a ceremonial center to be not only constructed but also planned beyond the scale of the individual house yard. Clark’s (2004a, 2004b) discussion of planning focuses on his proposed unit of measure. A different instance of what looks like planned regularity at spatial scales larger than any constructed form emerges in the evidence reviewed here regarding the orientation of buildings, platforms, and burials. Clark (2004a), noting the approximately perpendicular orientations of the ballcourt and the long axis of the buildings at Mound 6, reoriented the map of the site to follow the axis of the ballcourt, a move I would heartily endorse and support with further evidence. Orientations of the ballcourt and large buildings in Mound 6, Mound 1, and Mound 32 approximate either 35° east of north or 55° west of north. This kind of coordination indicates site planning at a large spatial scale (the Mounds 6 and 32 buildings are 400 m apart). Use of these orientations persisted from the early part of the Locona phase (Mounds 6 and 7) to the end of Ocós or early Cherla (Mound 1),

a period of 250 to 300 years. On the other hand, the recurring patterns of large-building orientation, though striking, do not seem to have been very precisely replicated, even at Mound 6. Further, the meager evidence from small structures suggests that only some of them conformed to the organizational regularities noted among larger buildings. The orientation of adult burials, though, is an intriguing addition to the picture. The burials are more variable than the large buildings, but the distribution is distinctly nonrandom, with orientations approximating that of the long axis of the Mound 6 buildings. In other words, burial orientation fit a larger scheme of site planning. The fit is quite rough, but such a pattern would seem reasonable if we imagine the circumstances in which people would have oriented a corpse in a grave. In sum, though I am more conservative in my estimates of human interventions in the landscape and though I have ignored completely the issue of units of measure, my basic assessment of Paso de la Amada is in agreement with that of Clark (2004a). The site was indeed constructed and arranged at scales appropriate to a ceremonial center.

DIFFERENTIAL FORMALISM My second goal is to move from discussing if Paso de la Amada was a ceremonial center to asking how it was a ceremonial center. The inquiry here is into how spaces were inhabited. To avoid what I foresaw as a fruitless and frustrating temple-versus-residence debate, I have been looking for differential formalism. It is possible to perceive two modes in the use of space at the site. In formal mode, spaces for activities were segregated (on versus beside a platform, indoor versus outdoor, patio versus discard area), buildings were terminated in systematic sequences and refurbished with continuity of alignment and function, and rituals included the placement of subfloor offerings. Formalized activity is particularly associated

paso de la amada as a ceremonial center

135

with platforms and with large buildings longer than 10 m. Certain rare ritual objects appear only in contexts in which other attributes of formalization are also present, yet domestic garbage indicates that people lived in these buildings. Any actions deserving the label formalized should have been constituted in contrast to other, informal actions. I developed an understanding of the specific content of the informal at Paso de la Amada by observing patterns at the Locona platform at Mound 32, which resonate with the plan of Structure 4 at Mound 6 to suggest that large buildings were treated as if they had a distinguishable front and back. The front was a setting for formal action, the platform highlighting the distinction between the interior of the building and the swept- clean patio before it. Certain informal activities—the digging of pits, the deposition of domestic garbage—were relegated to the back of the platform. Contrasts between the Locona and Ocós occupations emphasize the greater formality of the former and provide, from the latter, further material correlates for an informal use of space. In those cases, we expect lack of arrangement, spatial segregation, rites of termination, or selfconscious attempts to promote continuity in the use of space. Distinctions between indoor and outdoor would not have been marked or elaborated, and domestic garbage was deposited in whatever location was most convenient. In this section, I synthesize evidence for differential formalism, moving from individual excavation locales to larger spatial scales. The discussion is broken down by phase. As I mentioned at the outset, time emerges as a key variable in the character of ceremonial life at Paso de la Amada. The maps (see Figures 6.5 and 6.6) focus on the central part of the site, where most excavations have been located. To the evidence of my small-mound and off-mound excavations just reviewed, I have also added Mounds 6 and 7 (based on Blake 1991; Blake and Clark 1999; Blake and Feddema 1993; Blake et al. 2006; Clark 1994, 2004a; Hill 1999; Hill,

136

emergent complexity

Blake, and Clark 1998; Hill and Clark 2001). Following Clark (2004a), I have reoriented the map so that it corresponds to the orientation of the larger constructions. Material traces of formalization are indicated by symbols with white interiors, while traces of informal uses of space are shown with all-black symbols. Where something is known of the shape and orientation of platforms, those are drawn in with heavy lines (somewhat larger than their true size). A heavylined square indicates a platform with unknown orientation and dimensions.

Locona Phase In the simplest of terms, what I have proposed is that indicators of formal and informal activity should cluster separately. That claim works well for the Locona phase (Figure 6.5) once we qualify that simple version by noting the presence of informal discard areas segregated within the formal spaces of Mound 32. Mounds 6 and 32, both large buildings associated with domestic refuse, emerge as formalized spaces. Mound 13 seems likely to be something of the same sort of construction, but the excavated sample is too small to document formalism. Contrasting with Mounds 6, 32, and (probably) 13 are the informal occupations at Mound 12, Mound 1, and Area A, where houses seem to have been at ground level and smaller than their platformtop counterparts. Finally, there is the ballcourt and the possible plaza areas in the southwestern corner of the site. Several features of the southwestern corner suggest that this was already a special area in the Locona phase, a ceremonial core. Beyond the ballcourt, there is a long series of superimposed constructions in Mound 6, a pattern apparently unique at the site. Finally, there is Clark’s (2004a) astonishing report of a level Locona surface throughout his 215 m trench between Mounds 6 and 7, the basis for suggesting a plaza for public gatherings in this area (see Chapter 5, Figure 5.1). The plaza area may have been expanded incrementally into sur-

rounding bajos, as at Mound 11. There is, however, something strange about this ceremonial precinct. Given our interpretations of Mound 6 as a high-state residence, the Locona-phase ceremonial core was, in Clark’s (2004a:65) pithy phrase, “a plaza without a shrine.” For the Locona phase, it is noteworthy that other large houses besides Mound 6 were foci of formalized activity. The orientations of those buildings conformed to the layout of the ceremonial core. They were interspersed with informal habitation areas, including small houses that at least sometimes followed similar alignments. Large, platform-top houses were not closely linked to the ceremonial core. Mounds 32 and 13 are separated from the ballcourt and each other by old river channels— that is, what would have been seasonally flooded bajos. Mound 15, though similarly separated from the ballcourt, was nevertheless an unoccupied natural rise closer to that public facility than were the locations chosen for platform construction at either Mound 13 or Mound 32. Two other known Locona platforms, Mounds 4 and 50, shown in Figure 6.2, were located at even greater distances from the core (Clark 1994; Lesure 1997). If I am correct to extrapolate the formalism documented at Mounds 6 and 32 to other Locona platforms (Mounds 4, 13, and 50), then it would appear necessary to understand ceremonial activities at Locona-phase Paso de la Amada as not being lodged solely in the core but instead disbursed to multiple foci across the site. The platform-top buildings were nodes of formalization scattered across the landscape of Paso de la Amada. It is possible that Loconaphase settlement organization was modular, with each module consisting of a formal, large building that was often atop a platform— a headman’s house?— associated with a cluster of smaller, more informally occupied residences. In this scenario, Mound 6 could have initially been part of a module similar to the others, but its role would have changed as it was enlarged to monumental proportions.

In short, Paso de la Amada emerged as a ceremonial center during the Locona phase, but with some strange features. The framework of differential formalism seems to illuminate the patterning of archaeological features for this phase. I will elaborate in a moment on what insights that framework provides into ceremonial life at this time. First, though, I trace the eroding applicability of the framework in the Ocós and Cherla phases.

Later Locona and Ocós The big change at the site in late Locona to early Ocós is the stark disparity between monumental construction in the ceremonial core and reduced platform construction in other parts of the site. I have suggested that it was at this time that Mounds 6, 7, and 14 were united into a single complex with construction of the Mound 14 ridge and perhaps with the deposition of fi ll between Mounds 7 and 14 to create a plaza 200 m across (Figure 6.6, top). The massively constructed Mound 6 would have dominated the southwestern end of any such plaza. At the summit of this mound stood a large residential building, a chief’s house, indicating successful intergenerational transfer of power, authority, and community leadership from the last half of the Locona phase through Ocós. The ballcourt seems not to have been further expanded at this time. Its sides began gradually to erode (Hill et al. 1998), but it could have continued to be in use for some time. Nothing was going on in other parts of the site to match construction activity in the ceremonial core. There was no longer a formal platform at Mound 32. Architectural platforms were built in other locations, including Mounds 13 and 21, but those were simply not comparable to Ocós-era construction at Mound 6 and in the ceremonial core generally. Thus, the Ocós phase would have been a time of maximum differentiation between the built spaces of the southwestern core, where there were spaces whose function varied from public to residential, and the rest of the site.

paso de la amada as a ceremonial center

137

FIGURE 6.5 Summary of results: Locona phase

FIGURE 6.6 Summary of results. Top: Later Locona and Ocós phases; bottom: Cherla phase.

It is very significant, then, to find that patterns of differential formalism seem to have eroded at this time. At Mound 1, there are hints of formal termination of a small structure. More serious are trends away from formalized uses of space in multiple mound sequences. In Mound 21, a probable architectural platform with an associated patio was subsequently subjected to ad hoc expansion. A similar trajectory is documented in greater detail at Mound 32, where the formalized Locona space became, in the Ocós phase, an informal one. We might be inclined to ascribe these patterns to a growing disparity in status and authority, imagining that the ritual functions previously disbursed among multiple high-status residences had been transferred to the chief’s house at Mound 6. This is satisfying logic, except that the same trend toward informality in the use of space appears also at Mound 6, where even as the platform reached monumental proportions, platform-top burials appeared and trash-fi lled pits multiplied, appearing even to the “front” of the building. Thus, social differentiation in the Ocós phase is reflected in the centralization of construction activity. Evidence for how the spaces were inhabited reveals no such strong differences. There is, first, no convincing evidence among the artifacts for economic differences between Mound 6 and residences in other parts of the site, particularly Mounds 1, 12, 32, and Area A (Lesure and Blake 2002). Second, the differences in formalism that had characterized domestic life early on were also ebbing away. The character of ceremonial life at Paso de la Amada was changing.

Cherla Phase The Cherla phase (Figure 6.6, bottom) saw significant changes in the spatial organization of the site, in particular the abandonment of the ballcourt and Mound 6 (Clark 2004a; Hill, Blake, and Clark 1998). Constructions at Mounds 1, 12, and probably 14 indicate a shift in the focus of public activity. The platforms documented at

140

emergent complexity

Mounds 1 and 12 were shaped differently from earlier architectural platforms, and I have suggested that they supported some new class of nonresidential building. The lack of Cherla trash pits in the large areas exposed atop those platforms contrasts with the pervasiveness of such pits in comparatively small exposures elsewhere. The appearance of pits and burials atop mounds may indicate that earlier platforms were simply “mounds” by this time, indistinguishable from natural topographical features. The Cherla phase is the first in which differential access to jade ornaments, iron ore mirrors that were also probably costume components, and imported obsidian is identifiable among samples of domestic garbage. There also seems to have been a correlated difference between households in the use of clay earspools (Lesure 1995). Though the residential occupation of almost every high point might be taken as a break with earlier practices, it is worth recalling the orientation of the late Ocós or early Cherla Structure 1-2 and other attributes of formalism in that case, all indicative of some continuity of practice. The distinction between formal and informal ceases to be relevant at some point in the Cherla phase. One problem is that, since this was the last phase of major occupation at the site, Cherla surfaces tend to be near the modern ground surface rather than protected by layers of later fill. If I am correct in my current interpretation of the Mound 1 and Mound 12 platforms as being nonresidential— a claim that will require more excavation beside the platforms to evaluate conclusively—then it may well be that a distinction between residence and temple emerged in the course of a Cherlaphase reorganization of the site.

TOWARD A HISTORY OF PASO DE LA AMADA AS A CEREMONIAL CENTER My proposal is that the original ceremonial center of Paso de la Amada, during the first part of the Locona phase, is best understood

not as a scattering of temples and houses but rather as a landscape of formal and informal settings, constructed and arranged, differentiated and juxtaposed. There was already at this point a ceremonial core, but strangely, and as Clark has pointed out, it lacked a temple. Certain houses scattered across the site, particularly those whose construction required labor beyond that of household members, were loci for self- consciously formal activity, even as habitation at surrounding small houses was unselfconsciously informal in character. I will first spin out some of what I see as the implications of these patterns for understanding the par ticular qualities of early Paso de la Amada as a ceremonial center. Then I will turn to the question of change over time. The pattern of heightened formalism at large residences suggests that the ordinary activities of a segment of society—headmen, aggrandizers, or lineage leaders—were continuously on display. To whom? Presumably to their followers from whose labor they benefited, but likely as well to supernatural forces or entities. The benefits of high status—which centered on inhabited space and do not seem to have included the ability to retain unusual quantities of fancy pottery, imported obsidian, or jade ornaments—were assumed at certain costs. Those deserving high status seem to have behaved decorously and self- consciously even in the “privacy” of their own houses, perhaps because as beneficiaries of the labor of others they were continuously under public surveillance. Quite likely their actions were under supernatural surveillance as well, their own mediating function between people and gods being conducted through the continuous stream of their daily actions rather than merely in the fleeting moments of par ticular rituals. In such a system, there would be no real categorical division between temple and residence. Conceptions resembling the distinction between public and private would also likely be absent since the “private” lives of some but not all people were “public.” Such a system would

box in the wheeling- dealing aggrandizers of Clark and Blake’s (1994) model with duties, requirements, and constraints on innovative action. We have generally treated the buildings of Mound 6 as fundamentally high-status residences with certain templelike qualities. Based on the evidence reviewed here, I would suggest (speculatively) that it might be equally valid to flip the emphasis and imagine community leaders themselves, or maybe their bodies, as being akin to temple paraphernalia. Inhabiting a “temple” required an everyday existence lived as a continuous stream of pious action, entailing a nest of constraints that did not bind people living in small, ground-level houses. All this sounds a bit strange, but it is the result of an extended effort to understand the particular qualities of this early ceremonial center by privileging the archaeological record of the site, minimizing assumptions based on what is known from other areas or times, and allowing standard categorical distinctions (such as temple versus residence) to be altered in confrontation with the archaeological record under consideration. The goal was to move a step or two along a hermeneutic trajectory toward a better understanding of how the site’s original inhabitants categorized space and activity. Still, one theme that emerges very clearly is the importance of time. The strangeness at Paso de la Amada that has prompted me to develop the distinction between formal and informal appears to erode gradually over the course of the occupation. Though it would be nice to have more data, especially from the Cherla-phase occupation, the trajectory I see at this point is a gradual alteration in the character of ceremonial life, involving first the emergence of a categorical distinction akin to public versus private and then, as already suggested herein for the Cherla phase, temple versus residence. Table 6.1 assembles various lines of evidence for a three-stage history of Paso de la Amada as a ceremonial center. My excavations revealed only a few scattered pits from the late Barra occupation of the site in the eighteenth

paso de la amada as a ceremonial center

141

century b.c. and no evidence of significant construction. I understand Clark (2004a) to argue that this era pre- dated the emergence of the site as a ceremonial center, and my data would be consistent with that idea. The story here, then, begins in the seventeenth century b.c., in the first part of the Locona phase. Inspired by Clark’s assessment of this same era, I call this first stage of ceremonial life “the forging of community.” Elements of site planning that would persist for several hundred years in the orientation of large buildings and burials date to this epoch. The ballcourt was the principal nonresidential facility, and, following Hill and Clark (2001) and Clark (2004a:61), it seems likely that activities there and in the adjacent plaza areas were instrumental in the forging of a larger sense of identity in what was a rapidly growing community. Still, the ceremonial core lacked a temple, and my argument would be that the functions we might expect of a temple were dispersed across the site, lodged in the pious daily activities of those leaders who inhabited large, platform-top buildings. From the sixteenth through the fifteenth centuries b.c., ceremonial life at Paso de la Amada entered its second stage, that of a mature ceremonial center. Leadership functions that had previously been distributed across the site were increasingly concentrated in the hands of the inhabitants of Mound 6. The succession of structures there, on an ever- expanding platform, point to successful intergenerational transfer of authority—in other words, hereditary inequality (see Chapter 5 in this book; Blake and Clark 1999; Clark 2004a). Construction activities in this era were concentrated in the ceremonial core, and I would propose that the full Southern Plaza (Mounds 6, 7, and 14 as an integrated complex) dates to this time (Clark [2004a:60] places that development earlier). It is somewhat surprising that there is no evidence of interhousehold differences in artifacts at this time (Lesure 1995; Lesure and Blake 2002). It may be that there were subtle differences that are hard to pick up archaeologi-

142

emergent complexity

cally. I suspect that access to ritual knowledge was still the only ideologically sanctioned basis for social inequality, but I will not try to justify that claim here. Of more interest at the moment is the trajectory of decreasing interhousehold differences in the formality of domestic life, a pattern that includes Mound 6. In other words, as architectural disparities between the Mound 6 residence and other residences became greater, the character of daily life at Mound 6 became more informal, just as it did at Mound 32. That would seem contradictory, unless it signals larger changes in the character of ceremonial life at the site. As far as we know, the site was still without a temple (though we really have no idea what might be in Mound 14). The ceremonial functions that had previously been disbursed to the large residences of leaders across the site were probably now co- opted by the inhabitants of Mound 6. If those functions were now also enacted through specially constituted rituals in the Southern Plaza rather than through a formalization of daily existence at the Mound 6 residence, then the ebbing away of differential formalism may be explained. The domestic activities of community leaders were no longer cosmologically charged and under public surveillance. Both high- and low-status people occupied their houses “informally.” It may be appropriate to envision the emergence of a categorical distinction akin to public versus private with a “public” sphere now firmly lodged out in the Southern Plaza itself and the residence increasingly “private.” The third and final stage of the history of Paso de la Amada as a ceremonial center corresponds to the Cherla phase in the fourteenth century b.c. Major changes were under way during this final century of the site’s existence as a thriving community. The Southern Plaza was abandoned as a focus of public life; there may have been residential occupation atop the abandoned ballcourt. There was significant platform construction, less concentrated than during the late Locona or Ocós but still with a new center of activity that included Mounds 1,

TABLE 6.1 Elements of a History of Paso de la Amada as a Ceremonial Center SIZE OF + INDICATES STRENGTH OF PATTERN

Late Barra

Early Locona

Late Locona

Ocos

Cherla

+

+

+

The Forging of Community Orientations (35° E of N/55° W of N)

?

Construction activity in ballcourta Differential formalism in use of space

?

Dispersed construction of residential platformsb

+

+

+

+

+

+

+

+

+

+

The Mature Ceremonial Center +

Intergenerational transfer of authorityc

Spatial concentration of construction activityd Southern plaza as a public or ceremonial facilitye

+

+

+

?

+

+

+

+

+

Rank as an Embodied Quality? Nonresidential architectural platforms?f

?

?

+

Differential access to imported goodsg

+

Earspools as status differentiationh

+

Century b.c., calendar years a

18th

17th

16th

15th

14th

Hill, Blake, and Clark 1998; Hill 1999; Hill and Clark 2001. Lesure 1997. c Based on the interpretation of Mound 6 as a high-status residence, the long sequence of superimposed constructions there (Blake and Clark 1999), and the absence of such sequences in other mounds. d Based on Blake and Clark (1999), Blake et al. (2006), Clark (1994, 2004a), Lesure (1997, 1999), and arguments developed here concerning the Mound 14 ridge. e Based on Clark (2004a) as modified by arguments concerning chronology presented here. f Based on suggestions presented here concerning Cherla-phase platforms in Mounds 1 and 12, with the question marks referencing the suggestion that the Mound 14 ridge was an extensive platform with unknown function. g Lesure and Blake 2002. h Lesure 1995. b

12, and 14. If, as I have suggested here, these constructions constitute the first architectural platforms supporting nonresidential buildings— in other words, public buildings or temples— then the shifts in locations of construction were part of a structural change in the nature of ceremonialism. It is noteworthy as well that we see the fi rst (still modest) differences in access to imported goods, including personal ornaments, as well as striking differences between households in  the use of clay earspools. Those changes prompt me to tentatively call this stage “rank as an embodied quality” in Table 6.1. The growing differentiation involving personal ornaments raises the possibility that rank was increasingly taken to be an inseparable quality of personhood. I would suggest that it was the altered character of ceremonialism, with public ritual functions now removed from high-status residences and lodged in specially built structures, that released the final egalitarian-maintaining brakes (Blake and Clark 1999:58– 61) on the household accumulation of prestige items. With the creation first of a public versus private distinction and then a temple versus residence distinction, the people who were now commoners no longer held significant rights of surveillance over the private lives of the elite.

