The Archaeology of Greater Nicoya: Two Decades of Research in Nicaragua and Costa Rica
 1646421507, 9781646421503

  • 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

The Archaeology of Greater Nicoya

The Archaeology of Greater Nicoya Two Decades of Research in Nicaragua and Costa Rica

edited by

Larry Steinbrenner, Alexander Geurds, Geoffrey G. McCafferty, and Silvia Salgado

UNIVER SIT Y PRESS OF COLOR ADO

Louisville

© 2021 by University Press of Colorado Published by University Press of Colorado 245 Century Circle, Suite 202 Louisville, Colorado 80027 All rights reserved Manufactured in the United States of America



The University Press of Colorado is a proud member of the Association of University Presses.

The University Press of Colorado is a cooperative publishing enterprise supported, in part, by Adams State University, Colorado State University, Fort Lewis College, Metropolitan State University of Denver, Regis University, University of Alaska Fairbanks, University of Colorado, University of Denver, University of Northern Colorado, University of Wyoming, Utah State University, and Western Colorado University. ∞ This paper meets the requirements of the ANSI/NISO Z39.48-­1992 (Permanence of Paper). ISBN: 978-­1-­64642-­150-­3 (cloth) ISBN: 978-­1-­64642-­151-­0 (ebook) https://​doi​.org/​10​.5876/​9781646421510 Library of Congress Cataloging-­in-­Publication Data Names: Steinbrenner, Larry, editor. | Geurds, Alexander, 1974–­editor. | McCafferty, Geoffrey G., editor. | Salgado González, Silvia, editor. Title: The archaeology of Greater Nicoya : two decades of research in Nicaragua and Costa Rica / edited by Larry Steinbrenner, Alexander Geurds, Geoffrey G. McCafferty, Silvia Salgado. Description: Louisville : University Press of Colorado, [2021] | Includes bibliographical references and index. Identifiers: LCCN 2021015346 (print) | LCCN 2021015347 (ebook) | ISBN 9781646421503 (hardcover) | ISBN 9781646421510 (ebook) Subjects: LCSH: Indians of Central America—­Nicaragua—­History. | Indians of Central America—­Nicaragua—­Antiquities. | Indians of Central America—­Costa Rica—­Nicoya Peninsula—­History. | Indians of Central America—­Costa Rica—­Nicoya Peninsula—­Antiquities. Classification: LCC F1525 .A725 2021 (print) | LCC F1525 (ebook) | DDC 972.8/01—­dc23 LC record available at https://​lccn​.loc​.gov/​2021015346 LC ebook record available at https://​lccn​.loc​.gov/​2021015347 Cover illustrations: by Larry Steinbrenner (top); courtesy of Mi Museo, Granada, Nicaragua (bottom and spine).

LS: To Heather and Lyta, who have to put up with me AG: To the memory of Don Gustavo “Tapatilla” Villanueva Molina (1930–­2019), museum builder and pioneer of central Nicaraguan archaeology

Contents

Acknowledgments  xi 1. Introduction Larry Steinbrenner  3 Part 1: Redefining Greater Nicoya 2. Contact-­Era Pacific Nicaragua: Indigenous Groups and Their Origins Larry Steinbrenner  23 3. Indigenous Peoples of Pacific Nicaragua and Nicoya in the Sixteenth Century: A Historical Approach Eugenia Ibarra Rojas  47 4. A Critical Reevaluation of Pacific Nicaragua’s Late Period Chronology Larry Steinbrenner and Geoffrey G. McCafferty  67 Part 2: Projects and Surveys 5. The Managua Metropolitan Project: A Retrospective Frederick W. Lange  105 6. Twenty Years of Nicaraguan Archaeology: Results of the University of Calgary Projects Geoffrey G. McCafferty  125

7. The Development of Social Complexity in the Rivas Region, Pacific Nicaragua Karen Niemel Garrard  162 8. Social Practices at La Cascabel, A Village on Culebra Bay, ad 800–­1550 Ana Cristina Aguilar Vega  193 Part 3: Ceramics and Stone: Material Culture in Pacific Nicaragua 9. Polychrome Potting Traditions of Pacific Nicaragua, ad 800–­1300 Larry Steinbrenner  215 10. Ceramic Economy and Communities of Practice in Granada and Rivas, ad 1–­1300 Carrie L. Dennett  261 11. Exploring Technological Modifications: Variation of Non-­vessel “Objects” from El Rayo, Nicaragua Sacha Wilke  312 12. A Century in Stone: One Hundred Years of Lithic Analyses in Nicaragua Adam K. Benfer  331 13. The Development of Nicaraguan Rock Art Research Suzanne M. Baker  368 Part 4: Mortuary Practices in Greater Nicoya 14. Raising the Dead: Mortuary Patterns in Pacific Nicaragua Geoffrey G. McCafferty, Andrea L. Waters-­Rist, Sharisse McCafferty, Celise Fricker, and Jessica Manion  395 viii

CO N T EN T S

15. Osteoarchaeological Markers of Health and Identity at the Site of El Rayo, Nicaragua, ad 550–­1200 Andrea L. Waters-­Rist and Geoffrey G. McCafferty  433 16. Human Bone Artifacts: Ancestor Cults and Human Sacrifice in a Community of Mesoamerican Origin in Culebra Bay, Costa Rica Anayensy Herrera Villalobos and Felipe Solís Del Vecchio  469 Part 5: Conclusion 17. Greater Nicoya through the Looking-­Glass: Merging Culture History with Social Theory Alexander Geurds  505 Index  533

List of Contributors  549

CO N T EN T S

ix

Acknowledgments

In the preface to his own study of Nicaraguan and Costa Rican archaeology, the pioneering Central American archaeologist Samuel Lothrop observed that “it is needless to say that a book of this kind can be produced only with the aid and cooperation of many people” (1926:xx). What was true almost a century ago remains true now. As its subtitle suggests, this present volume is the culmination of archaeological research in Pacific Nicaragua and northwestern Costa Rica that extends back to the last decade of the twentieth century and that has involved scores of researchers. Its more recent genesis, however, can be traced to an early morning electronic symposium organized by Geoff McCafferty titled “Re-­ conceptualizing Nicaraguan Prehistory” at the Society for American Archaeology Annual Meeting in Sacramento in 2011. Following that session, most of the participants and several other interested Central Americanists gathered together over coffee to discuss what should be done next—­a conversation that continued through lunch, and then on for much of the rest of the day, and then over the next several years until the ultimate publication of this book. The editors would like to thank all the original symposium participants, our discussant Fred Lange, and Karen Bruhns for stimulating conversation. We also wish to heartily thank all of the contributors to this volume. Additional thanks to Paul Healy, who literally “wrote the book” on the archaeology of Nicaragua’s Rivas region; Adam Benfer, for his assistance in compiling

bibliographical information and for generating maps for several chapters; and to Heather Marcovitch, Carrie Dennett, and Danny Zborover, for their input on assembling the manuscript. In Nicaragua specifically, we would like to thank Mi Museo, the late Peder Kolind, and the invaluable Jorge Zambrana, as well as the Instituto Nicaragüense de Cultura (INC) and its archae­ology department for their formal support of our research. Geoff McCafferty would like to acknowledge the University of Calgary and the Social Sciences and Humanities Research Council of Canada (SSHRC) for funding several years of archaeological research in Nicaragua. Larry Steinbrenner would like to acknowledge Red Deer Polytechnic’s Professional Development committee for providing funding to work on this project. Alexander Geurds would like to thank the Netherlands Organisation for Scientific Research (NWO) for enabling him to work on this volume through his VIDI grant (2013–­2019). Collectively, we are all grateful to the two anonymous reviewers who provided enthusiastic feedback and useful critiques and to Darrin Pratt, Charlotte Steinhardt, and other staff of the University Press of Colorado for their support and guidance. La trayectoria milenaria de los pueblos indígenas de  Nicaragua y sus logros históricos han sido la mayor fuente de nuestra motivación para realizar este libro. Esta motivación también la han tenido los arqueólogos y ciudadanos nicaragüenses y de otras nacionalidades quienes han contribuido a la protección y al estudio del patrimonio histórico de Nicaragua.1 REFERENCES

Lothrop, Samuel K. 1926. Pottery of Costa Rica and Nicaragua. 2 vols. Memoir 8. New York: Museum of the American Indian, Heye Foundation.

1. The editors of this volume have been inspired by the millennia-­long history and achievements of the Indigenous people in the territory that we now know as Nicaragua. We also wish to acknowledge the many Nicaraguan citizens and archaeologists, as well as scholars from other nations, who have collectively contributed to the study and protection of Nicaragua’s cultural patrimony.

xii

Acknowledgments

The Archaeology of Greater Nicoya

1 Welcome to The Archaeology of Greater Nicoya, the first Introduction major publication in more than two decades to present new research focusing on one of Lower Central Larry Steinbrenner America’s most unique and confounding regions. The Archaeology of Greater Nicoya is a book that is long overdue. Scholars of Greater Nicoya will doubtlessly already be familiar with the handful of essential previous collections that have incorporated research from this archaeological subarea, a combined literature that is so sparse that it is not uncommon to refer its essential volumes by color: the “orange book” (i.e., The Archaeology of Lower Central America, Lange and Stone 1984), the “blue book” (Wealth and Hierarchy in the Intermediate Area, Lange 1992), the “green book” (Archaeology of Pacific Nicaragua, Lange et al. 1992), and so on. Inasmuch as the most recent of these collections—­Paths to Central American Prehistory (Lange 1996a)—­was published in 1996 and focused only partially on Greater Nicoya, we are confident that even specialists in Greater Nicoya will find more grist than chaff in the present volume, which is based, for the most part, on research carried out in the twenty-­first century. For nonspecialists, we hope that The Archaeology of Greater Nicoya will serve as a useful introduction to this fascinating region and perhaps inspire you to dive even deeper into the still-­ murky waters of its culture history. The uninitiated reader might well ask, “What exactly is Greater Nicoya, and what is exceptional about its archaeology?” As originally defined in the early 1960s DOI: 10.5876/9781646421510.c001 3

Figure 1.1. Greater Nicoya, indicating key sites discussed in this volume. The existence and potential boundaries of this archaeological subarea continue to be subjects of debate. Map by Larry Steinbrenner.

by archaeologist Albert Norweb (1961, 1964), Greater Nicoya comprised an archaeological subarea incorporating the supposedly “Mesoamericanized” portions of Pacific Nicaragua—­ effectively, Nicaragua’s entire west coast and its associated lakes—­and former Nicaraguan territory in northwestern Costa Rica, including the entire Nicoya Peninsula and most of the rest of the modern province of Guanacaste (figure 1.1). This subarea was associated with a supposedly homogenous material culture characterized especially by stunning “Nicoya Polychrome” pottery (discussed in detail in chapters 9 and 10), distinctive mortuary practices (chapters 14–­16), and a unique stone sculpture tradition, among other things. This material culture showed links to the north—­especially to El Salvador and Honduras but also to regions even farther afield—­and was also supposed to differentiate the subarea from non-­ Mesoamericanized territories to the east (in the highlands and Caribbean 4

S T EI N B R EN N ER

watersheds of Nicaragua and Costa Rica) and south, which were collectively associated with different, “southern/Circum-­Caribbean-­oriented” traditions of material culture. Norweb’s characterization was precedented centuries earlier by the first European reports of the country (cf. González Dávila 2002), which had described this area as “another Yucatán,” and reported that it was dominated by several different groups claiming Mesoamerican origin, including speakers of Otomanguean and Nahua languages (see chapters 2 and 3). Not surprisingly, this earliest recognition of connections between Greater Nicoya and lands to the north led early archaeological research in the nineteenth century (e.g., Squier 1852, 1853) to uncritically view Greater Nicoya as the southernmost extension of a Mesoamerican culture area—­a tendency that continued well into the twentieth century and that still persists, especially in Nicaragua. Studies that treat Greater Nicoya as a sort of “Mesoamerica Lite” have become increasingly rare as modern scholars have recognized that it is problematic to assume either that the Mesoamericans who migrated to Nicaragua and Costa Rica in prehistory managed to transport their own cultures completely intact and unchanged or that the autochthonous Chibchan-­affiliated populations that seemingly inhabited the area before the migrations (and who afterward continued to occupy the surrounding territories to the east and south) simply vanished following the arrival of the Mesoamericans or left no mark on the cultures that supplanted their own. As Fred Lange (e.g., Lange 1971, 1978) and other influential Central Americanists (e.g., Fonseca Zamora 1994; Hoopes 2005; Hoopes and Fonseca Zamora 2003) have long argued, and as the various scholars who have contributed to this volume (e.g., chapter 6) now recognize, the Mesoamerican aspects of Greater Nicoya have likely been greatly overemphasized in past research at the expense of earlier-­appearing lifeways. Indeed, models of Greater Nicoya prehistory that are based exclusively on old-­fashioned notions emphasizing population replacement and that make no attempt to explore autochthonous cultural development (see, e.g., chapter 3) ought to be finally consigned to one of the shell middens that are so ubiquitous along Greater Nicoya’s Pacific coasts. It is time for a new paradigm in the archaeology of Greater Nicoya. Even as modern scholars have come to accept that Greater Nicoya’s “Mesoamericanization” was likely not as complete as was once believed—­sometimes even going so far as to argue that Greater Nicoya was not part of Mesoamerica at all (e.g., Lange et al. 1992:272)—­they have also come to embrace the idea that there was actually substantial material culture variation across the entire extent of Greater Nicoya, which has proven to be a less homogenous I ntroduction

5

“subarea” than it was once imagined to be. By the 1970s, numerous scholars (e.g., Healy 1974, 1980; Lange 1971, 1978; Sweeney 1975) were already recognizing important differences in the archaeological records of Greater Nicoya’s northern and southern sectors—­that is, Pacific Nicaragua and Guanacaste-­ Nicoya.1 These differences should surprise no one, given the diversity of cultural groups that are documented as living in the area in ethnohistoric sources (chapters 2 and 3), and the evidence that has accumulated over subsequent decades (much of which is presented in this present volume) has made it even clearer that it is problematic to treat Greater Nicoya as a distinctive, full-­fledged “culture area” in the classic anthropological sense. While there may be a loose correlation between “Greater Nicoya” and the broad distribution of “Nicoya polychromes,” it no longer seems prudent to infer a base level of cultural homogeneity from this: Greater Nicoya now appears to be yet another region where we need to think twice about equating pots with people. In light of this realization, it is worth noting that cultural homogeneity would not necessarily be a prerequisite for the existence of a constellation of practice linking individual communities of practice in Greater Nicoya to one another as well as to other communities of practice located beyond the boundaries of Greater Nicoya. A constellation of practice—­as proposed for Lower Central America by Rosemary Joyce (2016; following Roddick 2009 and Wenger 1998)—­might have comprised groups who were not necessarily closely related culturally but who might have produced similar forms of material culture because they were connected through other kinds of interaction, such as trade or competition. So, if we can no longer treat Greater Nicoya as a homogeneous culture area, is it time to give up the label altogether? While that argument has been made by some scholars in the past (including one of the coeditors of this volume; i.e., Salgado González 1996), and is made here by at least one contributor (cf. chapter 3), perhaps we would be better served by recognizing that it remains useful as a sort of conceptual shorthand denoting a contiguous geographical region—­that is, the coastal plains of Nicaragua and Costa Rica west of the Central Highlands—­that seems archaeologically “different” from surrounding 1. These differences partially explain why Mesoamerican influence continued to be emphasized in work that was Nicaragua based (e.g., Paul Healy’s 1974 dissertation, which was based on excavations carried out by Gordon Willey and Norweb between 1959 and 1961), while it was also being downplayed in most studies that were based on fieldwork in more distant Costa Rica (e.g., dissertations by Lange 1971; Sweeney 1975; and Creamer 1983), where evidence for the persistence of Chibchan traditions seemed to be more visible.

6

S T EI N B R EN N ER

regions elsewhere in these two countries, even if we have moved beyond trying to explain this difference wholly in terms of Mesoamerican cultural affiliations or material culture homogeneity. If we choose to think of Greater Nicoya in these modified terms, then there can be no more objection to discussing the “archaeology of Greater Nicoya” than there would be to discussing the archaeology of any other well-­established geographic locale, such as “Nicaragua,” or “the Ulua Valley” or “Outer Mongolia.” This volume is divided into four parts. The chapters in Part 1, “Redefining Greater Nicoya,” are foundational in that they provide background information concerning Greater Nicoya’s Indigenous populations and the development of the standard chronological sequence that is used by most archaeologists studying the subarea. My chapter 2, “Contact-­Era Pacific Nicaragua: Indigenous Groups and Their Origins,” introduces the diverse Mesoamerican-­and Chibchan-­affiliated groups that dominated the cultural landscape of Pacific Nicaragua and northwestern Costa Rica at the time of contact and goes on to briefly review the various lines of evidence—­ethnohistoric, linguistic, and archaeological—­that have contributed to the ongoing debates regarding the nature and timing of the migrations that brought Mesoamericans (or Mesoamerican influences) into Greater Nicoya starting more than a millennium ago. Chapter 3, “Indigenous Peoples of Pacific Nicaragua and Nicoya in the Sixteenth Century: A Historical Approach,” by prominent Costa Rican ethnohistorian Eugenia Ibarra Rojas, argues that conceptualizing Greater Nicoya as a well-­delimited “culture area” impairs our ability to truly understand the dynamic social relationships that bound Chibchan and Mesoamerican-­originating populations together. In Ibarra’s view, precontact Pacific Nicaragua and northwestern Costa Rica represented a pluricultural confluence space, a crossroads for many different cultural groups that varied substantially through time and that were often intimately connected to other groups beyond the traditional boundaries of Greater Nicoya. Ibarra draws on colonial documents as well as archaeological evidence to explore not only how Chibchan practices and traditions might have endured following the arrival of Mesoamericans, but also how precontact sociopolitical structures could also have merged with and endured in later colonial institutions, such as cofradías. The final chapter in Part 1, chapter 4, “A Critical Reevaluation of Pacific Nicaragua’s Late Period Chronology,” is a key chapter that follows up on an earlier-­published paper by two of the coeditors of this volume (i.e., McCafferty I ntroduction