ACKNOWLEDGMENTS Field and laboratory work was made possible through grants to the author from the WennerGren Foundation for Anthropological Research, the H. John Heinz III Fund Grant Program for Latin American Archaeology, the University of California Institute for Mexico and the United States (UC MEXUS), a Fulbright (IIE) Fellowship, the Social Science Research Council, the New World Archaeological Foundation (NWAF, Brigham Young University), and the UCLA Academic Senate. John Clark, as director of the NWAF, generously provided facilities, equipment, financing, and other kinds of support too diverse to detail, and Michael Blake funded

144

emergent complexity

the 1993 season at Mound 12 under a grant from the Social Sciences and Humanities Research Council of Canada. Special thanks are due to Lorena Mirambell, Joaquín García Bárcena, Alejandro Martínez, and the Consejo de Arqueología of the Mexican Instituto Nacional de Antropología e Historia for permission to conduct the various seasons of work. Robert Rosenswig and Mary Pye commented on a previous draft of the paper.

REFERENCES Bell, Catherine 1992 Ritual Theory, Ritual Practice. Oxford University Press, New York. Bell, Catherine 1997 Ritual: Perspectives and Dimensions. Oxford University Press, New York. Blake, M., J. E. Clark, G. Michaels, M. W. Love, M. E. Pye, A. A. Demarest, and B. Arroyo 1995 Radiocarbon chronology for the late Archaic and Formative periods on the Pacific Coast of Southeastern Mesoamerica. Ancient Mesoamerica 6: 161– 183. Blake, Michael 1991 An Early Formative Chiefdom at Paso de la Amada, Chiapas, Mexico. In The Formation of Complex Society in Southeastern Mesoamerica, edited by W. R. Fowler, Jr., pp. 27–46. CRC Press, Boca Raton, FL. Blake, Michael, and John E. Clark 1999 The Emergence of Hereditary Inequality: The Case of Pacific Coastal Chiapas, Mexico. In Pacific Latin America in Prehistory: The Evolution of Archaic and Formative Cultures, edited by Michael Blake, pp. 55– 73. Washington State University Press, Pullman. Blake, Michael, and Vicki Feddema 1993 Paso de la Amada, un resumen de la excavaciones, 1990. In Primer Foro de Arqueología de Chiapas: CazadoresRecolectores-Pescadores y Agricultores Tempranos, pp. 75– 85. Serie Memorias, 4, Government of the State of Chiapas, Tuxtla Gutiérrez, Chiapas, Mexico. Blake, Michael, Richard G. Lesure, John E. Clark, Warren D. Hill, and Luis Barba 2006 The Residence of Power at Paso de la Amada, Mexico. In Palaces and Power in the Americas: From Peru to the Northwest Coast, edited by J. J. Christie and

P. J. Sarro, pp. 191–210. University of Texas Press, Austin. Bloch, Maurice 1974 Symbols, song, dance and features of articulation: Is religion an extreme form of traditional authority? Archives Européennes de Sociologie 15(1): 55– 81. Ceja Tenorio, Jorge Fausto 1985 Paso de la Amada: An Early Preclassic Site in the Soconusco, Chiapas. Papers of the New World Archaeological Foundation 49, Brigham Young University, Provo, UT. Clark, J. E. 1994 The Development of Early Formative Rank Societies in the Soconusco, Chiapas, Mexico. Ph.D. dissertation, Department of Anthropology, University of Michigan, Ann Arbor. Clark, J. E., and M. Blake 1994 The Power of Prestige: Competitive Generosity and the Emergence of Rank Societies in Lowland Mesoamerica. In Factional Competition and Political Development in the New World, edited by E. M. Brumfiel and J. Fox, pp. 17–30. Cambridge University Press, England. Clark, John E. 2004a Mesoamerica Goes Public: Early Ceremonial Centers, Leaders, and Communities. In Mesoamerican Archaeology, edited by Julia Hendon and Rosemary Joyce, pp. 43– 72. Blackwell, Oxford. Clark, John E. 2004b Surrounding the Sacred: Geometry and Design of Early Mound Groups as Meaning and Function. In Signs of Power: The Rise of Cultural Complexity in the Southeast, edited by Jon L. Gibson and Philip J. Carr, pp. 162–213. University of Alabama Press, Tuscaloosa. Connerton, Paul 1989 How Societies Remember. Cambridge University Press, New York. Hill, Warren D. 1999 Ballcourts, Competitive Games, and the Emergence of Complex Society. Ph.D. dissertation, University of British Columbia, Vancouver.

Hill, Warren D., Michael Blake, and John E. Clark 1998 Ball court design dates back 3,400 years. Nature 392: 878– 879. Hill, Warren, and John E. Clark 2001 Sports, gambling, and government: America’s first social compact? American Anthropologist 103: 331–345. Hodder, Ian 1992 The Haddenham Causewayed Enclosure—A Hermeneutic Circle. In Theory and Practice in Archaeology, pp. 213–240. Routledge, London. Hodder, Ian 1999 The Archaeological Process: An Introduction. Blackwell, Oxford. Lesure, Richard 1995 Paso de la Amada: Sociopolitical Dynamics in an Early Formative Community. Ph.D. dissertation, Department of Anthropology, University of Michigan, Ann Arbor. Lesure, Richard 1997 Early Formative platforms at Paso de La Amada, Chiapas, Mexico. Latin American Antiquity 8: 217–235. Lesure, Richard 1999 Platform architecture and activity patterns in an early Mesoamerican village in Chiapas, Mexico. Journal of Field Archaeology 26: 391–406. Lesure, R., and M. Blake 2002 Interpretive challenges in the study of early complexity: Economy, ritual and architecture at Paso de la Amada, Mexico. Journal of Anthropological Archaeology 21: 1–24. Love, Michael 2007 Recent research in the Southern Highlands and Pacific Coast of Mesoamerica. Journal of Archaeological Research 15: 275–328. Lowe, Gareth W. 1977 The Mixe-Zoque as competing neighbors of the early lowland Maya. In The Origins of Maya Civilization, edited by R. E. W. Adams, pp. 197–248. University of New Mexico Press, Albuquerque. Marcus, Joyce, and Kent V. Flannery 1996 Zapotec Civilization: How Urban Society Evolved in Mexico’s Oaxaca Valley. Thames and Hudson, London.

paso de la amada as a ceremonial center

145

SEVEN

A History of Disaster and Cultural Change in the Coatán River Drainage of the Soconusco, Chiapas, Mexico Gerardo Gutiérrez

I

n 1794 the capital of the Soconusco

province was moved from Escuintla to Tapachula. The reason for this change was a storm with strong winds that caused severe damage to Escuintla’s commerce and harm to its population (Pineda 1999:72 [1845]). These “strong winds” likely referenced a powerful hurricane that hit the area. Two hundred and eleven years later, another hurricane, Stan, provoked similar destruction in many parts of Tapachula. Given that hurricanes and other natural hazards are annual events in southern Mexico and Central America, one is led to ask if previous hurricanes caused similar damage to ancient population centers of the Soconusco. What other geological hazards could cause large-scale destruction in the region? How have human activities interacted with natural hazards, accelerating or preventing possible disasters? These questions are relevant because the Soconusco mangroves and coastal plain demonstrate clear evidence for the transition to sedentism (Voorhies 2004), as well as for the early emergence of complex societies (Clark 2004; Clark and Blake 1994). Thus it is important to

146

understand to what degree similar catastrophes may have affected these developments. This chapter is divided into three sections. First, I will address some theoretical approaches used to study disasters from the anthropological perspective, specifically focusing on disasters as “process” more than events (Hoffman and Oliver-Smith 2002:3). This approach sees disasters as the conjuncture of human activities and natural hazards embedded in natural and social systems that unfold as processes over time. This theoretical introduction is relevant for the second section of the chapter, where I will summarize the effects of Hurricane Stan in the Soconusco, specifically in the basin of the Coatán River. Through analysis of social and physical characteristics of this hurricane, we can understand how economic activities triggered the necessary conditions for a disaster to occur and how this process had a deep impact on all aspects of the regional society. Analysis of Hurricane Stan has provided an inductive model for detecting previous disasters in the region. It has evidenced how changes in the course of the river are frequent

and how they can rapidly cause the demise of previously important settlements and favor the emergence of others. Based on this analysis, in the third section of the chapter I will compare the sedimentation patterns left by Hurricane Stan with those that buried Cantón Corralito, a thriving archaeological site in the western margin of the Coatán River between 1900– 1000 b.c. (Cheetham et al. 2007:12). With this, I make an argument for the occurrence of an early disaster in the same hydrological basin later impacted by Hurricane Stan. Cantón Corralito has achieved some notoriety because of its strong connections to the early Olmec site of San Lorenzo Tenochtitlan in the Gulf of Mexico (Blomster, Neff, and Glascock 2005; Cheetham 2006; Cheetham et al. 2007; Pérez Suárez 2002:83; Pérez Suárez and Lesure 1998:33).

THEORETICAL APPROACHES TO THE STUDY OF DISASTERS1 Sociologists, meteorologists, geologists, and geographers are the specialists who have traditionally focused on studies of disasters and risk. This work has frequently occupied public officials or centers of study on natural disasters evaluating the impact on the productive infrastructure of an industrial society. These professionals are interested in the capacity for institutions to respond to stress and attend to emergency conditions. The guiding concept of these studies is the natural disaster itself, considered to be a misfortune that affects individuals and whole societies who happened to be in the wrong place at the wrong time. By labeling these catastrophic processes as “natural,” humans do not take responsibility for the making of the disaster, revealing a clear distinction between those processes caused by an external agent of society and those produced directly by human action. This has been the dominant philosophic approach in the study of disasters, portrayed masterfully by Voltaire in his poem Candide, blaming nature for the destruction of Lisbon during the earthquake of 1755. Nature (as controlled by God) is an agent at

work against humans (Olcina Cantos and AyalaCarcedo 2002:43). More recently, anthropologists and historians have engaged in the broader study of disasters (Oliver-Smith and Hoffman 2002:5). Their interest is based on understanding the intricate relationship between nature and society and how human action or inaction aggravates or facilitates the occurrence of disasters originally considered a strictly natural phenomenon. In this approach the theoretical tradition of anthropology has begun to question the false dichotomy between nature and society and has put front and center the role of human agency in the creation of disasters, which, for lack of a better word, are called “socionatural” (Oliver-Smith 2002:41; Wallace 1956). This perspective argues that disasters occur within the context of historically created patterns of vulnerability. The factors involved in the formation of these patterns of vulnerability are location, infrastructure, sociopolitical organization, distribution and production systems, and the ideology of a particular society (Oliver-Smith and Hoffman 2002:3). Philosophically, the origins of this approach would be found in Rousseau’s reply to Voltaire’s Candide. For Rousseau there is no such thing as an “evil” nature, and the causes of disasters lie in humans’ actions and choices in their search for never-ending “progress” and economic growth. It is humans who have constructed thousands of buildings and economic infrastructure along tectonic faults, in volcanic regions and alluvial fans and plains (Olcina Cantos and Ayala- Carcedo 2002:53). From this perspective, disasters and hazards can be converted into modular anthropological concepts. Operationally, Oliver-Smith and Hoffman (2002:4) define the concept of disaster as a process/event combining a potential destructive agent/force from the natural, modified, or built environment and a population in a socially and economically produced condition of vulnerability, resulting in a perceived disruption of the customary relative satisfactions of individual and social needs for physical survival, social order, and meaning.

disaster, cultural change in the coatán

147

This definition attempts to break down the dichotomy between nature and society by assuming that the destructive agent does not create the disaster; instead, it is the society that creates the appropriate conditions of risk and vulnerability that forge the socioeconomic contradictions within its par ticular social system. Hazard is defined by these same authors (Oliver-Smith and Hoffman 2002:4) as forces, conditions, or technologies that carry a potential for social, infrastructural, or environmental damage. A hazard can be a hurricane, earthquake, or avalanche; it can also be a nuclear facility or a socioeconomic practice, such as using pesticides. The issue of hazard further incorporates the way a society perceives the danger or dangers, either environmental and/or technological, that it faces and the ways it allows the danger to enter its calculation of risk.

The capacity for these hazards to cause damage can manifest itself in many forms: (1) the silent accumulation of lethal elements as a byproduct of human activities; (2) the constant removal of valuable resources from society that over time ends with devastated ecosystems and causes direct societal repercussions; or (3) the placement of human constructions in places where cyclical, large-scale physical forces occur, the natural byproducts of the geological, hydrological, and atmospheric processes of our planet. For most hazards, their capacity to harm derives from their potential to release suddenly large quantities of accumulated energy that exceed the scale and magnitude of human efforts to manipulate and control (Oliver-Smith and Hoffman 2002). Risk has been defined as the possibility that a specific territory or society is affected by a hazard of extraordinary magnitude, which then creates a major disruption in the usual organization and socioeconomic activities of a territory (Olcina Cantos and Ayala-Carcedo 2002:55). Risk always implies a random component, since it is the probability that an actual disaster will take place. We know that major hazards will eventually manifest themselves everywhere on the earth, but we do not know exactly when and how, so

148

emergent complexity

risk is the waiting period before fate strikes (Olcina Cantos and Ayala-Carcedo 2002:65). Vulnerability as a social concept is defined as “the characteristic of a person or group in terms of their capacity to anticipate, cope with, resist, and recover from the impact of a natural hazard” (Blaikie et al. 1994:9 as cited in Oliver-Smith 2002:28). Vulnerability has become the key concept in the social analysis of disasters, since it links the conditions of inequality and subordination in a society to very particular environmental forces (Oliver-Smith 2002:27–29). Disasters highlight the social structure of the affected group, in particular its capacity to overcome (or not) the tragedy, testing kinship relationships and alliances under stressful conditions. A disaster can result in social cohesion and unity or in conflict and disintegration by increasing segmentary tendencies within the social body. An anthropological approach to the study of disasters is precisely the analysis of an adaptation or maladaptation of a sociocultural system to its environment (Oliver-Smith and Hoffman 2002:8). In this line of research, it is necessary to study the knowledge and strategies of indigenous cultures to cope with recurrent hazards.

The Soconusco and Central America as a Region of Risk There is no space on earth completely free of hazards; however, there are some regions where cyclical hazards are more frequent and powerful than others. These spaces have been called territories or regions of risk, since their human occupations are exposed to recurrent catastrophes. Olcina Cantos and Ayala- Carcedo (2002:68) defined the concept of “risk-region” as a geographic space of known or discrete size where one or several hazards have an impact on a population, its settlements, and activities. The intersection between natural hazards and human groups in these regions of risk is such that constant disasters have put some checks on their social, political, and economic development. Central America, comprising nine countries from Panama to Southeastern Mexico,

presents par ticular intervallic hazards that have affected the social development of the groups living there. Therefore, it is possible to isolate it as a risk region on the basis of its more frequent and damaging disasters (Musset 1994; Preston 1996). Central American hazards can be classified in three groups: (1) those related to plate tectonics; (2) those related to cyclogenesis; and (3) those related to mass wasting.

Hazards Related to Central American Plate Tectonics In the region that constitutes Central America, two minor but active plates, the Cocos and the Caribbean, have a converging boundary along the Pacific coast of the Mexican state of Chiapas, Guatemala, El Salvador, Honduras, Nicaragua, Costa Rica, and northern Panama. The Cocos plate is created by the sea floor spreading along the East Pacific Rise and the Cocos ridge, specifically in the so- called Cocos-Nazca spreading system. The Cocos plate is pushed eastward, then collides and is dragged under the less dense overriding Caribbean plate at a  speed of 7.5 cm each year (Kerr et al. 1999; Meschede and Barckhausen 2000). The subducted portion of the plate heats up and traps superheated water under great pressure. In these conditions the subducted plate melts and produces ascending magma that has formed a continuous arc of at least eighty- one large volcanoes stretching from northern Panama to the border between Guatemala and Mexico. Hence, volcanic eruptions and constant earthquakes are the major geomorphological hazards for Central America, especially along the Pacific coast. Some examples of how these cyclical forces have affected human groups in this risk region are (1) the temporary abandonment of Izapa, Chiapas, around 38 a.d. by the Peléan-style eruption of the Tacaná Volcanic Complex (Macías et al. 2000:1244); (2) the destruction of Cerén village, El Salvador, by the eruption of the Loma Caldera volcano, circa 600 a.d. (Sheets 1992, 2002); (3) the vast damages suffered by Antigua, Guatemala, during the

earthquakes of 1717 (Peraldo and Montero 1996:297; Vázquez 1944:4:392) and subsequent earthquakes of 1773, which led to the shift in capitals from Antigua to Guatemala City in 1776 (Anderson 1908:484; García Acosta and Suárez Reynoso 1996:140); and (4) the destruction of Quetzaltenango, Guatemala, in 1902 by a major earthquake associated with the eruption of the Santa María Volcano (Anderson 1908:474; Gaitán 2005:1–11; García 1963:393–399).

Hazards Related to Cyclogenesis in Central America Tropical cyclones are the most powerful storms on earth and are potentially also the most destructive and violent tempests experienced by humans. Every year, from June to October, the Caribbean area and the Gulf of Mexico in the Atlantic Ocean, as well as the Pacific coast of Mexico and Central America, become a large region of cyclone genesis and hurricane landfalls (Christopherson 2006:231–236). In 1998, Hurricane Mitch left more than 12,000 dead in Central America, making it the deadliest Atlantic hurricane since tropical cyclones have been recorded. More historical and archaeological research is needed to assess the impact of ancient hurricanes on human groups and settlements from Southern Mexico to Panama.

Hazards Related to Mass Wasting in Central America Soil creeps, rockfalls, earthflows, debris avalanches, and mudflows have been labeled generically as mass-movement or mass-wasting processes because they all refer to the movement of a body of material, primarily propelled and controlled by gravity. Large debris flows have caused extreme loss of life and property damage all over the world (Christopherson 2006:418–425). Sudden mass movements of earth and water have also destroyed and buried an unknown number of human settlements throughout prehistory. The Nahuatl-speaking people of Mesoamerica had a special term for debris flow: Tetl Oatococ-

disaster, cultural change in the coatán

149

Quauitl Oatococ, which literally translates as “the current has washed away the rock and the tree.” This term was also used as a metaphor to refer to bad times with great hunger and disease (Sahagún 1969:258). The native sources of Mesoamerica also reported in the Codex TellerianoRemensis how an army of 1800 Aztec warriors was washed away by a debris flow in 1507, during its march to conquer the Mixtec area (Quiñones 1995:87, 229). The Central America Volcanic Arc, with its dozens of stratovolcanoes reaching heights of 3,000 m above sea level, provides the required variables for catastrophic mass movements to occur. The steep slopes of the volcanoes along the Pacific coast are cut by hundreds of deep gorges and ravines. Quaternary sedimentation has created large alluvial fans at the exit of these profound narrow canyons along the coast and in highland valleys. These gently sloping, cone-tofan-shaped landforms are highly exposed to major mass movements by a combination of volcanic, tectonic, or severe weather-related factors, and debris flows and flash floods actively operate on them (Larsen et al. 1999). Interestingly, since prehistoric times, human groups in Central America have settled in these alluvial fans, given their great fertility, availability of fresh water, gentle slopes, and temperate climate (Shook 1965). This intersection has proved fatal on many occasions. Perhaps the most famous historically recorded debris flow is the one that destroyed the first capital of Guatemala, Ciudad Vieja, in 1541. Reviewing the surviving accounts of the event, it seems that after several days of heavy rain, the water-saturated slopes of the Agua Volcano failed and created several mudslides that converged on their way down, creating a massive debris flow that killed some 600 people (Anderson 1908:484; Shillings et al. 2001; Vázquez, 1937:1:89– 92).