7

and Steinbrenner 2005), which reported a dozen new radiocarbon dates for the site of Santa Isabel in Rivas, by far the largest collection of 14C dates ever reported for a single archaeological site in Nicaragua. On the basis of this evidence and a reexamination of the preexisting radiocarbon database, that paper made the provocative and paradigm-­shifting argument that the long-­ established and widely accepted standard chronological sequence (see tables 4.1 and 4.3, this volume) for Greater Nicoya was seriously flawed. In particular, the cumulative 14C evidence indicated that virtually all archaeological sites in Greater Nicoya that had been previously attributed—­based on the presence of misdated diagnostic ceramics—­to the Ometepe period (conventionally dated ad 1350–­1550) were actually centuries older and more likely dated to the earlier Sapoá period (conventionally dated ad 800–­1350). Needless to say, the implications of this claim—­which, while accepted by many modern scholars and strongly endorsed in most chapters in this volume—­remain controversial but are far reaching. We examine some of these implications in this new chapter, which begins with a historical review of the development of the Greater Nicoya chronological sequence (in an attempt to understand how the current version of this sequence came to be out of sync with the actual 14C database upon which it is ostensibly based) and then revisits and reinforces the argument of the earlier paper with additional radiocarbon dates from more recent archaeological work in the Nicaraguan departments of Rivas and Granada. The chapters in Part 2, “Projects and Surveys,” provide summary overviews of major international archaeological projects in Pacific Nicaragua (several of which were directed by one of the coeditors of this volume), a major survey of archaeological sites in the department of Rivas, and a village site located on Costa Rica’s Bay of Culebra. Chapter 5, “The Managua Metropolitan Project: A Retrospective,” by Fred W. Lange, presents the first published summary in English of an important three-­year project directed by Lange that ran from 1995 to 1997 and that centered on Nicaragua’s capital of Managua and its environs. This project involved both North American and Nicaraguan archaeologists and—­unlike many earlier projects in Nicaragua in the mid-­twentieth century—­was guided by a specific interest in exploring local cultural evolution rather than Mesoamerican influence. While reports on the project’s first two seasons were originally published locally in Spanish (Lange 1995, 1996b) and later posted online on a now-­defunct website, the third season report was never released after funds earmarked for its publication were redirected to provide relief for Hurricane Mitch in 1998. As a result, information on the Managua Project has generally remained difficult to access, especially for 8

S T EI N B R EN N ER

scholars outside of Nicaragua.2 Lange’s retrospective succinctly summarizes key sites and findings of this pioneering project. Chapter 6, Geoffrey G. McCafferty’s “Twenty Years of Nicaraguan Archae­ ology: Results of the University of Calgary Projects” provides an overview of the various archaeological projects that McCafferty has directed in Pacific Nicaragua since 2000. Originally interested in the Mesoamerican migration question, McCafferty’s earliest University of Calgary–­based work in Nicaragua from 2000 to 2005 focused on domestic contexts at the long-­occupied Rivas-­ area site of Santa Isabel, first excavated in 1959 by Gordon Willey and Norweb (Healy 1974, 1980; Norweb 1961, 1964) and later identified by Karen Niemel (2003) as being the largest archaeological site in Rivas. Finding little compelling evidence for distinctively Mesoamerican lifeways at this site (which, contrary to the expectation that it was a contact-­period site, was apparently abandoned near the end of the Sapoá period), McCafferty relocated north to the department of Granada for a second project between 2008 and 2010. This latter project, which also focused on domestic contexts and similarly found only limited evidence for Mesoamerican migration, explored sites previously identified and visited by Silvia Salgado González (1996) during her dissertation survey of the department, including Tepetate, the dominant center in the area from the Sapoá period until contact (largely destroyed by urban development in the modern city of Granada), and El Rayo, a minor though better-­ preserved site on Lake Nicaragua’s Asese Peninsula, which provided evidence of the transition to the Sapoá period from the earlier Bagaces period (ad 300–­800). McCafferty’s subsequent work in the Granada area has included not only additional field seasons at El Rayo but also an investigation of the much-­ debated ceremonial site of Sonzapote on Zapatera Island in Lake Nicaragua, long renowned for its monumental sculptures. As discussed in this chapter, McCafferty’s projects have both helped to redefine the standard chronological sequence for Pacific Nicaragua by greatly expanding the radiocarbon database (see chapter 4) and have produced unique insights into many aspects of life in precontact Greater Nicoya, ranging from foodways to architectural traditions to mortuary patterns to personal adornment to craft specialization. Chapters 9, 10, 11, 14, and 15 in the present volume are all based on research linked to McCafferty’s projects. 2. Additional information about the Managua Project can also be found in master’s theses by Dickau (1999), Finlayson (1998), and Aggen (2007). The former study focused on ethnobotany, while the latter two were lithic oriented. Note that much of lithic analysis associated with this project (including work by Finlayson and Aggen) is also summarized in chapter 12, this volume. I ntroduction

9

Chapter 7, “The Development of Social Complexity in the Rivas Region, Pacific Nicaragua,” by Karen Niemel Garrard, provides a summary of Garrard’s yearlong survey of the lacustrine coast of the department of Rivas in the late 1990s, the foundation for her subsequent doctoral dissertation (Niemel 2003). Modeled after the survey methodology used by Salgado González (1996) in her own dissertation survey of the adjacent department of Granada as part of an effort to formulate a comparative settlement database incorporating multiple regions,3 Garrard identified four dozen sites in a 270 km2 area delimited by the Ochomogo River in the north, the modern city of Rivas in the south, and the Pan-­American Highway and coast of Lake Nicaragua to the west and east, respectively. The development of these sites over more than two millennia, from the earliest Orosí period to the most recent Ometepe period (as defined in the standard Greater Nicoya chronological sequence), is traced, and the emergence of certain key settlements—­most notably the eventual regional center of Santa Isabel—­is chronicled. The chapter concludes with a discussion of the challenges inherent in comparing the results of the Rivas survey with those obtained via other survey projects carried out in Nicaragua and northwestern Costa Rica since the 1970s, owing to varying research interests, differences in survey methodologies, and a general lack of fieldwork supporting survey work—­a lack that work in the twenty-­first century is only now beginning to address. Chapter 8, “Social Practices at La Cascabel, A Village on Culebra Bay, ad 800–­1550,” by Ana Cristina Aguilar Vega, presents the results of a smaller-­scale research project at a nucleated village site on the Papagayo Peninsula, located on the north side of the Bay of Culebra in Costa Rica’s Guanacaste Province. Research focusing on the Bay of Culebra (also the focus of Herrera Villalobos and Solís Del Vecchio’s chapter 16) has contributed enormously to our general understanding of Greater Nicoyan prehistory since the 1960s, with important projects being carried out at the sites of Papagayo, Nacascolo, and Vidor (e.g., Abel-­Vidor 1980; Baudez et al. 1992; Gutiérrez González 1993, 1998; Hardy 1983, 1992; Lange, Ryder, and Accola 1986; Moreau 1980, 1983, 1984; Norr 1991; Obando 1995; Solís Del Vecchio 1998, 2002; Solís Del Vecchio and Herrera 3. Salgado’s model was also emulated in a third doctoral dissertation settlement pattern survey carried out by Nicaraguan archaeologist Román Lacayo (2013) at roughly the same time. Román’s project focused on several municipalities in the department of Masaya, adjacent to both Rivas and Granada. Although a summary of this work is not provided in the present volume, the complete settlement dataset and dissertation can be accessed via the University of Pittsburgh Comparative Archaeology Database (http://​www​.cadb​.pitt​.edu/).

10

S T EI N B R EN N ER

Villalobos 2009; Wallace and Accola 1980). La Cascabel was one of several similar contemporary villages around the bay that appear to have been occupied by Chorotega migrants from Mesoamerica from the Sapoá period (ca. ad 800) until contact. Aguilar reports on the remains of shell deposits, craft production, architecture, sculptures, and funeral contexts and explores how all of these things played roles in maintaining a complex social hierarchy in the region. Part 3, “Ceramics and Stone: Material Culture in Pacific Nicaragua,” is the longest section in this volume and focuses almost exclusively on research carried out in Greater Nicoya’s northern sector (i.e., in Pacific Nicaragua or other parts of Nicaragua). The first two chapters deal with the distinctive archaeological ceramics that have long been recognized (as previously noted) as a hallmark of the Greater Nicoya archaeological subarea. My own chapter in this section, chapter 9, “Polychrome Potting Traditions of Pacific Nicaragua, ad 800–­1300,” summarizes some of the most essential findings of my doctoral dissertation research (Steinbrenner 2010). This research focused primarily on ceramics from the Santa Isabel site, which had previously provided much of the dataset used in Paul Healy’s influential Archaeology of the Rivas Region (1980). Healy’s book, which has served for many years as the de facto “handbook” for ceramics studies in Pacific Nicaragua, employed a type-­variety-­based typology to classify Rivas ceramics. My own Santa Isabel database (which included monochrome as well as polychrome ceramics, though only the latter are discussed in chapter 9) was also supplemented with a secondary database of polychromes from the Managua-­area San Cristóbal site (previously excavated by Wyss 1983) and a tertiary database of nearly 1,600 complete vessels derived from museum and private collections. My study of this material, which incorporated analyses of aspects of variation in vessel shape and form (such as orifice size and wall thickness) as well as decorative variation, and which was also informed by the modified chronological sequence discussed in chapter 4, allowed me to refine and critique Healy’s original typology while maintaining a type-­variety-­based approach. It allowed me as well to argue for a significant degree of similarity and continuity between supposedly unrelated Nicoya Polychrome ceramic types that have usually been treated, in previous research, as products of rival cultural entities, such as the aforementioned Otomanguean Chorotega and the Nicarao, the most prominent of Greater Nicoya’s ethnohistorically documented Nahua groups. Chapter 10, “Ceramic Economy and Communities of Practice in Granada and Rivas, ad 1–­1300,” based on Carrie L. Dennett’s recently completed doctoral dissertation at the University of Calgary (Dennett 2016), approaches the I ntroduction

11

study of Nicaraguan ceramics from a very different yet complementary perspective. Departing from the more traditional focus on vessel form and decoration that has characterized most previous ceramics-­oriented studies in Greater Nicoya (including my own work), Dennett’s work focuses instead on the sourcing and production/manufacture of ceramics. This approach (long overdue in this part of Central America) employs both petrographic analysis and instrumental neutron activation analysis (INAA) that builds on pioneering INAA-­ based studies of Ronald Bishop and Fred Lange (Bishop, Lange, and Lange 1988; Bishop et al. 1992; Bishop and Lange 2013) in the 1980s. Dennett’s robust dataset (which derived from McCafferty’s University of Calgary projects at El Rayo, Tepetate, and Santa Isabel) included 150 ceramic samples that were submitted for INAA compositional testing and nearly 200 thin sections from sherds that were subjected to petrographic analysis. Her findings challenge many long-­standing preconceptions concerning relationships between different key ceramic types, where and when these types were actually made, and who might have made them, and they also provide sophisticated new insights into the configuration of consumer distribution networks and communities of practice in Pacific Nicaragua over a span of more than a thousand years. Chapter 11, “Exploring Technological Modifications: Variation of Non-­ vessel ‘Objects’ from El Rayo, Nicaragua,” by Sacha Wilke, takes an interesting look at some of the less-­glamorous and least-­often-­studied artifacts that are commonly found in the archaeological record of Greater Nicoya and considers the implications of differing frequencies of these objects in different contexts, both spatial and temporal. Among the more utilitarian objects examined in this study are net sinkers (fairly ubiquitous at sites on the shores of Lake Nicaragua, such as El Rayo); ceramic balls (possibly rattles from hollow vessels or projectile pellets); various types of reworked sherds; and an assortment of other tools that were likely employed in textile production, including spindle whorls, weaving picks, needles, and awls. Other artifacts less likely to represent tools, such as figurines and items of personal adornment (including pendants, earspools, and beads), are also discussed. The final two chapters in this part both provide historical perspectives on two very different archaeological categories based on stone. Chapter 12, “A Century in Stone: One Hundred Years of Lithic Analyses in Nicaragua,” by Adam K. Benfer, reviews the literature focusing on Nicaraguan chipped and ground stone tools from the nineteenth century—­when collectors competed to supply “primitive” tools to American and European museums but made little attempt to understand these tools—­to the twenty-­first century, when processualist-­oriented analyses have become much more commonplace and 12

S T EI N B R EN N ER

many Nicaraguan archaeologists have adopted a standardized methodology inspired by the work of Georges Laplace (1974). Following this review, Benfer outlines research priorities for future lithic-­oriented work, such as the need to identify sources of raw materials, better understand operational sequences, conduct more usewear analyses, and develop a comparative database for identifying microbotanical residues. Chapter 13, “The Development of Nicaraguan Rock Art Research” by Suzanne M. Baker, provides an overview of work in a field that has often been marginalized in Greater Nicoya studies, in spite of the ubiquity of petroglyphs and pictographs in archaeological contexts distributed throughout Nicaragua. Baker provides detailed summaries of important work in this field by international scholars Wolfgang Haberland, Dominique Rigat and Franck Gorin, Laura Laurencich Minelli and Patrizia Di Cosimo, and others, as well as Nicaraguan scholars Joaquín Matilló Vila (the avocational pioneer of modern rock art studies in the mid-­twentieth century) and others and archaeologists associated with Nicaragua’s National Museum, including Jorge Zambrana Fernández and Rigoberto Navarro Genie. Baker, who has studied Nicaragua rock art since the 1980s, also summarizes her own work directing the Ometepe Archaeological Project, which has surveyed and identified more than 100 rock art sites on Ometepe Island in Lake Nicaragua (see Baker 2010 for an even more detailed account). Her chapter concludes with a call for the development of a common methodology for studying rock art in Nicaragua to facilitate future research. The final section of this volume, Part 4, “Mortuary Practices in Greater Nicoya,” begins with a comparison of mortuary practices at three sites on the shores of Lake Nicaragua in Pacific Nicaragua and then moves on to more specific studies of human remains from one of these lacustrine sites and at a Pacific Coast site in Costa Rica. Chapter 14, “Raising the Dead: Mortuary Patterns in Pacific Nicaragua”—­by Geoffrey G. and Sharisse McCafferty, Andrea L. Waters-­Rist, Celise Fricker, and Jessica Manion—­examines evidence from the three key sites excavated by the University of Calgary’s Santa Isabel and Granada projects: Santa Isabel, Tepetate, and El Rayo. Burial practices and offerings for all three sites are summarized, including primary extended burials in simple earthen and stone-­lined graves and secondary burials of partial remains in large ceramic vessels. The results of osteological analysis vis-­à-­vis age-­at-­death, sex, demography, dentition, and pathology from Santa Isabel and El Rayo—­the two sites where human remains were well preserved—­are also presented and compared. Chapter 15, “Osteoarchaeological Markers of Health and Identity at the Site of El Rayo, Nicaragua, ad 550–­1200,” by Andrea L. Waters-­Rist (Western University) and Geoffrey G. McCafferty, I ntroduction

13

looks more closely at skeletal and dental data from this habitation and mortuary site, which provided a notably larger sample of human remains for study than Santa Isabel. As a site spanning the transitional period when historically documented groups of Mesoamerican origin appear to have first arrived in Greater Nicoya, El Rayo provides unique insights into the health and dietary practices of the peoples who lived on the lakeshore both prior to and following the transition. The last chapter in this section, “Human Bone Artifacts: Ancestor Cults and Human Sacrifice in a Community of Mesoamerican Origin in Culebra Bay, Costa Rica,” by veteran Costa Rican archaeologists Anayensy Herrera Villalobos and Felipe Solís Del Vecchio, provides a fascinating and detail-­rich study of how the bones of ancestors (as well as possible enemies) recovered from the Jícaro site were transformed into wearable artifacts. Jícaro, like chapter 8’s La Cascabel site, is located on the much-­studied Papagayo Peninsula and has been identified by its excavators as being a Chorotega site during the Sapoá period. Herrera and Solís discuss burial practices and artifacts that range from beads made from human teeth to jaw and maxillary bones worn as pendants and bracelets. Evidence for manufacturing techniques is also covered. The chapter concludes with a discussion of how these artifacts mirror Mesoamerican practices related to ancestor cults and human sacrifice and how they might have been ritually used by male warrior-­leaders to emphasize their roles in the new social context created by the arrival of migrant populations in Greater Nicoya. Wrapping up The Archaeology of Greater Nicoya is chapter 17, “Greater Nicoya through the Looking-­Glass: Merging Culture History with Social Theory” by Alexander Geurds, one of the coeditors of the present volume. Geurds, who has been directing archaeological research in Central Nicaragua on the eastern flanks of Greater Nicoya since the mid-­2000s, reflects upon several common threads running throughout the book from a useful etic perspective, and, looking forward, proposes some new avenues of investigation for archaeologists working in Greater Nicoya, such as the roles of land-­and waterscapes and the sorely neglected transition to the colonial period. In closing, Geurds anticipates a future when archaeological research in Pacific Nicaragua and Nicoya will be rooted in empirical research and informed by social theory in equal measure. As even a cursory reading of the various chapters in this volume will make clear, our knowledge of the prehistory of Pacific Nicaragua and northwestern Costa Rica is growing rapidly, and many of the simplest facts that were taken 14

S T EI N B R EN N ER

for granted in the last century are now being challenged and replaced with a deeper, more sophisticated understanding of the region that we call Greater Nicoya. With an ever-­increasing number of investigators taking an interest in the archaeology of this region (not to mention the rest of Nicaragua and Costa Rica, areas that are also experiencing a boom in research), the twenty-­first century promises to be an exciting time for Greater Nicoya scholarship, but there are challenges ahead. Nicaragua’s current political turmoil (which began with protests against the government in the spring of 2018 and is ongoing as of this writing) and the much-­heralded yet perpetually delayed construction of a massive transoceanic canal across southern Nicaragua (first announced in 2013) are not the least of these. While an exploration of the potential environmental and economic consequences of the canal project in particular would take us beyond the scope of traditional archaeology, we cannot help but recognize that such an enormous undertaking, though potentially offering opportunities for extensive archaeological exploration via mitigation work, would almost assuredly result in the destruction of dozens or even hundreds of archaeological sites (many of which have yet to be identified) as well as the loss of everything that they might reveal. It is our fervent hope that such a worst-­case scenario will never play out, that the archaeological record of Nicaragua and Costa Rica will receive the attention that it so richly deserves, and that The Archaeology of Greater Nicoya will not prove to be the last major publication to focus on this exciting region, but the first of many. REFERENCES