Interaction of Natural Hazards and Culture in the Soconusco The three main categories of natural hazards overlap each other, and one way or another they have affected social organization and cultural

150

emergent complexity

development in Central America, especially along its Pacific slope and coastal plain, as many studies now suggest (see articles in Rose et al. 2004). Cultural adaptations to natural hazards of this particular risk region must be identified and explained so that we can understand the processes and variables that operate in the resilience or collapse of local social groups during moments of stress (Redman 2005). Given the high volcanic risks, Sheets observes a direct correlation between the degree of social complexity and vulnerability in prehistoric Central America. The more complex social structures became, and they more they relied on elaborately built environments and intensive agriculture, the more vulnerable indigenous societies were to sudden massive stress (Sheets 2004:110). Adaptation within the Central America risk region and to its hazards can be observed in many ways. For example, during the 1902 earthquakes that affected Quetzaltenango, all stone and adobe houses were destroyed, whereas the native houses of wattle-and-daub covered with thatch were minimally affected (Anderson 1908:476). This type of house also resists the yearly floods of the swampy coasts and sustains minimal damages. The regional settlement pattern is another adaptation, since places that manage to escape major catastrophes for prolonged periods tend to benefit and thrive at the expenses of those that are located in less fortunate spots. It is likely that human groups in Central America underwent cyclical periods of adaptation and maladaptation based on changes in productive systems and their relationship to the natural environment. For example, each year flooding events associated with hurricanes and powerful storms are expected in the lower basins of the Soconusco. The “Cordonazo de San Francisco”2 marks the height of rainy season flooding, when populations in the bajos (lowlands subject to flooding) took refuge on high ground, abandoning homes they had used for most of the year. Before the severe deforestation of the Soconusco region, this flooding was perceived as a positive event because it brought a wealth of comestible resources—including

FIGURE 7.1 The Coatán River Basin and the Soconusco region

fish, eels, turtles, lizards, and shrimp— as the rising waters disturbed the normal habitats of local fauna. Nowadays, in contrast, this flooding is perceived as a disaster by government officials as well as by the inhabitants of the zone. Regional economic changes, from construction techniques to the productive infrastructure to the devastation of mangroves and coastal and piedmont forests, have all led to the regional culture suffering a maladaptation in a system that had formerly had a certain fitness.

HURRICANE STAN AND ITS EFFECTS ON THE COATÁN RIVER, SOCONUSCO, MEXICO The Physical Settings of the Coatán River Basin The Coatán River possesses a basin that captures rainwater from approximately 652 square kilometers (km2) of surface. It is an interna-

tional basin shared by Mexico and Guatemala (Figure 7.1). The upper basin of the Coatán River, with its principal headwaters, lies in the municipality of Tacaná, department of San Marcos, Guatemala. The middle and lower basin are located in Mexico, in the municipalities of Cacahoatan, Tapachula, and Mazatán. According to Instituto Nacional de Estadística y Geografía (INEGI) maps, the geological matrix of the upper basin of the Coatán is composed primarily of an enormous batholith of granite and granodiorite from the Permian period of the Paleozoic Era (some 251 million years before the present). This batholith was partially covered by tuff, an extrusive igneous rock from the Tertiary period of the Cenozoic Era (between 5 and 1.6 million years before the present). The lower basin on the coastal plain was more recently created, principally in the Quaternary period (1.8 million year ago to the present), by alluvial deposition of eroded sediments from the upper basin transported to

disaster, cultural change in the coatán

151

the coast. The transport of material by the rivers that drained the Tacaná Volcanic Complex has formed a large alluvial apron from the Pompuapa to Suchiate Rivers. Large sand bars extend along the coast from the MexicanGuatemalan border to Barra San Simón, evidence of the constant renewal of sediments. Pluvial precipitation is high in this zone, especially along the slopes of the Sierra Madre oriented to the Pacific Ocean, with up to 4,000 mm of rain per year along the altitudinal tier between 400 and 1800 m above sea level (INEGI maps). At elevations higher than 2,400 m above sea level, precipitation rates begin to diminish to 2,800 mm per year. Along the coastline, precipitation rates are also low: 1,800 mm per year. In the Soconusco, seasonality is very marked, with a clear division between dry and rainy seasons. This clearly affected regional settlement patterns, since the area from the coastline inland up to 50 m above sea level has precipitation during the dry season ranging only between 75 to 150 mm (a maize plant needs at least 700 mm of rainfall throughout its development). This would lead to human settlements locating near permanent sources of fresh water, in par ticular the principal channel of the Coatán River. Maximal vegetation coverage in this zone occurs up to 2,000 m above sea level, with dense forests in humid temperate zones and a great variety of plant species (Richter 1993:13– 34). Pine and oak forests are found from 2,300 m above sea level to the peak of Tacaná. Pine has commercial value as planks, while oak is used to produce carbon (Cortina 2006:4). In the past, carbon production could have been important and perhaps was traded into the lowlands, while turpentine from pine could have been used for medicinal purposes. Until the nineteenth century the coastal plain was covered with perennial tropical rainforest, with trees 25 m in height, including rubber (Castilla elastica), ramon (Brosimum alicastrum), various species of burseras, mamey (Pouteria zapota), cajete (Dendropanax arboreus), chaperno (Ulmus

152

emergent complexity

mexicana), guava (Terminalia oblonga), and pochote (Ceiba petandra). Unfortunately, this vegetation has been devastated and substituted by grasses, seasonal and irrigation agriculture, and secondary vegetation. Along the coast, mangroves are still found, primarily two species: red mangrove (Rhizophora mangle) and white mangrove (Laguncularia racemosa).

The Changing Channels of the Coatán River During the investigation of Hurricane Stan, we became interested in the changing course of the Coatán River, especially in the lower basin. The coastal plain of the Soconusco had assumed its form during the Quaternary period, but especially during the Holocene epoch—that is, within the last 10,000 years. The Coatán River and all of its affluents in the upper basin are the principal agents of erosion on the northern and western slopes of the Tacaná Volcanic Complex and have transported millions of cubic meters from the summits of the sierra down to the coastal plain. In the middle and upper basin, the river has significant energy from high precipitation rates and the steepness of the slopes. Hence, it is capable of dragging and moving rocks of considerable size, but on exiting the narrow valley that separates the piedmont from the mountain slopes, it loses energy and begins depositing the materials it has thus far carried along. Large boulders remain trapped in the narrow canyon created by the river, which is located between the communities of Salvador Urbina and Veinte de Noviembre; the mediumsize boulders and gravel deposit themselves between Tapachula and Álvaro Obregón. From Álvaro Obregón to the mouth of the river into the Pacific Ocean, the river has less energy and is capable of transporting only sand, silt, and clays, which deposit differentially along the lower basin, creating a very dynamic delta (Clark 1994:49). This process has been elevating the coastal plain millimeter by millimeter and sometimes by meters, and it has caused permanent changes in the river basin. These

FIGURE 7.2 Delta of the Coatán River and its main channels during the Holocene period

changes are important, since the lack of rainfall on the coastal plain for the six months of the dry season leads human groups to follow (and settle) according to the movements of the river. Analyzing the ancient oxbow lakes in the lower alluvial fan of the Coatán River, I suggest at least seven major changes in the Coatán River channel, all of which indicate the river is moving progressively from north to south along its most recent delta (Figure 7.2). Channels 1 and 2 Two of the first channels of the Coatán seemed to have drained directly into the Encrucijada Mangrove Biosphere Reserve (also known as Pampa Cantileña). The first channel drained into Pampa Cantileña some 13 km to the northwest of the current mouth of the Coatán, south of the town of Vicente Guerrero. The second channel drained 10 km to the northwest of the current mouth of the Coatán, north of El Cocal.

These two channels crossed through Cantón Corralito and continued just north of Buenos Aires. Until we do further geomorphological coring, our hypothesis is that these two channels of the Coatán River were active during the Early and Middle Archaic periods. Channel 3 At some moment during the Late Archaic, these former channels changed and deviated to the south to Channel 4 (see below). This change of course was located 1.5 km to the west of the current community of Buenos Aires. What is interesting in this stage is that an ephemeral channel was formed to the south of the archaeological site of Paso de la Amada, famous for its early chief’s house and ball court (Blake et al. 2006; Clark 2004). The pools used to store water around Paso de la Amada evidence considerable human modifications, which would indicate an adaptation to an intermittent

disaster, cultural change in the coatán

153

flow that received abundant water only during the rainy season (see figures 27 and 28 in Clark 1994:127– 128). Channel 4 This channel is difficult to understand because of the large-scale canals built by the Mexican government in the area over the last fifty years. This channel used to pass to the south of Adolfo Ruíz Cortínez, coinciding at some points with the La Calentura estuary. Perhaps this basin is contemporaneous with the channel that periodically passed near Paso de la Amada. Clark has proposed that it was the last former channel of the Coatán River before it shifted to its present location (Clark 1994:166). As I will suggest later in this chapter, this channel may have become clogged by the sedimentation event that buried Cantón Corralito between 1200 b.c. and 1000 b.c. Perhaps it was at this moment that all the intermittent drainages of the river leading toward the channels running north, west, and south of Buenos Aires were definitively blocked.

nel (Channel 7) from Guanacastal to north of Balneario Barra de San Simón. At some point this short-lived channel ceased to function and the river moved to its current channel that runs south of La Victoria. After Hurricane Stan the channel changed its route again, moving more to the southeast. The Mexican government, acting against nature in order to preserve property boundaries, decided to return the river basin to its former position by dredging the principal channel and removing tons of sediment. This human intention to impede the natural movement of the Coatán is one of the largest anthropogenic interventions in the course of this river, at least in terms of the absolute movement of volumes of earth by using large-scale, earth-moving equipment. Nonetheless, it is not the fi rst anthropogenic intervention, since the bajos and river channel near Paso de la Amada demonstrate evidence of pond and pool modification and water management as early as the Locona phase (1700– 1500 b.c.) (Clark 2004:58– 59).

Channel 5

Hurricane Stan

This channel is marked by an intermittent current of the Arroyo Seco, a canal that runs from Cantón Corralito to the town of Mazatán. This, together with Channel 6 (see the next paragraph), would have formed at the end of the Cuadros phase when Cantón Corralito was buried by a large sedimentation event. It is unclear if this basin reincorporated itself into Channel 4 north of Mazatán or simply turned to the south and followed the current channel of the Coatán River (Channel 6).

Stan was the eleventh storm of the 2005 Atlantic hurricane season. It was embedded in a larger nontropical system of rainstorms that dropped torrential rains on Guatemala, El Salvador, and southern Mexico, causing floods and mudslides that led to some 2,000 deaths throughout Central America. On October 3, 4, and 5, 2005, many areas of southwestern Chiapas were inundated by the overflow of the Coatán River that was caused by extraordinary torrential rains fi lling its basin. Using the resources of the Tropical Rainfall Measur ing Mission Online Visualization and Analysis System (TOVAS) of NASA,3 I have estimated that during Hurricane Stan, the Coatán River Basin accumulated circa 235 million m3 of rainfall from October 1 to October 7, 2005. This figure would provide the 8.2 million inhabitants of New York City with water for two months. This rain washed away several million cubic meters of water-saturated soil. The slopes of

Channel 6 This is the current channel as seen in INEGI maps. This basin configuration was in place during the construction and occupation of the large sites of Guanacastal and Santa Cruz (El Triunfo), downstream from Mazatán (Clark 1994:158). Given the presence of a series of oxbows to the north of La Victoria, it is possible that the river ran through an ephemeral chan-

154

emergent complexity

FIGURE 7.3 Debris flows in the alluvial fan of the Coatán River after Hurricane Stan

the Tacaná Volcano failed, and dozens of landslides merged into large debris flows. Boulders, mud, and trees were carried by floodwaters downriver to the coastal plain and covered dozens of villages, towns, and sections of the city of Tapachula, destroying houses and agricultural lands (Figure 7.3). Communication and transportation routes suffered major damage, isolating the local population for days. More than 1,700 families in the municipality of Mazatán watched with horror as the muddy water engulfed their homes and buried their belongings (Programa Emergente Stan 2006). Hunger and infectious diseases became a serious threat for the victims, provoking uncertainty and social unrest. Studying this disaster process while it was unfolding became a priority for many research institutions with interests in Chiapas, including Centro de Investigaciones y Estudios Superiores en Antropologia Social (CIESAS) and the New World Archaeological Foundation (NWAF), which provided assistance in creating an ethnography of the event as well as a spatial analysis of the physical and socioeconomic factors related to this specific disaster.

The Effects of Hurricane Stan and Their Analysis My strategy for studying the Hurricane Stan disaster was to use qualitative and quantitative information to interpret and analyze all pertinent data, but to focus primarily on 1. Areas where sediments were deposited in the lower basin of the Coatán River; 2. Sedimentation stratigraphy left by the hurricane; 3. The volume of sediments deposited in the lower basin of the Coatán River; and 4. Long-term adaptive strategies of the population and future archaeological contexts created by Hurricane Stan. To obtain these data we needed to understand the scale of the larger sedimentation patterns as influenced by both human action and by the release of cyclical energy from extreme hydrometeorological and volcanic events in the region. This would help describe the geomorphological evolution of the Coatán River and place the developments of human societies in the Soconusco within this regional environmental framework.

disaster, cultural change in the coatán

155

To estimate the surface area affected by Hurricane Stan, aerial photos were used from before and after the event. Since satellite images of the study area were not available in the first weeks after the disaster, we rented a small aircraft to obtain high-resolution digital photos to measure the area affected by flooding and sedimentation. After being georeferenced, these images were compared to pre-Hurricane Stan images (Ortofotos Digitales, INEGI 1996), making it possible to evaluate specific damages to housing, fields, and infrastructure. From this analysis we found that the area directly affected by sedimentation in the lower basin of the Coatán River measured circa 69 square km, which makes up about 11 percent of the total catchment area of the Coatán River Basin. The area of the actual floodwaters is larger than this figure because many streams, channels, and so forth merged in the pampas, swamps, and lowlands of the coastal plain, making it difficult to obtain an accurate number, but it probably reached some 1,500 km2 of inundated lands between the Suchiate and Tilapia Rivers. The second goal was to understand the stratigraphic sequence of several recent catastrophes reported for the area. In one century, three major natural events have impacted the Chiapas coast and created what we call a disaster. These were the eruption of the Santa María Volcano in 1902; Hurricane Mitch in 19984; and Hurricane Stan in 2005. To investigate the kind of strata left by these events, a 2 × 2 m soil test pit was excavated. This unit was placed in Cantón de Guanacastal at coordinates 557490E, 1639653N (UTM [Universial Transverse Mercator], WGS84, Zone 15), located 170 m north of the Coatán River. The unit was excavated to 1.30 m. A brief description of its stratigraphy follows, beginning at the bottom, taking as a chronological marker the ash sediment of the 1902 Santa María event.5 The eruption of the Santa Maria Volcano left a very clear 10 cm stratum of ash (Figure 7.4). This was followed by brown soil formation, which accumulated 30 cm over a period of ninety-six years. Then, two distinct layers were

156

emergent complexity

created by debris flows from the two different hurricanes in a period of seven years. The first was caused by a meteorological event that took place either in late September or late October (most likely Mitch) that left a 30 cm layer of sediments. The second event was deposited by Stan, creating a 45 cm layer of sediments. The sedimentation pattern of the two hurricanes is very interesting, as each one leaves two identifiable layers, one composed of a clay loam mixed with sands (lower layer) and the other a finer sand and silt (upper layer). During the sevenyear period between Mitch and Stan, there was a 3 cm formation of rich, dark organic soil that contained an abundance of cacao and banana leaves in decomposition. With this understanding of the deposition patterns for Hurricane Stan, 258 auger cores were drilled from La Victoria to the Pandora Gravel in Alvaro Obregón. Although my team and I drilled a core approximately every 400 m on average, the sampling procedure was opportunistic, since some landowners did not permit access to their properties or some areas were simply unreachable because of the destruction of roads and bridges. These cores provided an approximate minimum volume of 28,234,767 m3 of sedimentation left by Hurricane Stan in the Coatán River Basin. This estimate does not capture the total volume of materials moved by the heavy rains of the hurricane, but it does measure the sediments left by floodwaters from the town of Alvaro Obregón to the mouth of the Coatán River. When we began our coring study (in April 2006), the sediment between the towns of Viva Mexico and Tapachula had already been disturbed by heavy machinery trying to reopen the former river channel. Thanks to a set of excellent oblique pictures taken by the office of the Protección Civil of Chiapas during the first week of the disaster, it was possible to estimate the thickness of the layer of sediments deposited by Hurricane Stan in some control points between Viva Mexico to the hydrological apex of the Coatán alluvial fan in Fraccionamiento Margaritas, a municipality of Tapachula. This estimate was based on how

FIGURE 7.4 Sedimentation events caused by major volcanic and hydrometeorological processes in the Soconusco, 1902–2005

deep the cars, houses, and bridges were buried by boulders and sediments (see Figure 7.3). With these additional data, I estimated a corrected figure of some 46,252,559 m3 of sediments brought by the torrents from the upper basin to the alluvial fan of the Coatán River. This is equivalent to 6,607,508 dump trucks with a cargo capacity of 7 m3 each (Figure 7.5). This significant deposition clearly affected regional topography, and currently I am in the process of estimating the changes in water flow patterns for the coastal plain in the Mazatán area. Despite the efforts of the Mexican government to bring the river channel back to its preStan position, the river channel has moved slightly southward and elevated its bottom.

This is creating problems for Cantón El Aguacate, a community that is now experiencing floods during heavy rains, which previously had not been the case. A similar flooding situation is being experienced on the outskirts of Ejido Buenos Aires to the north of the Coatán River. The prevailing idea when I began this project was that landslides and flooding were caused directly by an increase in rainfall due to global climatic change. Nonetheless, during the year of Hurricane Stan, I was able to estimate that it rained less than normal. According to the INEGI map of climate effects, historically the basin of the Coatán River captures 1.389 km3 of water during the rainy season

disaster, cultural change in the coatán

157

FIGURE 7.5 Area and volume of sedimentation in the lower basin of the Coatán River after Hurricane Stan

(May through October), but according to TOVAS,6 during the year of Stan the basin accumulated only 1.303 km3 of rain from May to October—representing an approximate deficit of 86.6 million m3 of rain. Hence, seasonally excess rainfall was not the primary explanation behind the disaster. What are the cultural implications to infer from this? Seventy interviews with the Mazatán population revealed that in the past heavy rains and annual flooding were part of daily life and that their society and material culture were adapted to these episodes. Previously, rains and flooding were perceived not as a catastrophe but as times of abundance; the opportunities for hunting and collecting local fauna greatly increased from the disturbance of their habitats. Iguanas and turtles were caught by the thousands, as well as fish and shrimp. Older infor mants recall incremental changes as local agriculture became more intensive and

158

emergent complexity

industrialized. Deforestation and single- crop agriculture in the coastal lowlands exposed large tracts of land to surface erosion. Additionally, the use of pesticides killed much of the fauna formerly collected during the rainy season. Traditional house construction was abandoned for concrete houses. Salaried jobs replaced subsistence agriculture. Similar processes were likewise occurring in the upper basin of the Coatán River. Unfortunately here, the lack of salaried jobs caused the pine and oak forests to be converted into cash crops. The denuded landscape was exposed to tropical sun and torrential rains like those that took place during Hurricane Stan, with an estimated twenty inches of accumulated rainfall per square meter in a one-week period. This situation provides an ecological model for understanding the interplay of factors that created the conditions for a disaster to occur:

1. Human destruction of the forest in the upper basin causes loss of topsoil; 2. During torrential storms, the eroded topsoil becomes mudslides that transport millions of metric tons of material down the river basin; and 3. In the past, the mangrove and swamp system of the lowlands, together with the tropical forest, would have mitigated the destructive force of a similar mudslide, acting as detention ponds and living walls, but since this system is now compromised by increasing economic activities, the impact on the human settlements is catastrophic. Activities generated within the context of an international economy have resulted in severe damage to the regional ecology of the Coatán River Basin, perhaps to the point of collapse. The stratigraphy in the lower basin presents two massive events of sedimentation within seven years (1998–2005). So as the millions of metric tons of topsoil are washed away in the upper basin, it increases the vulnerability of society in the lower basin in the face of everpresent natural events. The flooding and sedimentation of Hurricane Stan also produced a vast number of future archaeological contexts, and literally turned thousands of personal belongings into archaeological objects buried by millions of cubic meters of dirt. It is a sad reminder that behind each one of these objects there is a personal tragedy. In addition to the washed-away belongings, the mudslide created Pompeii–like archaeological contexts. Tools that would never have been left behind by their owners were often located in the exact spot where they were last used. These objects vary from cars to refrigerators as well as agricultural tools. Most interesting is the fact that few people are leaving their properties, despite the funding and in some cases new houses and land in other locations provided by the Mexican government. Instead, most people are staying and rebuilding over their old house

sites, creating a process of archaeological occupation and mound formation. Disasters are usually seen as moments of destruction, as the disintegration of orderly patterns binding together many aspects of modern material life. But they are also moments of social and material renewal: washed-away objects are replaced by new ones and destroyed houses are rebuilt, creating a totally new occupation. It is not a new migration with a different material culture; it is the same people renovating their lives and social ties. Disasters are also a moment of innovation. As an infor mant told us, “If the last storm brought me thirty centimeters of mud, then I will build a house over a platform of one meter so I can always stay one step ahead of the disaster.”

CANTÓN CORRALITO: THE EARLIEST REPORTED DISASTER IN MESOAMERICA Rock avalanches, debris flows, flooding, and volcanic lava and ash flows create very specific stratigraphic signatures that archaeologists can detect when one of these natural hazards has intersected with a human settlement in a catastrophic way (Manzanilla 1997:34–42). Therefore, the question to be addressed here is whether the archaeological record can reveal the role of human agency in the occurrence of ancient disasters in the same way that we have seen for Hurricane Stan. I will analyze the case of Cantón Corralito, probably the earliest disaster ever recorded for Mesoamérica. The archaeological site of Cantón Corralito was reported by John Clark as part of his archaeological survey in the municipality of Mazatán (Clark 1994:140– 162). The site was originally defined as two separate settlements: San Carlos (Mz-44) and Cantón Corralito (Mz28). Later these two were merged into one called Cantón Corralito (Mz-257). By evaluating the stratigraphy of water wells dug in the area, Clark identified the locale as an important Cuadros-phase (1300– 1200 b.c.) site; using also the ceramics, figurines, and sculpture found in the Mazatán area, he posited Cantón Corralito

disaster, cultural change in the coatán

159

as a community linked to the Gulf Coast center of San Lorenzo (Clark 1997; Clark and Pye 2000:234–236). Clark also noticed that much of the site was covered by a thick deposit of sterile sand. Following Clark’s initial work, Tomás Pérez Suárez and Richard G. Lesure continued the exploration of Cantón Corralito in 1997 (Pérez and Lesure 1998). Pérez Suárez excavated eight test pits to depths of almost 5 m on both sides of the main road connecting Álvaro Obregón to Mazatán (Pérez Suárez 2002:78). These test pits were complementary to those excavated by Clark in the mound of San Carlos, south of this same road (Clark 1994:140– 158). During his explorations, Pérez Suárez identified a complex stratigraphy with two main cultural deposits clearly separated by sedimentation events. One contained ceramics from the Cherla phase (1400– 1300 b.c.) and was covered by a thin layer of fi ne sand; above this was a  thick deposit of Cuadros-phase ceramics embedded in lenses of silt, fine sand, and clay.7 Cuadros-phase deposits were covered again by a thick layer of what Pérez Suárez reported as “fine yellow sand,” what the locals call arena roja, and what I will refer to as the “red sand” layer.8 In test pits 6 and 7, Pérez identified a few potsherds of the Jocotal phase (1200– 1000 b.c.) in the organic soil that formed between the red sand layer and the ash layer of the Santa María volcanic eruption (Pérez Suárez and Lesure 1998:16–24). From the materials recovered in 1997, Pérez Suárez confirmed the presence of Calzadas Carved and Limón Incised ceramics. This strengthened Clark’s initial hypothesis of a strong connection between Cantón Corralito and San Lorenzo in the Gulf Coast (Clark 1997; Lesure 2000; Pérez Suárez 2002:80). In 2004, under the auspices of Clark and the NWAF, David Cheetham undertook a more extensive exploration of Cantón Corralito, through the excavation of two long trenches, twenty-two test pits aligned in three transects, and more than ten scattered test pits. With this exploration Cheetham again corroborated the findings of the 1997 field sea-

160

emergent complexity

son, confirming a strong connection between Cantón Corralito and San Lorenzo beginning in the Cherla phase, but especially during the Cuadros phase (Cheetham 2006; Cheetham et al. 2007). He also observed the alluvial sediment layer covering the Cuadros-phase deposits9 and suggested, as had Pérez Suárez, that this red sand layer had buried the site during a flooding episode of the Coatán River (Cheetham et al. 2007:49– 51; Pérez Suárez 2002:91). In other publications, Cheetham has suggested that “one fateful stormy day about 1000 b.c. [uncalibrated C14 years] the nearby Coatán River began to rise. Before long the town was engulfed by the sands of its rapidly shifting channel, the residents no doubt fleeing to safer ground and a new life. Thus ended the occupation of what may well be the Americas’ first true colony” (Cheetham 2006:46). Although I ultimately believe this statement will be proved correct, the primary causes and the physical and social process behind this par ticular disaster have yet to be studied before we can refer to Cantón Corralito as an “aquatic version of Pompeii” (Cheetham 2007). Based on all the reports of the Cantón Corralito cultural deposits (Cheetham et al. 2007; Clark 1994; Pérez Suárez and Lesure 1998) and accidental findings of archaeological objects by the local population living in the modern settlement, it is possible to estimate that Cantón Corralito reached a site size of at least 31 hectares (ha) during the Cuadros phase.10 I observed in a nearby water canal that another segment of the site, or perhaps a different coetaneous site, is located 1 km to the north of the main concentration of archaeological deposits defined as Cantón Corralito. In this Santa Ana site (after the ranch where it is located), some ranchers have observed potsherds similar to those of Cantón Corralito to a depth of 1.50 m, precisely under a thick layer of the 50 cm of red sand. Based on the ranchers’ information, I estimate that Santa Ana’s archaeological deposits cover an area of at least 7 ha (Figure 7.6). These two sites seem to have been separated by an old channel of the Coatán River detected

FIGURE 7.6 Location of Cantón Corralito and Santa Ana sites

during the explorations of 1997 and 2004 (Cheetham et al. 2007:95; Pérez Suárez and Lesure 1998:19). Of the estimated 31 ha of Cantón Corralito, 47 percent (14.6 ha) of the site was engulfed by the red sand sedimentation event, creating a time capsule of preserved Cuadros and Cherla middens and burials. Unfortunately, the paved road connecting Mazatán to Alvaro Obregón was constructed over 2 ha of the buried archaeological deposits. The other 53 percent of the settlement was on higher ground, and the debris flow was deflected by the Barra-Locona mound of San Carlos (Clark 1994:142) and an associated platform built during the Cherla and Cuadros phases (Cheetham et al. 2007:100). The summit of another mound located around water well “D” and in test pit 18 of the Northern Transect excavated by Cheetham also escaped flooding and sedimentation (Cheetham et al. 2007:115). That portion of the site not covered by the sudden sedimentation event has been exposed to intensive erosion, bioturbation, looting, and construction of a modern settlement

over the culturally created high ground. Hence, the Cuadros and Cherla contexts are poorly preserved south of the road, but their older phases of occupation (Locona and Ocós) are in better shape (Clark 1994:140– 158). The estimated 7 ha of the Santa Ana site were completely buried, and archaeological research needs to be undertaken there to understand its development and connection to Cantón Corralito. If in fact this large flooding and sedimentation event ended the development of Cantón Corralito and its contacts with San Lorenzo, it would be the earliest disaster reported in the archaeological bibliography of Mesoamerica. This disaster would be earlier than 1. The eruption of the Xitle volcano that buried Cuicuilco during the Late Formative (Córdova, Martín del Pozo, and López Camacho 1994; Sanders et al. 1979); 2. The cultural deposits of Bezuapan and the Tres Zapotes region buried by the protoClassic eruption of Nixtamalapan (JaimeRiverón and Pool 2009:139);

disaster, cultural change in the coatán

161

3. The Ilopango and Loma Caldera eruptions in El Salvador (Sheets 1992, 2002, 2004) during the Early and Middle Classic periods; and 4. The burial of Tetimpa by the Popocatepetl volcanic ash flows in the Epiclassic period (Plunkett and Uruñuela 1998).

Drawing a parallel of the red sand debris flow event to the sedimentation patterns of Hurricane Stan, more than 30 km2 of forest and swamps north of the Coatán River, between the El Carmen and Buenos Aires communities, must have become wastelands for years. People would have moved to higher ground southeast of Cantón Corralito, favoring the development of the site of Ojo de Agua. After this sedimentation event, both Pérez and Cheetham report only a minor occupation on the high ground area of Cantón Corralito during the subsequent Jocotal phase (Cheetham et al. 2007:104; Pérez Suárez and Lesure 1998).

The Red Sand Catastrophe The sedimentation episode that blocked the northern channels of the Coatán River (see the previous section), besides being a major geomorphological event, seems to have also been a human disaster for the societies that lived at the end of the Cuadros phase, especially for those occupying Cantón Corralito and Santa Ana. During our analysis of Hurricane Stan, we detected an area of high ground that had not been affected by the flooding and debris flow. This area attracted my attention as a possible safe ground to build refuges for future flooding events; however, by analyzing aerial photographs and inspecting the area, I realized that this high ground was actually covering older channels of the Coatán River. When I obtained cores from this high ground area, I found a homogenous, medium- to fine-grained sand with a color ranging from intense red to light orange.11 To understand the stratigraphic deposition of the red sand layer and the blocking of

162

emergent complexity

the old river channels, a 2 × 2 m soil test pit was excavated at the coordinates 560360E, 1648253N (UTM, WGS84, Zone 15 North), 2 k east of the 90o turn in the Coatán River, close to the El Carmen Ranch. During excavation of the first 10 cm, I observed that a light brown soil had formed during the last one hundred years. Immediately beneath this soil was a 12 cm layer of ash accumulated during the eruption of the Santa María Volcano (Figure 7.7). Below the Santa María ash is a 35 cm layer of a rich, dark organic soil formed perhaps during hundreds of years under a forest canopy. Then a thick layer of 75 cm of the sterile red sand follows. This red sand then typically covers either a loam composed of a mixture of clay, sand, and silt or a layer of coarser gray fluvial sand. Applying the same sampling technique as that used for the Hurricane Stan coring, I drilled sixty cores along the northern ancient channels of the Coatán River, partially defining the edges, area, and volume of this red sand deposit. On learning that this layer was the same one that covered Cantón Corralito, it was possible to tentatively date it as an event taking place between 1200 b.c. and 1000 b.c. I also had the opportunity to use the stratigraphic information from archaeological explorations by John Clark of the NWAF-BYU in Mazatán to refine our current knowledge of this par ticular sedimentation event (Cheetham et al. 2007; Clark 1994; Pérez Suárez and Lesure 1998). By using these archaeological data, it was possible to increase the number of stratigraphic columns to 155, revealing more precisely the presence or absence of the red sand layer over an area of 70 km2. Unfortunately, the Hurricane Stan debris flow prevented a complete analysis of this earlier event, since Stan eroded or deeply buried the older sediments of the basin. Therefore currently it is not possible to define all of the red sand’s edges, its volume, and its trajectory between Tapachula and the Perla del Coatán.12 We are lacking data for the area east and south of the main road between Mazatán and Alvaro Obregón. Despite these problems, I have man-

FIGURE 7.7 Sedimentation events caused by major volcanic and hydrometeorological processes in the Soconusco, 1200 b.c.–2009

aged to estimate that an area of 30 km2 was covered by the red sand deposit to the north and west of the Mazatán-Alvaro Obregón road, with an approximate volume of 16,622,000 m3 (Figure 7.8). In future work, we hope to learn more about this red sand sedimentation event and its patterns, but it is likely that it was a debris flow event at least as large as that triggered by Hurricane Stan. I propose two potential hypotheses to explain this red sand event and its catastrophic process. Both hypotheses are not necessarily exclusive, and perhaps future research will be able to prove that they are complementary. Hypothesis 1 I suggest that during the Early Formative period (1900– 900 b.c.) the upper basin of the Coatán River was experiencing deforestation,

perhaps caused by incipient agriculture, which accelerated erosion rates in the upper basin and eventually led to a large process of mass wasting similar to the debris flow observed during Hurricane Stan. This extreme event destroyed and buried a large portion of Cantón Corralito and Santa Ana archaeological sites and marked the end of the social developments of the Cuadros phase in the Mazatán area. At first glance it would seem absurd to propose there was such an impact on a preindustrial or prehistoric system. Nonetheless, the Soconusco zone has witnessed plant cultivation activities since at least the Late Archaic (3500– 1900 b.c.) (Voorhies and Metcalfe 2007:176). I hypothesize here (as Lowe first did in 1971) that tubers were the first cultigens, followed by maize, and that slash and burn techniques caused a major impact on the pristine ecologi-

disaster, cultural change in the coatán

163

FIGURE 7.8 Area and volume of “red sand” sediment in the lower basin of the Coatán river, 1200– 1000 b.c.

cal systems of the Coatán River Basin between 2000 b.c. and 1000 b.c. All ecological systems possess a certain resistance to the impacts of human activities; however, these activities can begin a cumulative process of damage that, with time, reaches a critical level. In this case they resulted in erosion that led to a large debris flow. Perhaps this accumulation did not arrive at critical levels in the Coatán River Basin until it manifested itself in the Cuadros phase and buried the sites of Cantón Corralito and Santa Ana. This is not to say that erosion did not exist before the introduction of maize in the region, but undoubtedly the cultivation of maize accelerated erosion processes. Deforestation and erosion could have increased because of modifications in social organization that may have taken place in the Cuadros phase with the establishment of a foreign colony from San

164

emergent complexity

Lorenzo (Cheetham 2006:46). Perhaps the creation of a rudimentary tributary system would have increased the population’s use of resources in the upper basin. For this hypothesis to be viable, it mandates occupation in the upper basin of the Coatán River; unfortunately, this zone has not been systematically investigated. Other highland areas of Mesoamerica had important occupations in the Early Formative era, but currently we have no understanding as to what was occurring in the piedmont zone of the Soconusco at this time apart from secondary evidence in the fi lls of Mound 30 at Izapa (Ekholm 1969). If we consider the model of verticality and the control of natural resources as it operated in pre-Hispanic Peru, we see that Soconusco coastal populations could likewise have been interested in resources from the upper basin. If such were the case, social complexity

in the upper basin would be correlated to groups in the lower basin. I suggest here that a period of maladaptation occurred in the Soconusco during the Cuadros phase. The suggestion is derived from changes in the regional political organization introduced by external groups (immigrants from the Gulf Coast). For some reason the new order introduced by Olmec elites may have provoked an increase in productive activities in the Soconusco that manifested itself in the deforestation of the volcanic slopes of the upper basin of the Coatán river. This deforestation would have increased the vulnerability of Soconusco societies to a hydrometeorological phenomenon, which could have led to a shift in the Coatán River course, ultimately burying the archaeological site of Cantón Corralito, the primary center of the Cuadros phase. Soconusco society seemingly readapted itself after this event, reconstructing a major settlement on the other side of the river at Ojo de Agua during the Jocotal period (Pinkowski 2006: 46– 49; see Pye, Hodgson, and Clark’s Chapter 10 in this book). Hypothesis 2 Recent petrographic and geochemical analysis of Tacaná’s lava deposits suggests eruptive activity at 2905 ± 75 years b.p. (conventional date in Mora et al. 2004:336). This would correspond to the Cuadros-phase ceramic sequence of the Mazatán area, when the settlement of Cantón Corralito was active. It is possible that between 1206 b.c. and 986 b.c. (calibrated date ± 1s), a block-and-ash-flow volcanic eruption of the Tacaná Volcano burned the forests and clogged the channels of the upper basin of the Coatán River with loose pyroclastic material. During a subsequent hurricane season, millions of cubic meters of volcanic ash were carried away by a lahar, which eventually buried and destroyed the site of Cantón Corralito and other contemporaneous settlements along an ancient channel of the Coatán River. Geochemical analysis of the materials in the red sand layer is now being undertaken by volca-

nologists from the department of Geophysics of the National University of Mexico to see if these red sand materials of the lower basin are the same as the constituent materials of the volcanic event that affected the upper basin during the Cuadros phase.

CONCLUSIONS The sedimentation event that buried Cantón Corralito and Santa Ana may indicate that the vegetation in the upper basin of the Coatán River had been affected either by human agency or by the destruction of forests and the deposition of a volcanic ash flow. While more geomorphological investigation is required to test these hypotheses, the buried sites of Canton Corralito and Santa Ana indicate how the intersection between a human group (that elected to build a settlement on the edge of a river basin, deforested the upper basin of the Coatán River, or both) and a cyclical agent that released a large amount of energy (the Tacaná ash flow eruption, an associated large storm, or both) led to one of the most important Cuadros-phase settlements in the Mazatán area being buried under tons of sediments. I believe that the lesson learned by Soconusco people in the subsequent Jocotal phase is demonstrated at the site of Ojo de Agua, which was placed on high ground, 150 m away from the river channel, and consisted of large platforms elevated high above the surrounding terrain. The builders did an excellent job, as Ojo de Agua survived Hurricane Stan and its large constructions deflected the floodwaters. Ojo de Agua in the Jocotal phase represents an example of adaptation, which reflects the intentions of the infor mant who told us, “If the last storm brought me 30 cm of mud, then I will build a house over a platform of one meter, so I can always stay one step ahead of the disaster.” Human societies have always lived with hazards related to the powerful natural forces of our planet. When our economic activities and social organizations underestimate these risks, we become more vulnerable to them and thus the

disaster, cultural change in the coatán

165

creators of our own disasters. Archaeology in combination with geomorphology can contribute to an understanding of the causes behind past and modern disasters, and it is worthwhile to engage in this kind of research to understand better these prior disasters and learn from these experiences to alleviate future pain for others. NOTES 1. This section is based on the proposals of Susanna M. Hoff man and Anthony Oliver-Smith (2002), as well as those of Jorge Olcina Cantos and Francisco J. Ayala- Carcedo (2002). 2. This popu lar belief states that Saint Francis used the cord with which he tied his habit to whip or beat the clouds and cause all the rain to fall out of them, and that this led to heavy rains. October 4 is the date when this “cordonazo” or whipping occurred and marks the end of the rainy season. 3. http://disc.sci.gsfc.nasa.gov/giovanni/tovas _ snapshot.shtml. 4. Infor mants in the field were often unsure of the name of this event and the month. Some mentioned that it was the remnants of a storm named Javier that occurred in September 1998; others mentioned Hurricane Pauline (October 1997); and still others, Hurricane Mitch (October-November 1998). 5. An eyewitness account from the Soconusco relates that it rained ash for more than twenty-four hours starting on October 24. This rain killed livestock, polluted rivers, and covered all vegetation. After a few months, the ash still measured a halfmeter in depth, and it took years before it was covered over, resulting in a horizontal soil layer visible in stratigraphy throughout the Soconusco region (García 1963:395–399). 6. See note 3. 7. A combination of clay, fine sand, and silt is known locally as texcal or “barro to build houses.” Texcal with a higher clay content is considered better for construction purposes. 8. The color of this matrix depends on how much moisture it contains and what kind of light it is seen under, hence the variation in names. As mentioned, the local people define this matrix as arena roja (“red sand”), as oppose to arena blanca (“white sand”), which is a layer reported by Pérez below the deposit of Cherla-phase ceramics, between –2.90 and –4.90 m, in test pit 7 of Cantón Corralito (Pérez and Lesure 1998:23, figure 23). 9. Cheetham refers to it as “orangey brown, medium-size river sand.”

166

emergent complexity

10. With less information, Cheetham et al. (2007:20) estimated a more conservative figure of at least 25 ha. 11. The sand and silt deposited during this event was not necessarily “red” to begin with; the “red” color may have developed from the oxidation of manganese and iron minerals contained in the sands. 12. A 20 m profi le was cut by the torrents of water and debris during Hurricane Stan along the banks of the Coatán River, close to the immigration detention center of Tapachula. In the upper section of this profi le, there is a 1.5 m sedimentation deposit similar to the “red sand” layer observed in the lower basin. Future work will include the collection of a sample of this Tapachula sediment to compare with that of the lower basin.