Abel-­V idor, Suzanne. 1980. “Dos hornos precolombinos en el Sitio Vidor Bahia Culebra, Guanacaste.” Vínculos: Revista de Antropología del Museo Nacional de Costa Rica 6(1–­2):43–­50. Aggen, Kerry L. 2007. “Provenance Sourcing of Obsidian Artifacts from Managua, Nicaragua, Using Trace Element Geochemistry.” MA thesis, Colorado School of Mines, Golden. Baker, Suzanne M. 2010. The Rock Art of Ometepe Island, Nicaragua: Motif Classification, Quantification, and Regional Comparisons. Paris Monographs in American Archaeology 25. BAR International Series 2084. Oxford: Archaeopress and British Archaeological Reports. Baudez, Claude F., Nathalie Borgnino, Sophie Laligant, and Valerie Lauthelin. 1992. Papagayo: Un hameau précolombien du Costa Rica. Edited by Claude F. Baudez and Michael D. Coe. Paris: Centre d’études mexicaines et centraméricaines, Editions Recherches sur les Civilisations. I ntroduction

15

Bishop, Ronald L., and Frederick W. Lange. 2013. “Frederick R. Mayer’s Legacy of Research Support: The Prehispanic Ceramic Schools of Greater Nicoya.” In Pre-­ Columbian Art and Archaeology: Essays in Honor of Frederick R. Mayer, edited by Margaret Young-­Sánchez, 27–­46. Denver: Mayer Center for Pre-­Columbian and Spanish Colonial Art, Denver Art Museum. Bishop, Ronald L., Frederick W. Lange, Suzanne Abel-­V idor, and Peter C. Lange. 1992. “Compositional Characterization of the Nicaraguan Ceramic Sample.” In The Archaeology of Pacific Nicaragua, edited by Frederick W. Lange, Payson D. Sheets, Anibal Martínez, and Suzanne Abel-­V idor, 135–­162. Albuquerque: University of New Mexico Press. Bishop, Ronald L., Frederick W. Lange, and Peter C. Lange. 1988. “Ceramic Paste Compositional Patterns in Greater Nicoya Pottery.” In Costa Rican Art and Archaeology: Essays in Honor of Frederick R. Mayer, edited by Frederick W. Lange, 11–­44. Boulder: University of Colorado. Creamer, Winifred. 1983. “Production and Exchange on Two Islands in the Gulf of Nicoya, Costa Rica, A.D. 1200–­1550.” PhD diss., Tulane University, New Orleans. Dennett, Carrie L. 2016. “The Ceramic Economy of Pre-­Columbian Pacific Nicaragua (ad 1–­1250).” PhD diss., University of Calgary. Dickau, Ruth E. 1999. “Paleoethnobotany of the Lake Managua Region, Nicaragua.” MA thesis, University of Calgary. Finlayson, Kerri. 1998. “Stone Tool Industries of Managua, Nicaragua and Their Evolution through Time.” MA thesis, University of Illinois, Urbana-­Champaign. Fonseca Zamora, Oscar M. 1994. “El concepto de Area de Tradición Chibchoide y su pertinencia para entender Gran Nicoya.” Vínculos: Revista de Antropología del Museo Nacional de Costa Rica 18–­19(1–­2):209–­227. González Dávila, Gil. 2002. “Carta del Capitán Gil González Dávila al Rey de España dándole cuenta del descubrimiento de los territorios de Costa Rica y Nicaragua.” In Descubrimiento, conquista y exploración de Nicaragua: Crónicas de fuentes originales seleccionadas y comentadas por Jaime Incer Barquero, edited by Jaime Incer Barquero, 79–­99. Colección Cultural de Centro América, Serie Cronistas 6. Managua: Fundación Vida. Gutiérrez González, Maritza. 1993. “El aprovechamiento de la fauna en el sitio arqueológico Nacascolo, Guanacaste.” Licentiate’s thesis, Universidad de Costa Rica, San José. Gutiérrez González, Maritza. 1998. “La ictiofauna del sitio arqueológico Nacascolo, Bahía Culebra, Guanacaste.” Vínculos: Revista de Antropología del Museo Nacional de Costa Rica 22(1–­2):157–­187. Hardy, Ellen Teresa. 1983. “Burials and Possible Status Distinctions at Nacascolo, Costa Rica.” MA thesis, University of California, Los Angeles. 16

S T EI N B R EN N ER

Hardy, Ellen Teresa. 1992. “The Mortuary Behavior of Guanacaste/Nicoya: An Analysis of Precolumbian Social Structure.” PhD diss., University of California, Los Angeles. Healy, Paul F. 1974. “Archaeological Survey of the Rivas Region, Nicaragua.” PhD diss., Harvard University, Cambridge, MA. Healy, Paul F. 1980. Archaeology of the Rivas Region, Nicaragua. Waterloo, ON: Wilfrid Laurier University Press. Hoopes, John W. 2005. “The Emergence of Social Complexity in the Chibchan World of Southern Central America and Northern Colombia, ad 300–­600.” Journal of Archaeological Research 13(1):1–­47. Hoopes, John W., and Oscar Manuel Fonseca Zamora. 2003. “Goldwork and Chibchan Identity: Endogenous Change and Diffuse Unity in the Isthmo-­Colombian Area.” In Gold and Power in Ancient Costa Rica, Panama, and Colombia: A Symposium at Dumbarton Oaks, 9 and 10 October 1999, edited by Jeffrey Quilter and John W. Hoopes, 49–­89. Washington, DC: Dumbarton Oaks Research Library and Collections. Joyce, Rosemary A. 2016. “Rethinking Ceramics as Evidence of Regional Interaction.” Paper presented at the 81st Annual Meeting of the Society for American Archaeology, Orlando, Florida, April 6–­­10. Lange, Frederick W. 1971. “Culture History of the Sapoá River Valley, Costa Rica.” PhD diss., University of Wisconsin, Madison. Lange, Frederick W. 1978. “Coastal Settlement in Northwestern Costa Rica.” In Prehistoric Coastal Adaptations: The Economy of Maritime Middle America, edited by Barbara L. Stark and Barbara Voorhies, 101–­119. New York: Academic Press. Lange, Frederick W., ed. 1992. Wealth and Hierarchy in the Intermediate Area: A Symposium at Dumbarton Oaks, 10th and 11th October 1987. Washington, DC: Dumbarton Oaks Research Library and Collection. Lange, Frederick W., ed. 1995. Descubriendo las huellas de nuestros antepasados: El proyecto “Arqueología de la Zona Metropolitana de Managua.” Managua: Alcaldía de Managua. Lange, Frederick W., ed. 1996a. Abundante cooperación vecinal: La segunda temporada del proyecto “Arqueología de la Zona Metropolitana de Managua.” Managua: Alcaldía de Managua. Lange, Frederick W., ed. 1996b. Paths to Central American Prehistory. Niwot: University Press of Colorado. Lange, Frederick W., Peter Ryder, and Richard M. Accola. 1986. “Bay of Culebra Survey.” In Prehistoric Settlement Patterns in Costa Rica, edited by Frederick W. Lange and Lynette C. Norr, 25–­36. Journal of the Steward Anthropological Society 14(1–­2):1982–­1983. Urbana: University of Illinois. I ntroduction

17

Lange, Frederick W., Payson D. Sheets, Aníbal Martínez, and Suzanne Abel-­V idor. 1992. The Archaeology of Pacific Nicaragua. Albuquerque: University of New Mexico Press. Lange, Frederick W., and Doris Z. Stone, eds. 1984. The Archaeology of Lower Central America. School of American Research Advanced Seminar Series. Albuquerque: University of New Mexico Press. Laplace, Georges. 1974. “De la dynamique de l’analyse structurale ou la typologie analytique.” Rivista de Scienze Preistoriche 29(1):3–­7 1. McCafferty, Geoffrey G., and Larry Steinbrenner. 2005. “Chronological Implications for Greater Nicoya from the Santa Isabel Project, Nicaragua.” Ancient Mesoamerica 16(1):131–­146. Moreau, Jean-­François. 1980. “A Report on the Hunter-­Robinson and Sardinal Sites.” Vínculos: Revista de Antropología del Museo Nacional de Costa Rica 6(1–­2):107–­124. Moreau, Jean-­François. 1983. “L’adaptation maritime préhistorique au site Vidor, Costa Rica.” PhD diss., Université de Montréal. Moreau, Jean-­François. 1984. “Subsistance [sic] Changes on the Coast of the Greater Nicoya Region.” In Inter-­regional Ties in Costa Rican Prehistory: Papers Presented at a Symposium at Carnegie Museum of Natural History, Pittsburgh, April 27, 1983, edited by Esther R. Skirboll and Winifred Creamer, 129–­141. BAR International Series 226. Oxford: British Archaeological Reports. Niemel, Karen S. 2003. “Social Change and Migration in the Rivas Region, Pacific Nicaragua (1000 bc–­a d 1522).” PhD diss., State University of New York, Buffalo. Norr, Lynette C. 1991. “Nutritional Consequences of Prehistoric Subsistence Strategies in Lower Central America.” PhD diss., University of Illinois, Urbana-­Champaign. Norweb, Albert H. 1961. “The Archaeology of the Greater Nicoya Subarea.” Manuscript on file, Peabody Museum Archives, Peabody Museum of Archaeology and Ethnology, Harvard University, Cambridge, MA. Norweb, Albert Holden. 1964. “Ceramic Stratigraphy in Southwestern Nicaragua.” Proceedings of the 35th International Congress of Americanists 1, 551–­561, Mexico City, August 19–­­25, 1962. Obando, Carmen Patricia. 1995. “Childhood Stress and Bone Maintenance in Pre­historic Northwestern Costa Rica: An Analysis of Two Coastal Populations, Nacascolo and Vidor.” PhD diss., University of Colorado, Boulder. Roddick, Andrew P. 2009. “Communities of Pottery Production and Consumption on the Taraco Peninsula, Bolivia, 200 bc–­300 ad.” PhD diss., University of California, Berkeley. 18