REFERENCES Anderson, Tempest 1908 The volcanoes of Guatemala. Geographical Journal 31(5): 473–485. Blaikie, Piers, Terry Cannon, Ian Davis, and Ben Wisner 1994 At Risk: Natural Hazards, People’s Vulnerability, and Disasters. Routledge, New York. Blake, Michael, Richard Lesure, Warren D. Hill, Luis Barba, and John E. Clark 2006 The Residence of Power at Paso de la Amada, Mexico. In Power and Palaces in the Americas: From Peru to the Northwest Coast, edited by Jessica J. Christie and Patricia J. Sarro, pp. 191–210. University of Texas Press, Austin. Blomster, Jeff rey, Hector Neff, and Michael D. Glascock 2005 Olmec pottery production and export in ancient Mesoamerica as determined through elemental analysis. Science 307: 1068– 1072. Cheetham, David 2006 The Americas’ first colony? Archaeology 59(1): 42–46. Cheetham, David 2007 Cantón Corralito, Objects From a Possible Gulf Olmec Colony. Available at www.famsi .org/reports/05021/index.html. Cheetham, David, John E. Clark, Gregory Luna, Terry G. Powis, Tomás Pérez Suárez, Artemio Villatoro Alvarado, and Juan Carlos López Espinosa 2007 Proyecto Arqueológico Cantón Corralito, Chiapas, México: Temporada 2004.

Manuscript on fi le, Instituto Nacional de Antropología e Historia, Mexico. Christopherson, Robert W. 2006 Geosystems, 6th ed. Pearson, Prentice Hall, Upper Saddle River, NJ. Clark, John E. 1994 The Development of Early Formative Rank Societies in the Soconusco, Chiapas, Mexico. Unpublished Ph.D. dissertation, Department of Anthropology, University of Michigan, Ann Arbor. Clark, John E. 1997 The arts of government in early Mesoamerica. Annual Review of Anthropology 26: 211–234. Clark, John E. 2004 Mesoamerica Goes Public: Early Ceremonial Centers, Leaders, and Communities. In Mesoamerican Archaeology, edited by Julia Hendon and Rosemary Joyce, pp. 43– 72. Blackwell, Malden, MA. Clark, John E., and Michael Blake 1994 The Power of Prestige: Competitive Generosity and the Emergence of Rank Societies in Lowland Mesoamerica. In Factional Competition and Political Development in the New World, edited by Elizabeth Brumfiel and John Fox, pp.17–30. Cambridge University Press, England. Clark, John E., and Mary E. Pye 2000 The Pacific Coast and the Olmec Question. In Olmec Art and Archaeology in Mesoamerica, edited by John E. Clark and Mary E. Pye, pp. 217–251. Studies in the History of Art No. 58, National Gallery of Art, Washington, DC. Córdova, Carlos, Ana Lilian Martín del Pozo, and Javier López Camacho 1994 Paleolandforms and volcanic impact on the environment of prehistoric Cuicuilco, southern Mexico City. Journal of Archaeological Science 21: 585– 596. Cortina, Sergio 2006 Deforestación en los Altos de Chiapas: magnitud y causas: Recomendaciones para la planeación estratégica forestal. El Colegio de la Frontera Sur, Tapachula, Chiapas. Ekholm, Susanna 1969 Mound 30a and the Early Preclassic Sequence of Izapa, Chiapas, Mexico. Papers of the New World Archaeological Foundation No. 25, Brigham Young University, Provo, UT. Gaitán, Alfaro Héctor 2005 Memorias del Siglo XX, Tomo 2. Librería Artemis Edinter, Guatemala.

García Acosta, Virginia, and Gerardo Suárez Reynoso 1996 Los Sismos en la Historia de México, Tomo I. Universidad Nacional Autónoma de México, Centro de Investigaciones y Estudios Superiores en Antropología Social, Fondo de Cultura Económica, Mexico. García, S., and J. Mario 1963 Soconusco en la Historia. Mexico. Hoff man, Susanna M., and Anthony Oliver-Smith (eds.) 2002 Catastrophe and Culture: The Anthropology of Disaster. School of American Research Seminar, James Currey, Oxford. Instituto Nacional de Estadística y Geografía, Mexico (INEGI) 1996 Carta topográfica, imagen digital, escala 1:250,000. Cartas de Huixtla D1502 and Tapachula D1505, cartas geológicas, de afectos climáticos, topográficas y edafológicas. INEGI, Mexico. Jaime-Riverón, Olaf, and Christopher Pool 2009 The Impact of Volcanic Hazards on the Ancient Olmec and Epi- Olmec Economies in the Los Tuxtlas Region, Veracruz, Mexico. In The Political Economy of Hazards and Disasters, edited by Eric C. Jones and Arthur D. Murphy, pp. 133– 154. AltaMira Press, Lanham, MD. Kerr, Andrew C., Manuel A. Iturralde-Vinent, Andrew D. Saunders, Tanya L. Babbs, and John Tarney 1999 New plate tectonic model of the Caribbean: Implications from a geochemical reconnaissance of Cuban Mesozoic volcanic rocks. Geological Society of America Bulletin 111(11): 1581– 1599. Larsen, Matthew C., Gerald F. Wieczorek, L. Scott Eaton, Benjamin A. Morgan, and Heriberto Torres-Sierra 1999 Natural Hazards on Alluvial Fans: The Venezuelan Debris Flow and Flash Flood Disaster. U.S. Geological Survey Fact Sheet FS-103- 01, U.S. Department of the Interior. Available at http://pubs.usgs.gov/fs/fs-0103 -01/. Lesure, Richard 2000 Animal Imagery, Cultural Unities, and Ideologies of Inequality in Early Formative Mesoamerica. In Olmec Art and Archaeology in Mesoamerica, edited by John E. Clark and Mary E. Pye, pp. 193–215. Studies in the History of Art No. 58, National Gallery of Art, Washington, DC.

disaster, cultural change in the coatán

167

Lowe, Gareth W. 1971 Civilizational consequences of varying degrees of agricultural and ceramic dependency within the basic ecosystems of Mesoamerica. Contributions of the University of California Archaeological Research Facility 11: 212–248. Macías, J. L., J. M. Espíndola, A. García-Palomo, K. M. Scott, S. Hughes, and J. C. Mora 2000 Late Holocene Peléan-style eruption at Tacaná Volcano, Mexico and Guatemala: Past, present, and future hazards. Geological Society of America Bulletin 112(8): 1234– 1249. Manzanilla, Linda 1997 Indicadores arqueológicos de desastres: Mesoamérica, los Andes y otros casos. In Historia y Desastres en America Latina, vol. 1, edited by Virginia García Acosta, pp. 33– 58. Red de Estudios Sociales en Prevención de Desastres en América Latina (LA RED), CIESAS, and Intermediate Technical Development Group, Lima. Meschede, Martin, and Udo Barckhausen 2000 Plate Tectonic Evolution of the Cocos-Nazca Spreading Center. In Proceedings of the Ocean Drilling Program, Scientifi c Results, vol. 170, edited by E. A. Silver, G. Kimura, and T. H. Shipley. College Station, TX. Available at www-odp.tamu.edu/publications/ 170 _SR/ VOLUME/CHAPTERS/SR170 _07 .PDF. Mora, J. C., J. L. Macías, A. García-Paloma, J. L. Arce, J. M. Espíndola, P. Manetti, O. Vaselli, and J. M. Sánchez 2004 Petrology and geochemistry of the Tacaná Volcanic Complex, Mexico- Guatemala: Evidence for the last 40,000 years of activity. Geofísica Internacional 43(3): 331–359. Musset, Alain 1994 L’Amerique Centrale et les Antilles, une Approache Géographique. Masson, Paris. Olcina Cantos, Jorge , and Francisco J. Ayala- Carcedo 2002 Riesgos naturales. Conceptos fundamentales y clasificación. In Riesgos Naturales, edited by Francisco Javier Ayala- Carcedo and Jorge Olcina Cantos, pp. 41– 73. Ariel Ciencia, Barcelona. Oliver-Smith, Anthony 2002 Theorizing Disasters: Nature, Power, and Culture. In Catastrophe and Culture: The Anthropology of Disaster, edited by Anthony Oliver-Smith and Susanna M. Hoffman, pp. 23–47. School of American Research Seminar, James Currey, Oxford.

168

emergent complexity

Oliver-Smith, Anthony, and Susanna M. Hoffman (eds.) 2002 Introduction: Why Anthropologists Should Study Disasters. In Catastrophe and Culture: The Anthropology of Disaster, pp. 3–22. School of American Research Seminar, James Currey, Oxford. Peraldo, Giovanni, and Walter Montero 1996 La secuencia sísmica de Agosto a Octubre de 1717 en Guatemala. Efectos y respuestas sociales. In Historia de los Desastres en America Latina, vol.1, edited by Virginia García Acosta, pp. 295–324. Red de Estudios Sociales en Prevención de Desastres en América Latina (LA RED), CIESAS, and Intermediate Technical Development Group, Lima. Pérez Suárez, Tomás 2002 Cantón Corralito: Un Sitio Olmeca en el Litoral Chiapaneco. In Arqueología Mexicana Historia y Esencia. Siglo XX: En Reconocimiento al Dr. Román Piña Chán, edited by Jesús Navarro, pp. 71– 92. Instituto Nacional de Antropología e Historia, Mexico. Pérez Suárez, Tomás, and Richard G. Lesure 1998 Informe del Proyecto Arqueológico de las Aldeas a los Centros de Poder en la Costa de Chiapas. Manuscript on fi le, Instituto Nacional de Antropología e Historia, Mexico. Pineda, Emeterio 1999 [1845] Descripción geográfica del departamento de Chiapas y Soconusco. Consejo Estatal para la Cultura y las Artes de Chiapas y Fondo de Cultura Económica, Mexico. Pinkowski, Jennifer 2006 A city by the sea. Archaeology 59(1): 46–49. Plunket, Patricia, and Gabriela Uruñuela 1998 Appeasing the volcano gods. Archaeology 51(4): 36–42. Preston, David 1996 Latin America Development, Geographical Perspectives, 2nd ed. Longman, Essex. Programa Emergente Stan 2006 Programa Emergente Stan Reporte por Beneficiario con Tipo de Daño, 16 de Febrero, 2006. Manuscript on fi le, Instituto de Vivienda, Gobierno de Chiapas, Tuxtla Gutiérrez, Mexico. Quiñones, Keber Eloise 1995 Codex Telleriano-Remensis: Ritual, Divination, and History in Pictorial Aztec Manuscript. University of Texas Press, Austin.

Redman, Charles 2005 Resilience theory in archaeology. American Anthropologist 107(1): 70– 77. Richter, Michael 1993 Características fitomorfológicas como indicadores climatológicos en el sur de Chiapas. In Investigaciones Ecogeográficas Sobre la Región del Soconusco, Chiapas, edited by Michael Richter, pp. 13–34. Centro de Investigaciones Ecológicas del Sureste, Mexico. Rose, William I., Julian J. Bommer, Dina L. López, Michael J. Carr, and Jon J. Major 2004 Natural Hazards in El Salvador. Special Papers No. 375, Geological Society of America, Boulder, CO. Sahagún, Bernardino de 1969 Florentine Codex: General History of the Things of New Spain. Book 6: Rhetoric and Moral Philosophy, translated by Charles Dibble and Arthur Anderson. University of Utah and School of American Research, Salt Lake City. Sanders, William T., Jeff rey R. Parsons, and Robert S. Santley 1979 The Basin of Mexico: Ecological Processes in the Evolution of a Civilization. Academic Press, New York. Sheets, Payson 1992 The Ceren Site: A Prehistoric Village Buried by Volcanic Ash in Central America. Holt, Rhinehard, and Winston, Fort Worth, TX. Sheets, Payson 2002 Before the Volcano Erupted: The Ancient Cerén Village in Central America. University of Texas Press, Austin. Sheets, Payson 2004 Apocalypse Then: Social Science Approaches to Volcanism, People, and Cultures in the Zapotitlan Valley, El Salvador. In Natural Hazards in El Salvador, edited by W. I. Rose, J. J. Bommer,

D. L. López, M. J. Carr, and J. J. Major, pp. 109– 120. Special Papers No. 375, Geological Society of America, Boulder, CO. Shillings, S. P., J.W. Vallance, O. Matías, and M. M. Howell 2001 Lahar Hazards at Agua Volcano, Guatemala. U.S. Geological Survey Open-File Report 2001-432, U.S. Department of the Interior. Available at http://vulcan.wr.usgs .gov/ Volcanoes/Guatemala/Publications/ OFR01-432/framework.html. Shook, Edwin M. 1965 Archaeological Survey of the Pacific Coast of Guatemala. In Handbook of Middle American Indians, vol. 2, edited by Gordon Willey, pp. 180– 194. University of Texas Press, Austin. Vázquez, Francisco 1937– Crónica de la Provincia del Santísimo 1944 Nombre de Jesús de Guatemala de la Orden de N. Seráfico Padre San Francisco en el Reino de la Nueva España, 2nd ed., 4 vols. Biblioteca “Goathemala” de la Sociedad de Geografía e Historia de Guatemala, vols. 14 through 17, Guatemala. Voorhies, Barbara 2004 Coastal Collectors in the Holocene: The Chantuto People of Southwest Mexico. University Press of Florida, Gainesville. Voorhies, Barbara, and Sarah E. Metcalfe 2007 Culture and Climate in Mesoamerica During the Middle Holocene. In Climate Change and Cultural Dynamics: A Global Perspective on Mid-Holocene Transitions, edited by David G. Anderson, Kirk A. Maasch, and Daniel H. Sandweiss, pp. 157– 187. Elsevier, Amsterdam. Wallace, Anthony F. C. 1956 Human Behavior in Extreme Situations. National Academy of Sciences—National Research Council, Washington, DC.

disaster, cultural change in the coatán

169

EIGHT

La Blanca and the Soconusco Middle Formative Michael Love and Julia Guernsey

T

he greater soconusco saw the dra-

matic transformation of human societies during the second and first millennia b.c., with the transition from the first sedentary villages to urban settlements taking place in just 1,500 years (ca. 1900–400 b.c.).* Change during that time undoubtedly was constant, but it was not necessarily seamless. While manifestly the long-term trend was toward increasing complexity, there may have been episodes of turmoil leading to cycles in which political integration alternated with times in which centralized political power was weak (Love 2002a). The clearest evidence that we have of political cycling on the coastal plain is the frequent change of capital cities. Although the ultimate causes of cycling may prove to be climatological, because of low overall rainfall and wide variation in annual totals the effects of political cycling were evident in many areas of ancient

* All dates are in calendar years. For regional chronology, see Figure 1.3 in Chapter 1; for sites not shown on Figure 8.2 in this chapter, see Figures 1.1 and 1.2.

170

life. In each cycle we see dramatic evidence of changes in social organization, economics, and ideology. We can identify at least four cycles in the sequence prior to the development of archaic states during the Late Formative period: (1) Paso de la Amada (Locona-Ocos phases); (2) Canton Corralito (Cuadros Phase); (3) Ojo de Agua (Jocotal phase); and (4) La Blanca (Conchas phase). The fifth cycle in the Formative-period sequence involves the development of archaic states at El Ujuxte, Izapa, and Takalik Abaj (Love 2010, 2011, at press). During the Late Formative period, El Ujuxte was the dominant polity of the eastern Soconusco coastal plain, while Izapa and Takalik Abaj developed in the piedmont region. If we call the beginning of the sequence a simple chiefdom and the end points states, the second through fourth cycles would probably all be called complex chiefdoms of varying scale. We use these categories of social forms solely as an ordinal scale and do not wish to imply that they conform to neo-evolutionary idealized societal types. Indeed, the significant differences among the three complex chiefdoms illustrate

the problematic nature of neo- evolutionary typologies. Our focus in this chapter is the fourth cycle and the role of La Blanca in the long-term development of social complexity in the Greater Soconusco. La Blanca’s critical role is highlighted by our interpretation of it as the last chiefdom in the Soconusco Formative-period sequence. Without being teleological by taking for granted the later development of states, or promoting a tautology by representing defi nitions as conclusions, it is justifi able to ask if there were changes that took place during the hegemony of La Blanca that facilitated the subsequent development of states in the region. The changes evident at La Blanca were significant and certainly do not require exaggeration to make them interesting. We do not wish to engage in hyperbole and will avoid calling La Blanca a protostate; yet there are many aspects of La Blanca and its polity that are statelike. There is evidence of social stratification, economic specialization, and an ideology of rulership that many would find consistent with being a state. Statehood consists of more than these social traits, however, because states are not a type of society but instead systems of government. The separation of elites from commoners, the development of craft specialization, and even rulership existed in Mesoamerica by Early Formative times, but the institutionalized and bureaucratic form of government called the state did not. While La Blanca’s rulers and social elite may have wielded significant power and exerted control over massive amounts of labor, evidence for the multitiered system of regional administration characteristic of a state is weak. What did occur at La Blanca, however, was the growth of economic centralization and the promotion of a public ideology that transformed and strengthened the institutions of rulership. In that sense, La Blanca represents a pivotal moment in the longterm growth of political centralization that characterizes the Soconusco Formative period.

PRELUDE La Blanca developed in the eastern portion of the Soconusco, between the Río Naranjo and the Río Tilapa, the latter being the formal boundary of the Soconusco region. In some areas of the site there are aceramic levels with obsidian and groundstone that might indicate an Archaic-period use of the locale; but there are no radiocarbon dates to confi rm that proposition. There is more abundant evidence of occupation during the Locona and Ocos phases found in secondary contexts throughout the eastern portion of the site, but no pure or undisturbed deposits of those two periods have been found. The principal occupation of La Blanca began at about 1000 b.c., or early in the Middle Formative Conchas phase. The site was occupied throughout the Conchas phase but declined noticeably after 600 b.c., when the locus of the regional capital shifted to the site of El Ujuxte, some 12 km east of La Blanca. During the ensuing Caramelo, Cataluña, and Pitahaya phases (600 b.c.–a.d. 100), the site declined to the size of a small village. La Blanca’s rise to prominence followed the decline of the Jocotal-phase settlement of Ojo de Agua, Chiapas (Clark and Hodgson 2004). Although still poorly known, Ojo de Agua was part of the “second wave” of the Olmec interaction network (the term Olmec here is used to denote “a group of people, or peoples, who shared a suite of cultural practices” [Clark and Pye 2000: 217–218]). Ojo de Agua was also the first site in the region to build small, public temple-pyramids around a central plaza (Clark and Hodgson 2004). The extent of the site is uncertain, but probably exceeds 100 hectares (ha). The Jocotal phase as a whole is still not well understood, and the type of polity associated with Ojo de Agua is likewise enigmatic. Nonetheless, we know that Ojo de Agua declined sharply at the beginning of the Middle Formative period and that it was the last capital within the Mazatán district. Following its demise, the

la blanca and the middle formative

171

center of demographic, political, and economic events shifted east of the Río Suchiate, although we cannot yet deduce the reasons for this shift. What occurred was more than a simple west-toeast movement, however, because populations from the east, within the modern Guatemalan department of Retalhuleu, also moved toward the Río Naranjo (Love 2002b). With so little known about Ojo de Agua, it may seem a stretch to call the rise of La Blanca a dramatic break from the past, but it is also hard to characterize it as anything less. The scale of the site, its monumental constructions, and its regional polity vastly exceed local precedents and as such must represent a new order. For nearly 300 years, La Blanca was the largest site on the Pacific coast of Mesoamerica, and the size of its monumental architecture was never matched in coastal Guatemala. La Blanca’s overall size and scale of construction mark it as one of Mesoamerica’s most important sites in the Middle Formative period and an important part of the Olmec interaction network of 1000 b.c. to 600 b.c. Although we lack many specific data for the period immediately before La Blanca, its rise is marked by greater centralization of political power, greater social distance between elites and nonelites, economic intensification, and new ideological structures. Economically and politically, the Conchas phase marked a time when elites gained more power, as indicated by their ability to mobilize labor on a scale that was much greater than that of previous polities. The basis of the increased social power wielded by La Blanca’s elite was twofold, resting on economic and ideological pillars. Economically, an intensification of the economy was carried out by a great focus on high-yield subsistence resources, especially maize and dogs (Blake et al. 1992; Love 1999a, 2002b; Wake and Harrington 2002). Ideologically, elites at La Blanca, along with those in other regions linked to the Olmec network of interaction, created a system of rulership based on perceived abilities to manipulate supernatural forces and communicate with other levels of the Mesoamerican cosmos.