S T EI N B R EN N ER

Román Lacayo, Manuel A. 2013. “Social and Environmental Risk and the Development of Social Complexity in Precolumbian Masaya, Nicaragua.” PhD diss., University of Pittsburgh. Salgado González, Silvia. 1996. “Social Change in a Region of Granada, Pacific Nicaragua (1000 B.C.–­1522 A.D.).” PhD diss., State University of New York, Albany. Solís Del Vecchio, Luis Felipe. 1998. “Nuevos datos en la arqueología de Bahía Culebra, Guanacaste, noroeste de Costa Rica.” Vínculos: Revista de Antropología del Museo Nacional de Costa Rica 22(1–­2):1–­44. Solís Del Vecchio, Luis Felipe. 2002. “El criterio ‘facilidad de obtención’ para explicar las frecuencias relativas de moluscos de los depósitos de conchas precolombinos de la Península de Nacascolo.” Vínculos: Revista de Antropología del Museo Nacional de Costa Rica 27(1–­2):63–­80. Solís Del Vecchio, Luis Felipe, and Anayensy Herrera Villalobos. 2009. “Proyecto Arqueológico Jícaro. Informe de Laboratorio. Temporadas II (2006–­2007) y III (2007–­2008).” Manuscript on file, Museo Nacional de Costa Rica, San José. Squier, Ephraim G. 1852. Nicaragua: Its People, Scenery, Monuments, and the Proposed Interoceanic Canal. 2 vols. New York: D. Appleton and Co. Squier, Ephraim G. 1853. “Observations on the Archaeology and Ethnology of Nicaragua.” Transactions of the American Ethnological Society 3(1):85–­158. Steinbrenner, Larry. 2010. “Potting Traditions and Cultural Continuity in Pacific Nicaragua, ad 800–­1350.” PhD diss., University of Calgary. Sweeney, Jeanne W. 1975. “Guanacaste, Costa Rica: An Analysis of Precolumbian Ceramics from the Northwest Coast.” PhD diss., University of Pennsylvania, Philadelphia. Wallace, Henry, and Richard M. Accola. 1980. “Investigaciones arqueológicas preliminares de Nacascolo, Bahía Culebra, Costa Rica.” Vínculos: Revista de Antropología del Museo Nacional de Costa Rica 6(1–­2):51–­66. Wenger, Étienne C. 1998. Communities of Practice: Learning, Meaning and Identity. Learning in Doing: Social, Cognitive and Computational Perspectives. Cambridge: Cambridge University Press. Wyss, Sue Bursey. 1983. “San Cristobal Archaeological Site, Managua, Nicaragua: Site Report and Preliminary Ceramic Analysis.” MA thesis, Texas A&M University, College Station.

I ntroduction

19

Part 1

Redefining Greater Nicoya

2 The first documented account of Europeans meeting the Indigenous peoples of Greater Nicoya in their own territories dates to 1522, when Gil González Dávila, leading an exploratory expedition from Panama, landed on Costa Rica’s central Pacific Coast and traveled from there overland through Nicoya and Guanacaste into southwestern Nicaragua (Abel-­Vidor 1981; Newson 1987). While this expedition ultimately ended with González retreating back to Panama, within a decade other Spanish adventurers following in his footsteps had seized control of the region, effectively enslaving or otherwise domineering the Indigenous population and founding several towns (including Granada and León) on the sites of some of the most populous preexisting pueblos. The new colony encompassing northwestern Costa Rica and Pacific Nicaragua was initially treated by the Spanish more as booty than as something to be developed for long-­term settlement. In less than a generation the territory had been more or less thoroughly sacked, and modern estimates suggest that the population was reduced (through disease, warfare, and the illegal export of slaves to Peru) by as much as 95 percent, from over a half million at the time of contact to between 30,000 and 40,000 less than twenty-­five years later (Abel-­Vidor 1981:90; Newson 1987:336–­337). Not surprisingly, this genocide was accompanied by extensive social disruption, and most of the dominant precontact groups did not survive as coherent cultural entities into modern times. The loss of native lore, languages, oral

Contact-­Era Pacific Nicaragua Indigenous Groups and Their Origins Larry Steinbrenner

DOI: 10.5876/9781646421510.c002

23

histories, and traditions was tremendous—­even greater than that experienced in areas such as Mexico or Guatemala, where precontact knowledge was well preserved in comparison (Lange et al. 1992:15). While this catastrophic loss of traditions and lore makes it impossible to completely reconstruct the societies of precontact Greater Nicoya, a relatively small number of surviving ethnohistorical records allow us to paint a rough picture of life immediately before the conquest (see Abel-­Vidor 1980, 1981; Fowler 1989; Ibarra Rojas 1995, 2001; Incer Barquero 1993, 2002; Newson 1987; and Werner 2000 for useful bibliographies and summaries of this ethno­ historic database). This chapter draws on these sources and corresponding linguistic and archaeological evidence to provide an overview of the major linguistic groups occupying Greater Nicoya’s northern sector (i.e., Pacific Nicaragua) at the time of contact and their possible origins, in addition to discussing how new research necessarily challenges traditional interpretations of the archaeological evidence for migration. For additional discussion of some of these groups, see chapter 3 in this volume. LINGUISTIC GROUPS

Historical accounts portray contact-­era “Nicaragua” as a confusing mosaic of more than two dozen distinct, competing groups speaking as many as seven different languages (figure 2.1). Unfortunately, it is often difficult to differentiate between the various groups described in these accounts. While groups are usually defined by their languages and/or geographical location, major early chroniclers such as Gonzalo Fernández de Oviedo y Valdés (1478–­1557) and Antonio de Ciudad Real (1551–­1617)—­who both visited Nicaragua in the sixteenth century, albeit decades apart—­rarely use the same terms to describe linguistic groups, are not always consistent in their own usage of terms, and often discuss cultural practices generically without specifying the groups to which they belong (Lothrop 1926:45).1 Despite such problems, however, the various historical accounts are in general agreement that at the time of European contact there were three large linguistic groups in Greater Nicoya claiming Mesoamerican origin (cf. Constenla Umaña 1991, 1994; Ibarra Rojas 2001; Lothrop 1926; Werner 2000). These groups almost certainly would have 1. Oviedo, for example, uses the terms “Nicaragua” and “Chorotega” to name the languages modern scholars would call Nahuat and Chorotega, while Ciudad Real refers to the same languages as “corrupt Mexican” and “Mangue”; Oviedo also refers to Chorotega and Orosí as distinct languages in one instance but then applies these terms to two groups possessing the same language in another (Abel-­V idor 1981:88).

24

S T EI N B R EN N ER

Figure 2.1. Linguistic groups of Greater Nicoya (cf. Lothrop 1926:24 and Constenla Umaña 1994, redrawn as figure 3.2, this volume). Map by Larry Steinbrenner.

subsumed rival groups that would not necessarily have self-­identified as members of a common cultural entity based on language alone. They included the following: 1. the Chorotega, or Mangue, speakers of Chorotega, a now-­extinct Otomanguean language related to Chiapanec, an extinct language of highland Chiapas (Lewis 2009). Chorotega speakers represented the largest Contact- ­E ra Pacific N icaragua I ndigenous G roups and T heir O rigins

25

population claiming Mesoamerican origin and collectively comprised groups located mainly in central Pacific Nicaragua (the Mangue proper) and northwestern Costa Rica (the Orotiña, the Nicoya, and the Orosí). Traditionally, it has been assumed that they also inhabited the Isthmus of Rivas prior to the later arrival of the Nicarao. Despite the tendency for scholars to treat “the Chorotega” as a monolithic entity with a common origin (see following discussion), relations between the rather broadly dispersed Chorotega-­speaking groups remain mostly undefined and unknown. 2. the Nicarao, Nahuas who dominated the Isthmus of Rivas south of the Río Ochomogo. Nahua speakers (not the most numerous but by far the best documented group in Nicaragua) were also settled in modern Chinandega in Nicaragua’s northwest (the Nahuatlatos), on the eastern shore of Lake Nicaragua, and in central Guanacaste (the Bagaces). The term “Nicarao” has come to be applied to all of these groups, which are not usually differentiated in historical accounts. However, there is no clear evidence that these various groups self-­identified as a single cultural group, and some of the groups may have been unrelated to the Rivas Nicarao. The Rivas group spoke a (now extinct) dialect called Nahuat, subtly different from the Nahuatl of Central Mexico. 3. the Maribios, a poorly known group mostly confined to the modern department of León (northwest of Lake Managua) who spoke a now extinct Otomanguean language related to Chiapanec/Chorotega but more closely related still to Tlapanec (still spoken in Guerrero, Mexico). The Maribios (who actually may have outnumbered the more dispersed Nahuas; see Werner 2000:18–­19) are sometimes referred to in the literature as the “Subtiaba” based on the fact that some Maribios speakers lived in the Indigenous town of that name upon which modern León was founded. Although modern linguists (e.g., Constenla Umaña 1994:204) seem to prefer Subtiaba as the name for the language, Maribios is probably the more historically correct usage for the group (Lothrop 1926:11; Werner 2000:ix), and both terms tend to be used interchangeably for both group and its language. A smaller group of Maribios, the Maribichicoa, lived beyond Greater Nicoya around Ocotal in the department of New Segovia, apparently having relocated there shortly before the Conquest (Constenla Umaña 1994:205).