172

emergent complexity

LA BLANCA AS A CITY A new map of the site (Figure 8.1), under development since 2003, has allowed us to disprove earlier characterizations of La Blanca as having little structure or planning (cf. Love 1999b). Greater access granted by the current landowners in the area has also allowed us to expand estimates of the site’s overall size significantly. The core of La Blanca is a ceremonial precinct of about 100 ha. Low habitation mounds extend over at least 200 ha, but surface materials are found over nearly 300 ha, stretching from 1 km north of Mound 1 to over 1.4 km south. Although the southern portion of the site was heavily damaged by road construction, including the excavation of enormous borrow pits from which fill material was extracted, excavations by Edwin Shook in 1972 recovered Middle Formative domestic refuse from that southern portion. A number of factors complicate the task of determining more precise boundaries. First, there is the extensive damage to the site caused by road construction in 1972 and 1973. Second, road grading certainly destroyed habitation mounds in the southern sector, but grading also may have dispersed surface materials. Third, the ability to survey the identified fringes of the site has been curtailed by some landowners who are as yet unwilling to grant access to their property. Despite the difficulties in specifying the precise size of the settlement, the known extent and density of occupation at La Blanca matches or exceeds the size of settlements identified as urban elsewhere in the world (Love 2011), and we therefore consider it to be a city.

Site Planning and Ceremonial Construction The central ceremonial zone of La Blanca was raised and leveled with an estimated 2 m of fi ll throughout (Figure 8.1 insert). Based on excavations in Mound 25, which revealed a Conchas A building at the southern end of the leveled area buried under 2 m of fi ll, this construction episode took place at the end of the Conchas A subphase, very soon after 900 b.c.

Mound 4

Mound 1

Mound 5

Sunken Plaza

West Acropolis

Joyas Group

Great Plaza

South Plaza Mound 3

Mound 225

4.

5

5

Detail of Central Zone

4

4.5

3

La

ta ne gu 3

Mound 2

5 6

7

4

8

2

3

Mound 5

Mound ound 4

Mound 6

L a gu

5 .5

8

net

Mound 1

eta1

5

6

gun La

Ca rn C Modern

R ad rn Ro Modern

nal

a 2

Mound 3

3

5

3 5 4

3

4 3

7

6

7

6

LA BLANCA

Nm

5

OCOS, SAN MARCOS

ag

5

5

Contour Interval 50 cm Contour Less than 50 cm 0

100

East Acropolis

Mound 6

West Plaza

Mound 2

200

METERS

FIGURE 8.1 Map of the northern portion of La Blanca, with detail of the central zone (inset)

Within this large area of leveled terrain, two more elevated areas were constructed—the East Acropolis and the West Acropolis. The East Acropolis consists largely of buildings identified as elite residences, although Mound 5 (mapped in 1985 but now destroyed), with a height of over 5 m, may be a public or ceremonial construction. Atop the West Acropolis a number of large public mounds once stood. Several were destroyed in 1972 but others remain, including Mound 2, the largest standing structure at the site, and a possible E-group complex that includes Mound 6 and the three low mounds east of it. The public mounds at La Blanca were constructed on an axis aligned with the peak of the Tajumulco volcano, the tallest peak in Central America with an elevation of 4221 m. The central axis of the site is a line that passes from Mound 3 through Mound 4, which is a long structure oriented 22 degrees east of magnetic north. The largest ceremonial construction at the site, Mound 1, was one of the earliest monumental temple-pyramids in Mesoamerica. The mound originally stood over 25 m in height and measured 150 × 100 m at its base. It was built without stone, using a sophisticated rammed earth technique and a capping layer of selected dark brown clay. The greatest bulk of the mound was raised in a single construction episode, with subsequent enlargements accomplished by the application of relatively thin strata of earth. Mound 1 was largely destroyed in 1972 for road construction, but investigations in 2003, 2004, and 2008 found that a total of over 4.5 m of its base survives. Ceramics within the earliest phase of construction indicate that Mound 1 was built during the Conchas B subphase, sometime after 900 b.c. (Love et al. 2005), and was probably undertaken at the same time the West Acropolis was constructed. We can now recognize four distinct plazas within the central precinct, one on each side of Mound 1. The south plaza is bounded by Mound 1 to the north, Mound 3 to the west, Mound 25 to the east, and the unnumbered ridge to the south. We have long assumed the southern

174

emergent complexity

ridge to be a natural feature but now suspect that it may be an artificial construction. The West Plaza is formed by Mound 1, Mound 3, and the Stela Mound (now destroyed, but recorded by Shook). The Sunken Plaza lies east of Mound 1 and is bounded to the east and north by elite residential mounds. The Great Plaza is bounded by Mound 1, the East Acropolis, and the West Acropolis. It may be open to the north, but one might also view it as being bound to the north by the Joyas Group. The Joyas Group was only recently mapped because we had previously been denied access to the property on which it lies. The confirmation that it is a Middle Formative complex extends the boundaries of the site 1 km farther north than previously suspected. There are continuous, but light, materials found between the site core and the Joyas Group, at about the same density as found on any nonmound area of the site core. The Joyas Group itself contains a public mound (Mound JG-1, 5 m tall) and more than a dozen small mounds that apparently are domestic. Mound JG-1 was built close to, but not precisely on, the line of sight from Mound 1 to Tajumulco. There is evidence of another mound, now destroyed, that originally stood directly in that line of sight. Another small ceremonial mound, named the CE (Chico Esquivel) Mound, may mark the southern limit of the site. That mound, which is as yet unmapped, lies 1.4 km south of Mound 1. Coring by Hector Neff (2005) found Middle Formative habitation in the area, underlying Late Classic materials, and the salvage excavations by Shook in 1972 documented Middle Formative occupation nearby. Although the mound, which is approximately 2 m in height and 15 × 40 m in extent, has not been excavated, a Middle Formative date is suggested by its placement: a straight line connects the CE Mound, the center of Mound 1, and Mound JG-1. The formality and scale of La Blanca’s site core is impressive, even in the decimated state in which it now stands. The massive Mound 1 was surely an awe-inspiring sight for residents

in and around La Blanca, visible for a distance of over a dozen kilometers. The ability of La Blanca’s elite to control substantial resources— including both material wealth and labor— and to deploy them in sophisticated engineering projects speaks not only of substantial power but of a well- organized managerial apparatus. The design and building of monumental architecture using sophisticated engineering is evidence of a group of skilled engineers and architects supported by either the personal wealth of rulers or by taxes levied on the populace.

Monuments and Sculpture The patronage of specialists at La Blanca is also evident in the corpus of sculptural art (Love 2010; Love and Guernsey 2007). At present only two fragments of stone sculpture are known from La Blanca. Monument 1 is a head carved in the Olmec style of sculpture, depicting an anthropomorphic figure with supernatural attributes. Monument 2 is a knee fragment from a standing human figure. Other monuments, notably a stela, are rumored to have been present in the West Plaza of the site but have not been located. Shook (personal communication 1985) related that he was told of a stela that was destroyed by local residents who thought the mica in the matrix of the stone was gold; Shook never saw the rumored stela himself. Other important but smaller artworks from the site may include the Young Lord sculpture (also known as “Slim”) and a saurian image now in the Museo del Libro in Antigua, Guatemala (Fields and Reents-Budet 2005: cat. no. 29). Monument 3 is a unique, earthen, quatrefoilshaped altar (Love and Guernsey 2007). The quatrefoil symbol is long-lived in Mesoamerica and is linked to elite power and the ability of rulers to communicate with the supernatural, especially the underworld (Guernsey 2010). Monument 3 comes from the western edge of Mound 9, an elite residence, and fronts the Sunken Plaza, east of Mound 1. Its placement was thus liminal between a presumably public space and

a supposedly private residential location. This monument and its associations will be discussed more extensively later in this chapter. La Blanca Monument 4 is a fragment of what was originally a very large ceramic sculpture of similar size and shape to La Blanca Monument 1 (Love 2010). The recovered portion of Monument 4 is the nose and eye of a supernatural creature, or perhaps a human with supernatural attributes. The eye is represented by a hole encircled by red paint, and the mouth is rendered as a curl motif incised within an inverted cleft highlighted by a resist technique. If the comparison to Monument 1 is correct, the head of Monument 4 would have originally been about 40 cm high. It is possible that Monument 4 is a piece of a mask rather than a freestanding sculpture, but the thickness of the ceramic (over 4 cm) means that the complete piece likely would have been far too heavy to wear.

Central La Blanca as the Locus of Public Ritual Excavations in the public areas of La Blanca to date have been limited to investigations of Mound 1. During the 2008 season we found evidence of structures associated with Mound 1 on both the north and south sides, at the base. On the south side we found floors and post molds, indicating that a large roofed structure was placed at the bottom of the ramp used to access the sanctuary that presumably sat at the pinnacle of the pyramid. On the north side we found a floor and post molds that were buried within the first stage of construction of the pyramid. That is, the construction of the pyramid covered this earlier structure, which was associated with a feature that included dense quantities of shells, the only shell recovered at the site to date (see Bove 2005:98 and Shook 1971 regarding the absence of shell in most archaeological sites in this region). The structure underlying Mound 1 may have been an earlier temple platform, or it may have been a platform built to dedicate the site before

la blanca and the middle formative

175

Río Naranjo Survey Conchas Phase 900 - 600 B.C. Regional Center

Secondary Center Village Single Residence Mangrove

El Infierno

Pampa

5

0

iate

Kilometers

La Zarca

RÌo Naranjo

RÌo Such

La Blanca

Pa

cif

ic Oc ea n

FIGURE 8.2 Middle Formative sites of the Río Naranjo district

Mound 1 was raised. In either case, it marks the only evidence that we have to date of public rituals possibly associated with Mound 1. The materials from the features associated with the platform are still being catalogued, so it is too early to offer any firm interpretations of the nature of those rituals. However, dense quantities of bone suggest that food preparation and feasting accompanied the rituals, and quantities of small pebbles, similar to the caches found at Late Formative El Ujuxte (Love 2002b), suggest that rites of divination or calendrical reckoning may have been performed.

176

emergent complexity

THE REGIONAL SYSTEM OF LA BLANCA The regional system of La Blanca has at least three and perhaps as many as five levels (Figure 8.2). The first level, of course, is La Blanca itself. The second level consists of two centers, El Infierno and La Zarca, each of which has one large public mound over 15 m in height. Izapa may also have been a secondary center of La Blanca during the Conchas phase (Love 2002b). The third level is composed of villages with between two and ten residences, and the fourth level is formed by isolated residential

mounds. Rosenswig (2005) proposes another level within the regional hierarchy, below the secondary centers and above the villages. He views Cuauhtemoc as such a site, having a single public mound approximately 5 m in height, and has identified other sites east of the Río Suchiate with similar mounds (see Rosenswig, Chapter 11 in this book). The place of these sites within the La Blanca polity, if indeed they were part of the polity, is nebulous for the moment. The most conservative classification would see three levels of hierarchy in La Blanca’s regional system, but there may be as many as five. A settlement hierarchy of this nature would be viewed by most archaeologists as indicative of something less than a state and might best be classified as a complex chiefdom. The secondary centers, while having monumental architecture, are not large settlements, and it is likely that most governmental functions were handled directly from La Blanca. The apparent lack of a well-developed regional system of administration is a key factor in judging La Blanca to be a chiefdom rather than an archaic state. Admittedly, at present we do not have much data by which to judge the political and economic relationships between La Blanca and its hinterland. Most of the data on social relationships of the Middle Formative La Blanca polity come from domestic excavations at La Blanca, supplemented by Rosenswig’s data from Cuahuatemoc. The earlier excavations at La Victoria by Coe (1961), although vital for initially defining the Conchas phase, yielded little information on economic activities or social differentiation.

HOUSEHOLD DIFFERENTIATION AT LA BLANCA The vast majority of work undertaken at La Blanca has focused on the analysis of household economy and ritual. The central theoretical objective has been to examine the relationship between the household and the centralized political and economic authority of a complex chiefdom. This objective entails discerning the nature of changes that took place in the

household as a response to the growing social complexity and increasing political centralization that characterized the Formative period. Eventually, these differences will be examined in light of other evidence for the nature of the regional polity, local relationships with other neighboring zones, and changing social identities of elite and nonelite sectors within the society. The relationship between households and emerging centralized political systems has been the subject of many studies in diverse parts of the world (Hastorf and D’Altroy 2001; Sahlins 1972; Wattenmaker 1998). Most studies have focused on the household as an economic unit and examined how household production is affected by the tribute demands typically made by emerging elites. Characteristically, one finds evidence for economic intensification in the form of more productive types of agriculture, craft specialization, or reorganization of labor that leads to surplus production (although Spielman 2002 provides an alternate logic for intensification). The La Blanca project is following a similar approach and views economic intensification at the household level to be in part a response to elite demands for tribute in kind and labor. To more precisely model how intensification occurred, we have collected detailed data on household economic activities, including faunal and plant utilization as well as the consumption of key commodities. While economically oriented perspectives have provided key insights to understanding early complex societies, they are incomplete in at least two aspects. First, there is a clear bias in focusing on the household solely as an economic unit, without viewing it as a place of ritual and a locus for the construction of identity. Both these factors, and others, need to be integrated into household archaeology side by side with economic analysis. Certainly there are archaeological studies of household ritual and the role of the household in the construction of social identities (see, for example, Hendon 1999 and Joyce 2000), but few have thoroughly explored ideology as an area for conflict between

la blanca and the middle formative

177

centralized power and individual households. A second omission is that of agency. Many studies view households as compliant entities with no options other than to accede to elite demands. Household strategies often are considered only with respect to which strategies will be used to satisfy the demands for increased production, and evidence for resistance to those demands often has been overlooked (but, for exceptions to this, see Joyce, Bustamante, and Levine 2001; McAnany 1995, 2002; and Scott 1990). The La Blanca study views household members as active agents who seek to achieve their own goals in the face of many demands, including those of the developing centralized government. Household economic strategies and ritual activities are analyzed conjointly to see how they together reflect the negotiation of relations with the institutions of centralized power. To date we have sampled domestic assemblages from ten residential contexts at La Blanca. The analysis of the data collected is still ongoing; hence the discussion presented here is decidedly preliminary. That said, this initial analysis looks at aggregate household data for all portions of the Conchas phase. It also includes data from Operation 31, which has both Conchas and post- Conchas (Caramelophase) occupations.

Household Samples and Excavation Strategy The strategy pursued at La Blanca has been to sample house lots rather than to excavate houses. The data collection strategy defined each low mound as a house mound and its environs as a house lot. We drew a stratified random sample from a 400 m2 area, centered on the highest point on each house mound. House lots were divided in five units: the center of the mound and four quadrants, and a 2 × 2 m unit was placed randomly in each of the five areas. Those five 2 × 2 m units constitute the primary sample within each house lot, but additional units were excavated if time permitted, or if they were needed to expose important

178

emergent complexity

features. The excavated matrix from the primary sampling units was wet-screened through 3 mm mesh, with 10 liters from each sample reserved for flotation and screening through 1mm mesh.

Household Differentiation and Socioeconomic Scaling We cannot yet trace the trajectory of social differentiation in the Soconusco through the course of the entire Formative period. However, we can be certain that by the Late Formative, society was highly stratified, with rulers representing an elite that dominated the bulk of society; the palace of a ruler has been tentatively identified at El Ujuxte (Love, at press). Between the Early Formative and Late Formative, however, there are large gaps in the data, so that we cannot be sure when the local institutions of rulership emerged and when social stratification became rigid. There are household data from Paso de la Amada for the Early Formative period (Lesure 1999; Lesure and Blake 2002), but such data are lacking for the Cuadros and Jocotal occupations of Canton Corralito, Ojo de Agua, and Cuauhtemoc (we distinguish here between data from household units and those from more generalized “domestic deposits”). Thus, the comparative framework for the La Blanca data is limited. The Early Formative data from Paso de la Amada suggest limited differentiation among households (Lesure and Blake 2002). Artifactual and faunal assemblages show few consistent differences among households, with architecture being the strongest discriminator. Lesure (1999) did, however, note differences in the amounts of jade found in seven contexts at the site. Excavations in residential zones at La Blanca show significant social differentiation at the household level (Love 1991, 2006). Elite households are marked by high densities of prestige goods, including jade, mica jewelry, and fine paste ceramics decorated with elaborate iconography. The elite residences also have higher densities of obsidian and greater num-

TABLE 8.1 Comparison of Household Scores for Six Artifact Classes THE HIGHEST-RANKING SCORE IN EACH ARTIFACT CLASS IS IN BOLD

Jade (g/M3)

Fauna (g/M3)

Obsidian (g/M3)

Blade (PCT by count)

Figurines (N/M3)

Duck/jaguar effigy bowls (counts)

25

0.03

3.39

9.54

8.7

0.30

0/0

26

0.18

31.21

18.87

25

1.40

0/0

27

0.13

23.50

27.39

27

2.57

0/0

31

0.27

23.90

18.4

17

0.58

0/0

32

0.32

44.31

36.5

16

2.68

1/1

34

0.1

NA

11.49

13.3

NA

3/2

35

0.03

NA

17.23

27.3

NA

6/1

36

0.09

NA

NA

NA

NA

1/1

37

NA

NA

NA

NA

NA

3/0

38

NA

NA

NA

NA

NA

3/0

Operation

bers of cores, suggesting that they controlled long- distance exchange. Based on a preliminary analysis of the household data from La Blanca, we can discuss a few elements that are indicative of household variability: differences in prestige goods, differences in household histories, differences in construction, and differences in faunal materials. Data are not yet available for all artifact classes across all households, so the number of households in each comparison varies. These data are summarized in Table 8.1. Differences in Prestige Goods The best discriminator among households at La Blanca is the quantity of jade. Jade has been found in all households excavated to date, largely in the form of small disk beads. These beads are usually less than 5 mm in diameter, so they are found only when 1/8-inch (3 mm) screens are used. A petrographic analysis of the jade is pending, but we have attempted to distinguish between jade and other forms of greenstone by a simple scratch test. A quantitative analysis of all greenstone artifacts is under way as of this writing.

The distribution of jade is highly skewed, making this variable an excellent means of scaling socioeconomic status. Jade may have been the means by which perishable agricultural surpluses were transformed into durable wealth. Small jade beads, and perhaps other greenstone beads as well, would have been an ideal means of preserving wealth and conceivably might have functioned as currency. The importance of these forms of wealth has been underestimated, we believe, simply because so few archaeologists in Mesoamerica have used fine screens and hence have not recovered these small objects. The differential distribution of these small jade beads is extremely important because it provides firm evidence that Middle Formative elites were able to convert surpluses into private wealth. Such direct evidence for the economic basis of elite power is lacking for earlier periods, although it may eventually be found. Some of the surpluses generated by economic intensification during the Middle Formative were used for public projects, such as construction and monument erection, but large portions appear to have been used by the elites for their own purposes.

la blanca and the middle formative

179

Obsidian Obsidian density also shows significant differences across households. Jade density is imperfectly correlated with obsidian density, but the top-ranked households on each measure are the same. We also examined the relative frequency of prismatic blades in each household, simply dividing the total weight of obsidian by the weight of prismatic blade fragments. We had expected high-status residences to show greater relative frequency of blades, reflecting control of the blade production process through patronage of craft specialists or simply greater consumption of what is thought to have been a valuable commodity. Such is not the case, and there does not appear to be a correlation between household wealth and blade consumption found in these data. The chief discrepancy between the ranking on blade consumption and the other measures is that Operation 32 ranks in the middle of the pack for blade consumption, despite ranking first in jade density and in overall obsidian density. One possible explanation is that some nonblade material in Operation 32 is the product of blade manufacture. There is evidence of blade production in Operation 32, with subsequent reduction of the prismatic cores via bipolar flaking. There is also evidence of core correction within that operation, so that we know blade production took place in that household (Matthew DesLauriers, personal communication 2008). Thus a simple ratio of blades to nonblades may not reflect the social and economic dimensions of obsidian production and use, and further analysis is necessary to clarify the situation. Household Construction We have yet to excavate a complete house floor at La Blanca, so there are no data on household size or architectural elaboration. The only data we have so far on differences in house construction deal with the quality of materials used and the amount of labor invested in the

180

emergent complexity

enterprise. These data are largely subjective, but we hope to substantiate them in the future through particle-size analysis of floor samples. The households that are considered to be elite are found in Operations 27, 32, and 37. In each of these households we found wellpreserved remains of the house. In all operations we found well- constructed floors, and in Operation 32 we also found remains of wall sections. In these three cases the house floors were made of well- compacted materials with high clay content. In contrast, the floors in less highly ranked residences consisted of nothing more than packed earth. An exception is Operation 25, where the house floor was well constructed and built of a very sandy loam. Operation 25, however, is a Conchas A dwelling (the only one identified so far), so the temporal dimension may be a cause of the differences. Household Histories Elite households tend to have been occupied for longer periods and to have undergone multiple renovations. Among the three highestranked households, Operation 27 had three episodes of construction, Operation 32 also had three and Operation 37 had four. Three house lots of “commoners” excavated farther from the site center appear to have had single occupations. We illustrate this point by a comparison of Operation 27 with Operation 34. Operation 27 has a long occupational history, covering three subphases, Conchas B to D. Artifacts were in high densities to a depth of nearly 3 m. The house was renovated once during each of the subphases, as reflected in three distinct house floors and three peaks in artifact density. In contrast, Operation 34 (classified as a commoner residence on the basis of jade) was occupied only during Conchas C. A small fragment of a single house floor was found at a depth of approximately 1 m below the modern surface. Artifact densities peaked at this depth also, and there is a unimodal distribution of artifact density.

A complete analysis of the La Blanca faunal assemblage is now under way by Thomas A. Wake at UCLA. Pending the results of his analysis, the only data available are for the amount of faunal remains recovered in five of the households excavated (Table 8.1). These data show a significantly higher density of bone in Operation 32, nearly 50 percent more than that of the second-ranking household. Recall that Operation 32 is the household ranked highest on jade density and overall obsidian density. The meaning of these data will not become clear until the complete suite of faunal remains is analyzed for all households, but a working hypothesis clearly must be that Operation 32 was the locus of feasting. Such feasts may have been held in conjunction with rituals around Monument 3 or perhaps in the adjacent Sunken Plaza. Rosenswig (2005, 2007) proposes that there was a significant shift in feasting patterns from the Early to the Middle Formative, and he believes that elites gained greater control over feasting activities as a means of solidifying their political power. However, the distribution of fancy serving vessels at La Blanca appears to be widespread, indicating that most, if not all, households held some kind of feast. That is, there do not appear to be differences among households in the distribution of putative feasting vessels at La Blanca, but differences in the scale of feasting may be reflected in the density of faunal remains and perhaps in the kinds of food served.

lespie 2002; Marcus 1998, 1999). Cyphers (1993) and Joyce (2003) have proposed that fi gurines were used to mark important milestones in the life history of individuals, such as transitions in age grades, and Joyce further suggests that they could have been utilized in  public contexts adjacent to residential zones. Whatever their precise meaning, handmade ceramic figurines are omnipresent in residential contexts at La Blanca in very high densities, which indicates the great importance of household ritual in Middle Formative times. There are differences in figurine densities among households (Table 8.1), but we cannot say definitively whether there are differences in the types of representations. At present, our only data on figurines compare the relative frequency of human and animal representations among them; those data show little differentiation among households (Love and Guernsey 2006). During the 2006 and 2007 seasons we found fragments of large hollow figurines associated with an elite residence (Operation 37). The fragments that we can mend appear to be from large seated figures lacking primary sexual characteristics. We have not yet tabulated the frequency of these figurines, but their distribution is very limited, both temporally and spatially. We cannot say definitively that hollow figurines do not appear in other residences, but we certainly have more fragments of such fi gurines in that single residence than in all others.

Household Ritual

Altars

At least four different types of materials can plausibly be linked to household ritual: figurines, altars, incensarios, and pottery vessels of various forms. Most scholars seem to accept that household rituals involved the use of ceramic figurines. The nature of those rituals is unclear, but many interpret these figurines as artifacts used in the domestic context, especially in ancestor veneration (Grove and Gil-

Within the ceramic assemblage of the Conchas phase are small numbers of heavy cylindrical vessels with flat, solid tops. These were called “stools” by Michael Coe (1961), a designation that has been generally accepted. Our interpretation, however, is that they were used as portable altars within the household. We base our interpretation on the fact that many of the vessels are too thin to support the weight of a human body and

Faunal Utilization

la blanca and the middle formative

181

FIGURE 8.3 Ceramic altar from Operation 34. Vessel height is 50 cm. Drawing by Julia Guernsey.

that a very high percentage show evidence of burning on their upper surfaces. We do not yet have complete data on the distribution of the form, but it is fairly widely distributed and without doubt found in nonelite households within our sample. The altar illustrated in Figure 8.3 came from Operation 34 in a residence that ranks low on most indices of social status. Despite the household’s rank, the altar is decorated with elaborate iconography that includes a cleft maize kernel, which shows up in four places. The other prominent design element is a thick vertical line interrupted at regular intervals with what Joralemon (1971) identified as “cleft ended vegetation.” The context of the altar suggests that people of even lower socioeconomic status at La Blanca had access to objects laden with iconographic significance (Guernsey and Love 2008b). Duck, Duck, Jaguar There are many forms of iconographic representations found on the pottery of La Blanca, including what may be motifs that anticipate calendrical glyphs such as the Ajaw, Lamat, and K’in signs of the Maya calendar (Figure 8.4). There also are many representations of supernatural creatures, including cleft-headed beings, were-jaguars, bats, and unidentified composite creatures.

182

emergent complexity

The putative calendrical signs are so rare as to be unhelpful in analyzing household differentiation, although they are found only in elite households to date. Many other representations are frequent but are still in the process of analysis. At this moment we have data only on two classes of recurring images found on effigy bowls: ducks (or some other long-beaked or snouted creatures) and jaguars (or perhaps some other type of blunt-snouted creature) (Figure 8.5). Twenty-two effigy bowls with modeled representations of these two entities have been found in household excavations. Although there are differences among the households in terms of the number and kind of effigy vessels found, the differences are not understandable in terms of an elite versus nonelite dichotomy. Operation 32, by most measures the highestranking household, has one of each. Other highranking households (Operations 26, 27, and 37) show no patterning. Operation 35, which ranks low on most indices of household status, has the highest total number of effigies and onethird of all duck representations. Given the overall low numbers of these representations, not too much should be made of these data, but the wide distribution of the images is significant insofar that it indicates that their use occurred in households of all socioeconomic levels.

Monument 3 and Household Ritual

FIGURE 8.4 Possible prototypical calendrical glyphs on La Blanca pottery

The evidence presented above indicates little differentiation among households in figurines, altars, and incensarios, although the hollow figurines found in Operation 37 may be unique to that household. The presence of Monument 3 in Operation 32, however, marks that location not only as elite but probably as the residence of the site’s ruler. The quatrefoil as an enduring symbol in Mesoamerica is consistently associated with rulership and with elite communication with the supernatural (Fash 2005; Gillespie 1993; Grove 2000; Guernsey 2010; Love and Guernsey 2007). Monument 3 contains channels that drain into the basin at the center of the quatrefoil and was thus probably used to contain liquids, perhaps akin to the basins used for rites of divination in Formative-period Oaxaca (Marcus 1999). The rites associated with a symbol of rulership, however, would not have been those carried out in other households. The higher density of faunal remains found in Operation 32 may also be indicative of greater, and perhaps different, feasting events in that household.

Summary

FIGURE 8.5 Effi gy bowl images: Jaguar (A) and duck (B)

The La Blanca household data suggest that significant differences in wealth were present by the Middle Formative period. Thus, in addition to having authority over the labor of others, as indicated by the enormous public constructions undertaken in the early Middle Formative period, the La Blanca elite also had significant economic power and the ability to convert perishable agricultural surpluses into durable wealth in the form of jade beads and other prestige goods. Although the La Blanca elite benefited from an ideology that privileged their ability to communicate with aspects of the supernatural, most La Blanca households performed similar rituals in the domestic context. Whatever religious strictures governed the veneration of ancestors and auguries of the future, they

la blanca and the middle formative

183

appear to have been shared among all sectors of the social spectrum. We know that public rituals took place at La Blanca, be we do not yet know the precise nature of those ceremonies, nor who was present at them. The tremendous size of the Great Plaza certainly could have accommodated large numbers of people, yet the presence of a roofed sanctuary at the southern base of Mound 1 suggests that smaller numbers may have been privileged to witness certain rituals or portions of complex rites.

LA BLANCA’S ROLE IN THE EVOLUTION OF SOCIAL COMPLEXITY IN THE GREATER SOCONUSCO While most of the available data from La Blanca are compatible with labeling it a chiefdom, there are some traits that anticipate the emergence of archaic states in the Late Formative period. First, there is evidence of a strong centralized government. The concentration of population in the Río Naranjo zone, drawing population from both the Mazatán zone to the west and the Río Jesus zone from the east, may be evidence of coercive resettlement (Love 2002a). The three or four levels of La Blanca’s regional system, while less than statelike, indicate a growth of administrative structures relative to the two-level systems of the Early Formative. The massive mobilization of labor to level the central zone of the site, construct the East and West acropoleis, and raise Mound 1 is also evidence of a strong centralized government with a well- developed administrative apparatus and specialized engineers. Secondly, there is evidence of social stratification at La Blanca, based not just on prestige and authority, but on economic power. The differences in wealth between the Eastern Acropolis residences and others at the site show a great gap between the haves and the have nots. We have not yet excavated complete houses at La Blanca, so we cannot label any of the elite residences as palaces, although there are differences in the quality of construction.

184

emergent complexity

Third, the iconography of La Blanca shows the presence of an ideology of rulership. The quatrefoil, Monument 3, is clearly linked to a pan-Mesoamerican Middle Formative system of elite authority. Some representations of other supernatural figures, such as the were-jaguar, and glyphlike motifs, such as the Ajaw and Lamat signs, are limited to elite residences at La Blanca. Images from the figurine assemblage may depict rites of transformation in which rulers took on the qualities of supernatural beings, a theme that appears at other sites in their corpus of monumental art (Guernsey 2006). Rulership in and of itself is not indicative of a system of government, however, and other evidence needs to be weighed when deciding how to classify a given case. The size of the system and the complexity of the governmental apparatus are key to defining the existence of both chiefdoms and states. On both those scores the La Blanca polity cannot be called a state. The overall size and suggested population are too small, and the second-level centers do not appear to be large settlements with significant administrative complexes. They have no public buildings other than their principal mounds; therefore they may have acted merely as ritual centers rather than as governmental ones. As in the Early Formative, much Middle Formative ritual was still centered strongly in the household, and it can be viewed as an effective means of commoners’ resistance to elite ideologies of power. Nonetheless, in the construction of large public spaces and monumental architecture we can see the beginnings of a movement of ritual away from the household and into the public domain in an attempt to undercut resistance to the expansion of centralized power. Building on the base begun by La Blanca’s elite, Late Formative rulers were successful in appropriating rituals that had been carried out in domestic settings during Middle Formative times (Guernsey and Love 2008a; Love 1999b, 2002a). New rituals were carried out in public spaces where rulers could control the agenda and present “official transcripts” that legitimated their power.

There is still much to learn about the economy of La Blanca, but evidence of craft specialization is very limited, perhaps including only fine paste ceramics (Love 2002b; Tejeda and Neff 2004). Elites may have controlled some aspects of obsidian distribution, such as fine prismatic blades, but households also procured obsidian for casual production via other networks (Jackson and Love 1991; Love and Jackson 1998). In comparison to the economy of El Ujuxte, that of La Blanca appears to be much less centralized (Love, at press). Nonetheless, we can see the beginnings of centralization that in the Late Formative allowed the elite to further consolidate their power and extract larger surpluses to finance the mechanisms of the state. Despite the evidence of a pronounced social hierarchy, control over labor, and an ideology of rulership, La Blanca is best described as a complex chiefdom. Although the elite may have been powerful, and perhaps wealthy, the institutions of government were poorly developed. Mechanisms of financing governmental operations do not appear to have existed beyond the ability to draw upon the labor of the populace. The elite of La Blanca may have been able to extract a profit from the long- distance trade routes that ran along the Pacific coast, but that possibility is purely inferential. The La Blanca polity did, however, successfully subjugate a large hinterland from which it could draw labor and resources. In that respect, it established the basis for centralized power that was to become more pronounced and more developed in the years that followed its decline. REFERENCES Blake, Michael, Brian Chisholm, John E. Clark, Barbara Voorhies, George Michaels, and Michael W. Love 1992 Prehistoric subsistence in the Soconusco region. Current Anthropology 33: 83– 94. Bove, Frederick J. 2005 The Dichotomy of Formative Complex Societies in Pacific Guatemala: Local Developments vs. External Relationships. In New Perspectives on Formative Mesoamerican Cultures, edited by Terry G. Powis, pp.

95– 110. BAR International Series 1377, Hadrian Books, Oxford. Clark, John. E., and John G. Hodgson 2004 A Millennium of Wonders in Coastal Mesoamerica. Paper presented at the Annual Meeting of the American Anthropological Association, Atlanta, GA. Clark, John E., and Mary E. Pye 2000 The Pacific Coast and the Olmec Question. In Olmec Art and Archaeology in Mesoamerica, edited by John E. Clark and Mary E. Pye, pp. 217–251. National Gallery of Art, Washington, D.C. Coe, Michael D. 1961 La Victoria, an Early Site on the Pacific Coast of Guatemala. Papers of the Peabody Museum of Archaeology and Ethnology, vol. 53. Peabody Museum, Cambridge. Cyphers, Ann 1993 Women, rituals, and social dynamics at ancient Chalcatzingo. Latin American Antiquity 4: 209–204. Fash, Barbara 2005 Iconographic Evidence for Water Management and Social Organization at Copán. In Copán: The History of an Ancient Maya Kingdom, edited by E. Wyllys Andrews and William L. Fash, pp. 103– 138. School of American Research, Santa Fe, NM. Fields, Virginia M., and Dori Reents-Budet (eds.) 2005 Lords of Creation: The Origins of Sacred Maya Kingship. Los Angeles County Museum of Art, Los Angeles. Gillespie, Susan D. 1993 Power, Pathways, and Appropriations in Mesoamerican Art. In Imagery and Creativity: Ethnoaesthetics and Art Worlds in the Americas, edited by Dorothea S. Whitten and Norman E. Whitten, Jr., pp. 67– 107. University of Arizona Press, Tucson. Grove, David C. 2000 Faces of the Earth at Chalcatzingo, Mexico: Serpents, Caves, and Mountains in Middle Formative Period Iconography. In Olmec Art and Archaeology in Mesoamerica, eds. John E. Clark and Mary E. Pye, pp. 277–295. National Gallery of Art, Washington, DC. Grove, David C., and Susan D. Gillespie 2002 Middle Formative domestic ritual at Chalcatzingo, Morelos. In Domestic Ritual in Ancient Mesoamerica, edited by Patricia Plunkett, pp. 11– 19. Monograph 46, Cotsen Institute of Archaeology Press, University of California, Los Angeles.

la blanca and the middle formative

185

Guernsey, Julia 2006 Ritual and Power in Stone: The Per for mance of Rulership in Mesoamerican Izapan-Style Art. University of Texas Press, Austin. Guernsey, Julia 2010 A Consideration of the Quatrefoil Motif in Preclassic Mesoamerica. Res: Anthropology and Aesthetics 57/58 (spring/autumn): 75– 96. Guernsey, Julia, and Michael Love 2008a Cerámica y piedra: Relaciones entre alfarería, figurillas y escultura en el Preclásico de la Costa Sur. In XXI Simposio de Investigaciones Arqueológicas en Guatemala, 2007, edited by Juan Pedro Laporte, Bárbara Arroyo, and Héctor E. Mejía, pp. 953– 969. Ministerio de Cultura y Deportes, Instituto de Antropología e Historia, and Asociación Tikal, Guatemala. Guernsey, Julia, and Michael Love 2008b La Blanca in the Middle Formative Olmec World. Paper presented at the 2008 Annual Meeting of the American Anthropological Association, San Francisco. Hastorf, Christine, and Terrence D’Altroy 2001 The Domestic Economy, Households, and Imperial Transformations. In Empire and Domestic Economy, edited by T. D’Altroy and C. Hastorf, pp. 3–25. Kluwer, New York. Hendon, Julia A. 1999 The Pre- Classic Maya Compound as the Focus of Social Identity. In Social Patterns in Pre- Classic Mesoamerica, edited by David Grove and Rosemary Joyce, pp. 97– 125. Dumbarton Oaks Research Library and Collection, Washington, DC. Jackson, Thomas L., and Michael W. Love 1991 Bladerunning: Middle Preclassic obsidian exchange and the introduction of prismatic blades at La Blanca, Guatemala. Ancient Mesoamerica 2: 47– 59. Joralemon, Peter D. 1971 A Study of Olmec Iconography. Studies in Pre- Columbian Art and Archaeology No. 7, Dumbarton Oaks Research Library and Collection, Washington, DC. Joyce, Arthur, Laura Arnaud Bustamante, and Marc N. Levine. 2001 Commoner power: A case study from the Classic Period collapse on the Oaxaca coast. Journal of Archaeological Method and Theory 8 (4): 343–385. Joyce, Rosemary A. 2000 Heirlooms and Houses: Materiality and Social Memory. In Beyond Kinship: Social

186

emergent complexity

and Material Reproduction in House Societies, edited by Rosemary A. Joyce and Susan D. Gillespie, pp. 189–212. University of Pennsylvania Press, Philadelphia. Joyce, Rosemary A. 2003 Making something of herself: Embodiment in life and death at Playa de los Muertos, Honduras. Cambridge Archaeological Journal 13: 248–261. Lesure, Richard 1999 On the Genesis of Value in Early Hierarchical Societies. In Material Symbols: Culture and Economy in Prehistory, edited by John E. Robb, pp. 23– 55. Center for Archaeological Investigations Occasional Paper No. 16, Southern Illinois University, Carbondale. Lesure, Richard, and Michael Blake 2002 Interpretive challenges in the study of early complexity: Economy, ritual and architecture at Paso de la Amada, Mexico. Journal of Anthropological Archaeology 21: 1–24. Love, Michael W. 1991 Style and Social Complexity in Formative Mesoamerica. In The Formation of Complex Society in Southeastern Mesoamerica, edited by William R. Fowler, Jr., pp. 47– 76. CRC Press, Boca Raton, FL. Love, Michael W. 1999a Economic Patterns in the Development of Complex Society in Pacific Guatemala. In Pacific Latin America in Prehistory: The Evolution of Archaic and Formative Cultures, edited by Michael Blake, pp. 89– 100. Washington State University Press, Pullman. Love, Michael W. 1999b Ideology, Material Culture, and Daily Practice in Preclassic Mesoamerica: A Pacific Coast Perspective. In Social Patterns in Pre- Classic Mesoamerica, edited by David Grove and Rosemary Joyce, pp. 121– 153. Dumbarton Oaks Research Library and Collection, Washington, DC. Love, Michael W. 2002a Domination, Resistance, and Political Cycling in Formative Pacific Guatemala. In The Dynamics of Power, edited by Maria O’Donovan, pp.214–237. Center for Archaeological Investigations Occasional Paper No. 30, Southern Illinois University, Carbondale. Love, Michael W. 2002b Early Complex Society in Pacific Guatemala: Settlements and Chronology of the Río Naranjo, Guatemala. Papers of the New

World Archaeological Foundation No. 66, Brigham Young University, Provo, UT. Love, Michael W. 2006 Middle Formative Household Diversity at La Blanca, Guatemala. Paper presented at the Seventy-First Annual Meetings of the Society for American Archaeology, San Juan, Puerto Rico. Love, Michael W. 2010 Thinking Outside the Plaza: Varieties of Preclassic Sculpture in Pacific Guatemala and Their Political Significance. In The Place of Stone Monuments in Mesoamerica’s Preclassic Transition: Context, Use and Meaning, edited by Julia Guernsey, John E. Clark, and Bárbara Arroyo, pp. 149– 175. Dumbarton Oaks Research Library and Collection, Washington, DC. Love, Michael W. 2011 Cities, States and City-State Culture in the Late Preclassic Southern Maya Region. In The Southern Maya in the Late Preclassic: The Rise and Fall of an Early Mesoamerican Civilization, edited by Michael Love and Jonathan Kaplan, pp. 47–45. University Press of Colorado, Boulder. Love, Michael W. At press Early States in the Southern Maya Region. In Early Maya States, edited by Loa Traxler and Robert J. Sharer. University Museum, University of Pennsylvania Press, Philadelphia. Love, Michael, Donaldo Castillo Valdéz, Rene Ugarte, Brian Damiata, and John Steinberg 2005 Investigaciones arqueológicas en el Montículo 1 de La Blanca, Costa Sur de Guatemala. In XVIII Simposio de Investigaciones Arqueológicas en Guatemala, edited by Juan Pedro Laporte, Barbara Arroyo, and Hector E. Mejía, pp. 959– 969. Ministerio de Cultura y Deportes, Instituto de Antropología e Historia, Asociación Tikal, Foundation for the Advancement of Mesoamerican Studies, Guatemala. Love, Michael W., and Julia Guernsey 2007 Monument 3 from La Blanca, Guatemala: A Middle Preclassic earthen sculpture and its ritual associations. Antiquity 81: 920– 932. Love, Michael, and Thomas L. Jackson 1998 Intercambio y consumo de obsidiana en la Costa Sur Occidental durante el Preclasico Medio. In Taller Arqueologica de la Region de la Costa Sur de Guatemala, edited by Christa Schieber de Lavarreda, pp. 95– 108.