Other languages of possible Mesoamerican origin spoken in Nicaragua (e.g., “Huave” and “Poton”) are named in historical sources, but very little information about these is available (cf. Constenla Umaña 1994:194 and Werner 26

S T EI N B R EN N ER

2000:13–­17).2 Patrick Werner (2000:11–­13) suggests that “Poton” may have referred to the language of the Putun Maya, a hypothesis that is consistent with his argument that the “Chondales” described by Oviedo and other contemporary chroniclers lived in northeastern Chinandega and may have been “Mayoid.” This idea is somewhat out of step with the current school of thought (cf. Constenla Umaña 1994; Ibarra Rojas 1994; Van Broekhoven 2002), which holds that the Chondales/Chontales—­possibly several different groups conflated into a single “other” population—­probably spoke Matagalpa, an extinct language usually associated with Nicaragua’s Central Highlands, farther to the east and beyond the bounds of Greater Nicoya (Constenla Umaña 1994:195). However, Werner (2000:14–­15) argues convincingly that the term “Chontal”—­ the word derives from the Nahuat chontalli, meaning “foreigner” and was also commonly applied by other Nahua groups in different parts of Mexico to various unrelated groups (Ibarra Rojas 2001:21)—­was not used to describe Indigenous people east of Lake Nicaragua until the nineteenth century and was used in colonial times to refer only to groups in western and northwestern Nicaragua living in very close proximity to Nahuas, Chorotega, and Maribios on a trade route to Honduras. In short, Werner proposes that scholars have confounded two separate groups of “foreigners”: a colonial-­era group of possibly Mesoamerican origin living in the west/northwest on the fringes of Greater Nicoya and a different group of Matagalpa speakers living at greater remove in the Central Highlands. However, the presence of two conflated groups does not rule out the possibility that the colonial-­era Chondales may have been Matagalpa speakers rather than Mayas, since, when all is said, almost nothing is known about Poton other than that it was spoken in the area where the Chondales lived. Matagalpa belonged to the Misumalpan language family, which is closely related to the very ancient Chibchan language family that once dominated Central America from Costa Rica to Colombia and Venezuela (see below). The two language families and several others—­including the extinct Lenca family, formerly spoken by populations neighboring Greater Nicoya in El Salvador and Honduras (Constenla Umaña 1991:17), and also proposed by Werner (2000:17) as an alternate candidate for the language of the colonial-­era Chondales—­are often grouped into a larger Macro-­Chibchan or Chibchan-­Paezan family (Campbell 1997:167). Other Misumalpan languages include most of the languages indigenous to Nicaragua’s Atlantic 2. Constenla Umaña (1994:194) suggests that Huave may have been the language of the Tacacho, a group that lived in a town called Yacacoyaua located in Maribio territory. Contact- ­E ra Pacific N icaragua I ndigenous G roups and T heir O rigins

27

Watershed, including Ulua and Sumo. Languages belonging to the Chibchan family proper were also spoken: in Greater Nicoya’s southern sector, speakers of the now-­extinct Corobici language lived along Lake Nicaragua’s southeastern shore, on the lake’s Solentiname Islands, along the San Juan River, and across the Cordillera de Tilarán as far as the Gulf of Nicoya (Constenla Umaña 1994:198–­200; Lothrop 1926:17).3 Chibchan-­and Chibchan-­affiliated groups such as the Matagalpa and Corobici are even more poorly documented in the contact-­period historical sources than the Greater Nicoyan groups claiming Mesoamerican origin, and almost everything that is known about these groups at contact is based on historical accounts from later times (see Van Broekhoven 2002) or on passing references to them found in accounts of better-­documented linguistic groups such as the Nicarao and Chorotega. ORIGINS AND MIGRATION MYTHS

According to their own histories, the various Mesoamerican-­related groups in Greater Nicoya were relatively recent arrivals in an area with an archaeological record of earlier occupation that dates back to at least 2000 bc (Haberland 1986; Niemel 2003; Salgado González 1996). Given that linguistic and genetic evidence has now been used to identify a wide-­ranging Chibchan tradition extending across much of Lower Central America and northern South America, with roots potentially reaching as far back as the Holocene (Fonseca Zamora 1994; Hoopes 2005; Ibarra Rojas 1990; Salgado González 1996:49–­53), it is not unlikely that Greater Nicoya’s earliest archaeological record reflects an original occupation by Chibchans ancestral to—­or at very least, related to—­the Matagalpa and/or Corobici who lived on Greater Nicoya’s periphery at the time of contact.4 This supposition finds support in generic similarities between Greater Nicoya’s early material culture (especially ceramics) and the 3. As Constenla Umaña notes, the Corobici of the contact period appear to have been the ancestors of either Nicaragua’s modern Rama (who now live along the San Juan and near the Atlantic Coast) or Costa Rica’s modern Guatusos (who live along the San Juan’s southern tributaries).

4. Constenla Umaña (1994:197) suggests that the Matagalpa were the original inhabitants of Pacific Nicaragua and rejects Incer’s (1993:291) alternative suggestion that the autochthonous residents may have been the Miskitos, Misumalpans whose own oral traditions—­suspect, in Constenla’s view—­tell of migrating to the Atlantic Coast after being expelled from Pacific Nicaragua in the ninth century. Another possibility is that the autochthons of Pacific Nicaragua were related to the poorly known Lenca. However, Dennett and Platz (2017) have recently argued that these “autochthons” were not connected to any of these groups (see note 11, this chapter, for additional discussion).

28

S T EI N B R EN N ER

contemporary material culture of surrounding regions (Bonilla Vargas et al. 1990; Hoopes 1987, 1994). Unfortunately, however, the surviving histories do not provide a Matagalpan or Corobici perspective on this question. In contrast, ethnohistorical accounts recorded by the Spanish provide a great deal more information about the arrival of Nicarao and the Chorotega, though they are silent concerning the Maribios. The various accounts—­the most important of which are found in Oviedo (citing Francisco de Bobadilla), Motolinía (a.k.a. Toribio de Benavente), and Juan de Torquemada—­are often confusing and occasionally contradictory, providing dates that range (based on comparisons by later scholars with linguistic and archaeological evidence; see following discussion) from about the fifth century ad to as few as eighty years before the arrival of the Spanish (Fowler 1989:32). Of these accounts, Bobadilla’s (cited in Fernández de  Oviedo y Valdés 1976:311, 327–­328) remains the only one for which a clear source is provided (Salgado González 1996:23). Bobadilla’s report was based on interviews conducted shortly after the conquest with Nicarao nobles, priests, and huehues (elders) in Teoca, one of several plazas located near modern San Jorge in Rivas ruled by the eponymous teyte (chief ) Nicaragua (Werner 2000:6). Bobadilla’s informants related that their ancestors had come from towns called Ticomega and Maguatega in the direction where the sun sets, fleeing masters who had treated them poorly, and that they could no longer recall when this migration had occurred, since it had been before their own lifetimes. The locations of these towns—­to say nothing of the country in which they might have been found—­remain unknown, though Walter Lehmann’s (1920:2:1006) speculative and “linguistically weak” (Fowler 1989:32) identification of these origin places with Central Mexican towns in Morelos (Tecoman) and Puebla (Miahuatlán) is unfortunately often treated as an established “fact.” Motolinía’s account is contemporary with Bobadilla’s but more suspect, since he does not provide sources and since he is writing from a Mexican perspective (he was a missionary in the New World but may have never actually visited Nicaragua) that is questionable in other respects (e.g., a tradition that all of the major peoples of Mexico are descended from a common ancestor is presented as fact). As a result, his account, while potentially incorporating true events, appears to be at least partially speculative—­perhaps representing Motolinía’s private theory of how Nahua speakers ended up in Nicaragua. This version of the story tells of a migration from an unknown part of New Spain spurred by a famous four-­year drought: “A great number of canoes and boats went by way of the South Sea and landed in Nicaragua . . . They made war upon the natives who were settled there and defeated them and overthrew their power. The invading Nahuas Contact- ­E ra Pacific N icaragua I ndigenous G roups and T heir O rigins