Ministerio de Cultura y Deportes, Guatemala. Marcus, Joyce 1998 Women’s Ritual in Formative Oaxaca: Figurine-Making, Divination, Death and the Ancestors. Memoirs of the Museum of Anthropology 33, University of Michigan, Ann Arbor. Marcus, Joyce 1999 Men’s and Women’s Ritual in Formative Oaxaca. In Social Patterns in Pre- Classic Mesoamerica, edited by David Grove and Rosemary Joyce, pp. 67– 96. Dumbarton Oaks Research Library and Collection, Washington, DC. McAnany, Patricia A. 1995 Living With the Ancestors: Kinship and Kingship in Ancient Maya Society. University of Texas Press, Austin. McAnany, Patricia A. 2002. Rethinking the Great and Little Tradition Paradigm From the Perspective of Domestic Ritual. In Domestic Ritual in Ancient Mesoamerica, edited by Patricia Plunket, pp.115– 119. Monograph 46, Cotsen Institute of Archaeology Press, University of California, Los Angeles. Neff, Hector 2005 El Sub-proyecto Clásico Tardío: Prospección por métodos geofísicos. In Proyecto La Blanca/El Ujuxte (PROBLALUX) Investigaciones Arqueológicas de La Blanca, Guatemala 2003–2004: Informe Final, edited by Michael Love and Donaldo Castillo. Report submitted to the Instituto de Antropología e Historia de Guatemala. Rosenswig, Robert M. 2005 From the Land Between Swamps: Cuauhtémoc in an Early Olmec World. Ph.D. dissertation, Department of Anthropology, Yale University, New Haven, CT. Rosenswig, Robert M. 2007 Beyond identifying elites: Feasting as a means to understanding Early Middle Formative society on the Pacific coast of Mexico. Journal of Anthropological Archaeology 26: 1–27. Sahlins, Marshall 1972 Stone Age Economics. Aldine, Chicago. Scott, James 1990 Domination and the Arts of Resistance: Hidden Transcripts. Yale University Press, New Haven, CT.

la blanca and the middle formative

187

Shook, Edwin M. 1971 Inventory of Some Preclassic Traits in the Highlands and Pacific Guatemala and Adjacent Areas. In Observations on the Emergence of Civilization in Mesoamerica, edited by Robert F. Heizer and John A. Graham, pp. 70– 77. Contributions of the University of California Archaeological Research Facility No. 11, Department of Anthropology, University of California, Berkeley. Spielmann, Katherine A. 2002 Feasting, craft specialization, and the ritual mode of production in small-scale societies. American Anthropologist 104: 195–207. Tejeda, Ana, and Hector Neff 2004 Laser Ablation-Inductively Coupled Plasma-Mass Spectrometry (LA-ICP-MS) Analysis of Fine Paste Ceramics of La

188

emergent complexity

Blanca, Guatemala. Paper presented at the Sixty-Eighth Annual Meeting of the Society for American Archaeology, Montreal. Wake, Thomas A., and Lady R. Harrington 2002 Vertebrate Faunal Remains From La Blanca, Guatemala. In Early Complex Society in Pacific Guatemala: Settlements and Chronology of the Rio Naranjo, Guatemala, edited by M. W. Love, pp. 237–252. Papers of the New World Archaeological Foundation No. 66, Brigham Young University, Provo, UT. Wattenmaker, Patricia 1998 Household and State in Upper Mesopotamia: Specialized Economy and the Social Uses of Goods in an Early Complex Society. Smithsonian Institution Press, Washington, DC.

part three

Beyond the Individual Study Area Grappling With Issues of Scale

NINE

Early Formative Transitions in Settlement and Subsistence at Chiquiuitan, Guatemala Molly Morgan

U

nderstanding early sedentary

society and the concomitant cultural transitions that occurred in the Formative period have been some of the most difficult problems of Mesoamerican archaeology. Along the Pacific coastal region of Mexico and Guatemala, increasing evidence demonstrates significant advances in subsistence, technology, trade, symbolic systems, and social relations. Much of the recent work on these topics is documented in the chapters of this volume. This chapter adds to the growing Pacific coast data by addressing settlement and subsistence at the site of Chiquiuitan in Guatemala. I describe data collected at Chiquiuitan between 2006 and 2009 from a small site at the southeastern edge of the coastal culture area. These data offer another example of Early and Middle Formative development. The area around Chiquiuitan was previously studied through a regional survey project that revealed an increasingly complex settlement during the Formative period (EstradaBelli 1998, 2002). More recent research builds upon the regional study by providing a

smaller-scale perspective on settlement and subsistence shifts through a focus on specific household developments at Chiquiuitan. This chapter describes the gradual development of Chiquiuitan from an Early Formative resource extraction locale to a fully sedentary, foodproducing community by the Middle Formative. The interpretations made here are important to consider in relation to those from the Soconusco region and other sites on the Guatemala coast. Taken together, the information that is presently being compiled from the various regions along the Pacific coast, including Chiquiuitan, offers a nuanced reconstruction of this critical period in Mesoamerican prehistory.

THE SITE OF CHIQUIUITAN Chiquiuitan is located in the Chiquimulilla coastal estuary ecological zone in southern Guatemala. North of the site lies the fertile coastal plain and the beginning of the slope to the Sierra Madre volcanic chain of the Guatemalan highlands less than 20 km away. Rivers

191

FIGURE 9.1 Map of Chiquiuitan showing the locations of mounds discussed in the text, as well as modern villages and landscape features

flow out of the mountains and form wetlands and lagoons that are connected by man-made and natural canals all along the coastal edge. The rivers nearest to Chiquiuitan are the María Linda, 15 km to the west, and the Esclavos, 15 km to the east. Chiquiuitan occupies the seasonally inundated land near the Chiquimulilla Lagoon, 3 km west of the modern town of Monterrico. Its soils are composed of sand and heavy clays. The hot tropical climate involves a six-month dry winter followed by concentrated rains, especially between July and September. The site itself is composed of twenty-two broad and flat earthen mounds varying in size between 50 and 150 m in diameter and 1 and 4 m in height. The site layout comprises a center of twenty mounds arranged in an irregular fashion in this relatively flat area of the coastal plain, including Mound 13, located slightly to the west, and two other mounds slightly removed to the east (Figure 9.1). The northern and western boundaries of the site are created by an artificial road and canals that delineate the southern extent of a man-made irrigation

192

beyond the individual study area

system that may have destroyed additional mounds of the site. The Chiquimulilla Lagoon is located to the east. Estuaries, mangrove forests, the Chiquimulilla Canal, and an extinct beach dune separate Chiquiuitan from the Pacific coast 1 km to the south. Barrier ridges such as the one to the south of Chiquiuitan are found all across the coastline and show evidence for progradation since the stabilization of the sea level after about 5500 b.c. (Kennett, Voorhies, and Martorana 2006; Voorhies 2004). The only previous survey and excavation project conducted at Chiquiuitan took place between 1995 and 1997, when Francisco Estrada-Belli identified the site as part of a regional Global Information System (GIS) study (Estrada-Belli 1998; Estrada-Belli, Kosakowsky, and Wolf 1998). That survey project helped define the occupational chronology of the region between the María Linda and Paz rivers and over the coastal plain and piedmont (Estrada-Belli 1999, 2002). Furthermore, Laura Kosakowsky conducted a preliminary ceramic analysis, which identified diagnostic attributes and established

a basic chronological framework (Kosakowsky and Estrada-Belli 1997; Kosakowsky, EstradaBelli, and Neff 1999). The chronology established by Estrada-Belli and Kosakowsky (Kosakowsky 2002; Kosakowsky, Estrada-Belli, and Pettit 2000) included three phases: the Early Formative Huiscoyol, the Early Formative Cangrejo, and the Middle Formative Tamarindo.* Chiquiuitan is the only known Huiscoyol-phase site between the two rivers María Linda and Paz (Estrada-Belli 1998). Based on surface remains and test pits, Estrada-Belli (1999) suggested that the site was initially composed of five widely spaced mounds, including Mounds 13, 24, 27, 34, and 36. In the Cangrejo phase, the community expanded from five to eleven inhabited mounds. New neighboring sites emerged at Pulido/Canal, Salinas Santa Rosa, Palosadentro, and Aguadulce along the coast to the east, and at Ujuxte1 on the piedmont. These are believed to be minor Early Formative sites with low occupation, each consisting of only a few mounds with Cangrejo sherds on the surface. By the Tamarindo phase, Chiquiuitan had become a significant center in the region, with nineteen identified mounds in the site core. Other sites that were previously established in the region grew in size through this phase as well, but none came near the density of settlement at Chiquiuitan (Estrada-Belli 1999).

PACHI RESEARCH While Estrada-Belli’s early project provided important preliminary results for understanding Chiquiuitan within the region, the Proyecto Arqueológico Chiquiuitan (PACHI) sought a more targeted understanding of this early community. By focusing on specific mounds that demonstrated evidence for early occupation, PACHI studied community developments at * For comparison with Soconusco chronology, see. Figure 1.3 in Chapter 1 of this book; for locations of sites discussed, see Figures 1.1 and 1.2. Dates discussed in this chapter are in calendar years.

the small scale. This research considered material evidence from household contexts to clarify how the founders of Chiquiuitan first established a sedentary community, as well as what subsistence choices were made throughout that process. In the first season, conducted in March and April of 2006, archaeologists excavated test pits in two of the mounds believed to be the earliest at the site (Morgan and Valle 2006). Excavations at Mound 24 penetrated 4 m of disturbed soil, uncovering mixed cultural materials that included sherds from all three ceramic phases. Excavations at Mound 27 located intact stratigraphic levels and identified twenty- one superimposed floors, fills, and other architectural features. This platform appears to have been constructed in the late Huiscoyol phase and was also occupied in the Tamarindo. The finds from this preliminary season aided in the augmentation of the site’s chronology (Morgan and Valle 2007a) and provided material for initial ceramic analysis. Research continued in 2007 with intensive excavations on additional mounds to examine residential architecture and the debris from domestic practices (Morgan and Valle 2007b). Levels dating to the late Cangrejo and Tamarindo phases at Mound 34 revealed wellpreserved house floors, hearths, middens, and circular features made of sand and clay (Velásquez López 2007). Excavations were not continued to lower levels at this location in order to conserve architectural features, in the hopes of future investigation. At Mound 13, five excavation units were placed over the mound (Figure 9.2) to gain greater coverage of the horizontal space (Morgan 2007b). Three of these excavations revealed layers of construction fill for the platform. Two units encountered floors and were expanded to reveal larger surface areas. Significant features include two flexed burials placed on floors and covered with construction fill, as well as middens, hearths, and storage pits. Excavations at this mound uncovered intact stratigraphic

settlement, subsistence at chiquiuitan

193

FIGURE 9.2 Topographic map of Mound 13 showing the locations of the five suboperations excavated by PACHI in 2007. Grid in meters; contour interval 10 cm; magnetic north at top. Map by Jon C. Lohse, PACHI 2008.

deposits from all three occupational phases. Huiscoyol-phase deposits from the lowest reaches of these excavations are especially important in evaluating the function of the site in the Early Formative. Lastly, recent research at Chiquiuitan aimed to bolster the radiocarbon chronology from the site. In addition to the two radiocarbon results from Estrada-Belli’s project (Kosakowsky, Estrada-Belli, and Pettitt 2000), thirteen new dates from carbon samples have provided important data needed to refine this chronology (Morgan and Valle 2007a) and consider it in a regional context. These dates have clarified the Huiscoyol and Cangrejo phases of the Early Formative occupation, fulfilling one of the main objectives of the recent project (Table 9.1).

Early Formative Huiscoyol Phase (1450–1250 b.c.) Microbotanical analysis of a core taken from sediments near the Chiquimulilla Lagoon has previously provided evidence for the presence of humans in the Chiquiuitan area and the concomitant effects on the Early Formative flo-

194

beyond the individual study area

ral landscape (John Jones, personal communication 2007). Those studies show a decrease in pollen content for arboreal species and an increase in charcoal frequencies from sediments dating to 1413 b.c., indicating human presence in the area. Mangrove forests also may have declined shortly thereafter. Although the mangroves were reduced in number, plants that grew in open habitats such as grasses, herbs, and sedges exhibit pollen increase, further demonstrating land clearing. In the work conducted by PACHI, eight excavation levels can be securely dated to the Huiscoyol phase. These deposits were located in the lowest levels of excavations in Mounds 13, 24, and 27 and consisted of thin layers of construction fills, a dirt floor, and a shallow pitlike round feature (Morgan 2007b; Morgan and Valle 2006; Velásquez López 2007). At Mound 13, the earliest levels were located in Suboperation 7-1, at the center of the mound, in which floor and fi ll layers are only 10 to 20 cm in thickness (Figure 9.3). Diagnostic attributes found on ceramic artifacts from Early Formative Chiquiuitan reflect general trends in material culture that

Tamarindo

AA86162

AA86165

CHI 07- 05- 03 Mound 13

CHI 05- 01- 09 Mound 27

CHI 04- 01- 09 Mound 24 Charcoal

Charcoal

Charcoal

Charcoal

Charred material

Charred material

Carbonized wood

Charcoal

Carbonized wood

Charred material

Charcoal

Charred material

Charcoal

Charred material

Charred material

Material

Uncalibrated date

1180 +/− 40 b.c. 1120 +/− 40 b.c. 1061 +/− 35 b.c. 1050 +/− 40 b.c. 1048 +/− 40 b.c. 990 +/− 50 b.c. 985 +/− 65 b.c. 967 +/− 36 b.c. 940 +/− 65 b.c. 920 +/− 40 b.c. 910 +/− 40 b.c. 895 +/− 62 b.c. 843 +/− 34 b.c. 840 +/− 34 b.c. 831 +/− 39 b.c.

Conventional radiocarbon age B.P.

3130 +/− 40 3070 +/− 40 3011 +/− 35 3000 +/− 40 2998 +/− 40 2940 +/− 50 2935 +/− 65 2917 +/− 36 2890 +/− 65 2870 +/− 40 2860 +/− 40 2845 +/− 62 2793 +/− 34 2790 +/− 34 2781 +/− 39

note: Dates OXA7779 and OXA7780 were published by Kosakowski, Estrada Belli, and Pettitt (2000).

Tamarindo

AA86160

CHI 05- 01- 07 Mound 27

Cangrejo

Tamarindo

AA86161

CHI 06- 01- 05 Mound 34

Cangrejo

Beta231366

Test Pit 7 Layer 15

CHI 07- 01-15 Mound 13

Cangrejo

CHI 07- 05- 07 Mound 13

Cangrejo

Cangrejo

AA86166

Test Pit 5 Layer 11

Beta231367

Cangrejo

OXA7779

CHI 04- 01-19 Mound 24

CHI 07- 01-22 Mound 13

CHI 04- 01-17 Mound 24

CHI 07- 04- 09 Mound 13

CHI 05- 01-19 Mound 27

CHI 07- 01-24 Mound 13

Field context

OXA7780

Huiscoyol

Beta226987

Cangrejo

Huiscoyol

AA86164

Beta226988

Huiscoyol

Beta226989

AA86163

Huiscoyol

Huiscoyol

Beta231368

Associated ceramic complex

Laboratory sample number

TABLE 9.1 Chiquiuitan Radiocarbon Data

2947–2846

2946–2854

2946–2856

3040–2872

3008–2926

3060–2940

3084–2929

3070–2946

3210–2996

3165–3006

3262–3142

3264–3142

3266–3162

3355–3255

3400–3328

Calibrated years B.P.

997–896 b.c.

996–904 b.c.

996–906 b.c.

1090–922 b.c.

1058–976 b.c.

1120–1000 b.c.

1134–979 b.c.

1120–996 b.c.

1260–1046 b.c.

1215–1056 b.c.

1312–1192 b.c.

1314–1192 b.c.

1316–1212 b.c.

1405–1305

1450–1378 b.c.

Cal 14 C 1 sigma

1016– 831 b.c.

1016– 840 b.c.

1021– 841 b.c.

1211– 890 b.c.

1131–914 b.c.

1190–1140 b.c., 1140–920 b.c.

1270–906 b.c.

1133–922 b.c.

1318–974 b.c.

1314–1002 b.c.

1386–1121 b.c.

1386–1123 b.c.

1386–1188 b.c.

1430–1256

1496–1311 b.c.

Cal 14 C 2 Sigma

FIGURE 9.3 Photo of the north profi le of the deepest section of the excavation unit in Suboperation 7-1, showing stratigraphic cultural levels discussed in text

are witnessed all along the Pacific coastal region, but with clear indications of local characteristics. Of the 26,886 sherds that were studied from excavations done in 2006 and 2007, forms could be securely identified in 3,239 (12 percent), and no whole vessels were found (Morgan 2007a). The Huiscoyol phase is characterized by the tecomate form (95 percent), a jar with no neck. Most tecomates exhibit a highly smoothed or burnished surface and are

unslipped or contain evidence of a self-slip (Figure 9.4, A–D). Some have a red painted band around the rim and are similar to the diagnostic Michis tecomates found in the Locona and Ocós phases at sites to the west (Arroyo 1994; Clark and Cheetham 2005; Pye 1995). Another surface treatment seen on Huiscoyol tecomates is plastic decoration, including shell edge impressions on the entire vessel exterior. Large, thick supports, a diagnostic trait for the regional Ocós phase,

FIGURE 9.4 (facing page) Photos and drawings of Formative-period ceramics from Chiquiuitan. Typical ceramics of the Huiscoyol complex: (A) Chiquimichis red- on-black globular tecomates with red rim bands, tool punctations, and an appliquéd and impressed ball decoration; (B) Chiquimichis red- on-black tear-shaped tecomates with fl at tool impressions and red rim bands; (C) Chiquimichis red- on-black sherd with rocker shell impression, circular smoothing, and red slip, and plain globular tecomates with red rim bands; and (D) Chiquimichis natural variety with zoned decorations. Typical ceramics of the Cangrejo complex: (E) Costeño tecomate with impressed “deer print” design; (F) signature appliquéd and molded designs of possible crab parts, frog or lizard face, and possible human face; (G) Cangrejo red effi gy of human face with Olmec-style eyebrow; and (H) raised and decorated rim band of a Cangrejo tecomate. Typical ceramics of the Tamarindo complex: (I) a variety of vessel forms, including (from top left) closed or incurving walled dish or bowl, water jar with tall neck, water jar with bolstered neck angle, double-angle open dish or bowl, closed dish or bowl with exterior bolstered upper wall and interior bolstered rim, flat bottomed open dish, open-walled or flaring dish or bowl with averted lip, water jar with wide neck angle; (J) Tamarindo natural incised rim sherd from a vertical-walled dish or bowl and with horizontal lines and cross-hatching design; (K) earspools found in Tamarindo deposits: top red, bottom black; (L) Tamarindo orange incised sherd; (M) Chiqui fi ne red incised sherd; (N) Tamarindo buff or cream incised sherd; and (O) drawing of a Tamarindo white incised rim sherd. Scale bar shows 25 cm total, with 5 cm units.

196

beyond the individual study area

TABLE 9.2 Marine Shell Fauna Counts and Frequencies Recorded for Different Time Phases

Huiscoyol Cangrejo

Tamarindo

Taxa

Common name

Mollusk type

Arcidae Anadara tuberculosa

Ark clam

Bivalve

Arcidae Anadara grandis

Ark clam

Bivalve

1

0.5

11.1

Arcidae Anadara tuberculosa

Ark clam

Bivalve

79

2.2

52.3

Arcidae Anadara mazatlanica

Ark clam

Bivalve

2

0.1

1.3

Noetiidae Noetia reversa

Clam

Bivalve

2

0.1

1.3

Veneridae Chione

Venus clam or cockle shell

Bivalve

1