29

remained and settled there, and . . . that was only a hundred years ago, more or less” (Motolinía [1540] 1950:31). Most scholars (e.g., Healy 1980; Lothrop 1926) are inclined to see Motolinía’s account as pertaining to the Nicarao, given that he explicitly offers his story as an explanation for how Nahua speakers ended up in Nicaragua and also comments on their numerous population. If true, this account would therefore place the arrival of the Nicarao in the fifteenth century. Recent climatological studies (e.g., Cleaveland et al. 2003; Metcalfe and Davies 2007) have established the fact of a fifteenth-­century megadrought in Mexico that could easily have spurred the migration recorded by Motolinía. William Fowler (1989:33), who favors an earlier arrival for the Nicarao, suggests this account might relate to the “Subtiaba” (i.e., Maribios) rather than the Nicarao, but the Maribios were not Nahuas. However, it should not be taken for granted that all Nahuas in Greater Nicoya necessarily arrived as part of a common migration event. An alternative possibility that does not appear to have been previously explored is that the seagoing Nahuas of Motolinía’s account represent one of the groups living outside the Rivas area, such as the Nahuatlatos of Chinandega. Torquemada’s work, based on both native and Spanish sources (probably including lost works of Motolinía; cf. Werner 2000:2), was published in 1615, almost a century after the discovery of Nicaragua by the Spanish. Fowler (1989:34) feels that Torquemada provides the “most informative historical account” of the migrations, a sentiment shared by many other scholars (e.g., Salgado González 1996:25), though others (e.g., Werner 2000:2) find Torquemada to be speculative. His work (Torquemada [1615] 1975–­1983: 1:452–­454; partially translated in Fowler 1989:34–­36) is exceptional in providing origins for both the Nicarao and the Chorotega. Torquemada relates that the “Nicaraguas” (i.e., Nicarao) and the “Nicoyas” or “Mangues” (i.e., the Chorotega) both anciently lived in the Soconusco (Xoconochco) region of the modern Mexican state of Chiapas. The Nicaraguas, described as being originally from “Anahuac” (i.e., the Valley of Mexico), lived toward the Pacific Coast, while the Nicoyas, described as descendants of “Chololtecas” (i.e., people from the powerful Central Mexican center of Cholula), lived inland toward the mountains. Invaded and defeated by “Olmecas” from the northwest (identified by most scholars as the Olmeca-­Xicalanca) and resentful of paying tribute, both groups secretly pulled up stakes and traveled en masse south along the Pacific Coast through Guatemala and eventually into El Salvador. Members of this exodus who settled in areas along the way became the Pipil, a Salvadoran Nahua group that Fowler (1989) associates intimately with the Nicarao, going so far as to treat the “Pipil-­Nicarao” as a single cultural entity. 30

S T EI N B R EN N ER

While still in Guatemala, a dying leader prophesied that the Nicoyas would lead the way farther south into Central America but would ultimately come under the domination of bearded white masters because of their evil ways. For the Nicaraguas, the leader prophesied that they would end up in a land already occupied by Nicoyas opposite an island with two mountains (a succinct description of Rivas and the volcanic island of Ometepe), where they would betray the Nicoyas, though the Nicaraguas too would ultimately also come to serve bearded men. This betrayal comes to pass later in Torquemada’s account, which indicates that after long wanderings in Central America that took the Nicaraguas across to the Caribbean and as far south as Panama, they arrived in Rivas and defeated the previously established Nicoyas through treachery. Refugees from the resulting battle are said to have fled south into the area subsequently known as Nicoya (i.e., Costa Rica), though Torquemada also establishes that Nicoyas had previously also settled in areas farther to the north near Xolotlán (Lake Managua) before the arrival of the Nicaraguas. Torquemada’s account has little in common with Motolinía’s (which would seem to argue against the idea that it draws on a lost work by that earlier author) but is consistent with Bobadilla’s inasmuch as both indicate that the “Nicarao” arrived in Nicaragua fleeing oppression. Various scholars (e.g., Abel-­ Vidor 1980:161–­162; Chapman 1974:94; Fowler 1989:36; Nicholson 1973:491) have tended to see this story as a sort of “official history” of the Nicarao that justifies their betrayal of the Chorotega by portraying it as foretold and therefore inevitable. Suzanne Abel-­Vidor (1980:161), more critical than most, also suggests that this account “reveals details that strongly suggest that it is a politicized version incorporating post-­conquest events as well as Prehispanic legend.” She cites Nicholson, who observed that many events in Torquemada are “consciously or unconsciously” fitted “into stylized patterns heavily influenced by religious and cosmological preconceptions” (Nicholson 1973:491, cited in Abel-­Vidor 1980:162). Indeed, it is difficult to avoid comparing the tale of a long-­suffering tribe of Nahuas wandering through the wilderness of Central America in search of a promised land—­a promised land that their leader will never live to enter—­with the biblical story of Exodus. The similarity between these stories cautions against taking Torquemada’s attractively detailed account at face value. In particular, Torquemada’s attempt to provide a common origin for all Nahua groups in Central America (Nicarao, Pipils, etc.) now seems problematic, and his own account includes details that raise flags for the careful reader. For example, Torquemada specifically notes that a population of Nahuas living in a town near the Desaguadero (the mouth of the San Juan River), “hablan en lengua mexicana no tan corruta como estotra de los pipiles” Contact- ­E ra Pacific N icaragua I ndigenous G roups and T heir O rigins

31

(i.e., “spoke a less corrupt form of Mexican than the Pipils”). (Torquemada [1615] 1975–­1983:1:454). Given that this town is described as being settled by migrant Nahuas after the foundation of the Pipil towns in Guatemala and El Salvador, it seems incongruous that its inhabitants speak better Nahuatl than the Pipil! A joint migration involving all Central American Nahuas is not favored by most modern scholars with the exception of Fowler (1981, 1989) and is also disputed by modern linguistic studies, as discussed below. These factors provide good reasons to be suspicious about most of the affiliations between the Rivas Nicarao and other Central American Nahua groups established in Torquemada’s account. On the other hand, there is no particular reason to question the detail that Chorotega speakers were established in Rivas before the Nahuas, or that the latter defeated the former and established control of territory in the Rivas area, since these details agree with other historical accounts. There is also no reason to summarily dismiss Torquemada’s claims that the ancestors of the Nicaraguas and the Nicoyas/Mangues came from Anahuac and Cholula before settling in the Soconusco, though it is worth noting that these ultimate Central Mexican origins do not form part of the narratives of either Bobadilla or Motolinía—­a fact that is sometimes obscured when scholars uncritically assume Lehmann was correct in identifying Bobadilla’s towns with Central Mexico or that Motolinía’s migrating Nahuas must have come from this region. With respect to the more immediate Soconuscan origin, it should be reiterated that the Chorotega language was related to the Chiapanec language once spoken in the highlands of this area and also noted that the Spanish were already aware of a connection between the Chiapanec people and Nicaragua in early colonial times. Carlos Navarrete (1966:5), summarizing the ethnohistory of the Chiapanecs, cites a 1571 alegato (statement) which claimed that (according to the Chiapanecs’ Zinacantec rivals), “the Chiapanec were a people intrusive in the Central Depression [of Chiapas], having come from Nicaragua and having appropriated lands which did not belong to them.” Navarrete adds that in the early seventeenth century, Antonio de Herrera y Tordesillas also reported that the Chiapanec “came anciently from the province of Nicaragua” and that in the early eighteenth century Francisco Ximénez (1929) reported much the same. However, Navarrete himself rejects these claims based on stronger evidence that the Chiapanec-­Chorotega connection ran north to south, and he notes that the Chiapanec responded to the Zinacantec “by saying that they were natives (eran naturales) of the province of Chiapa [sic] since time immemorial, and that, even if they were not, they had been settled in the place for more than a thousand years and had furthermore given their name to the 32

S T EI N B R EN N ER

province” (Navarrete 1966:5). An alternative interpretation that remains to be explored is that both versions of the story are partially true: it is possible that Chorotega speakers left Chiapas for the reasons described in Torquemada but that a group returned years later (after their defeat by the Nicarao in Rivas?) to either recolonize their ancestral homeland or perhaps to seek refuge with distant Chiapanec kin who had remained behind. DATING THE MIGRATIONS: ETHNOHISTORIC AND LINGUISTIC PERSPECTIVES

Torquemada does not provide dates for the arrival of the Chorotega or Nicarao, giving only a frustratingly vague estimate of when the two groups lived together in Chiapas—­that is, “seven or eight ages, or lifetimes of old men, ago” (Torquemada [1615] 1975–­1983:1:452)—­without specifying how long it took them to get to Greater Nicoya.5 The other two chroniclers are equally vague, though in different ways. Bobadilla’s account can be linked to a specific Nahua group (the Rivas Nicarao) but only tells us that they arrived long enough ago for their descendants to lose track, which would seem to imply at least several generations rather than a comparatively recent arrival. Motolinía’s account, in contrast, seemingly provides a more specific date (i.e., 100 years before contact) but cannot be securely linked to a specific Nahua group. As a result, previous attempts to date the various migrations have therefore depended largely on (a) tying these events to major historical events elsewhere in Mesoamerica, (b) glottochronology-­based estimates for the divergence of dialects such as Chorotega and Nahuat from their Mexican counterparts, and (c) archaeological evidence of cultural change. With respect to the historical approaches (summarized and reassessed in Fowler 1989:36–­49), various scholars (e.g., Borhegyi 1965; Jiménez Moreno 1959) have generally hypothesized that the migrations were spurred by Central Mexican phenomena with far-­reaching effects, such as the fall of Teotihuacan, the expansion and collapse of the Toltec empire, and other later events occurring during Mesoamerica’s turbulent Postclassic period (see also Chapman 1974:75–­76; Healy 1980:22; Hoopes and McCafferty 1989; Lothrop 1926:8). However, as Fowler (1989:38) has observed, the hypotheses of scholars such as Jiménez Moreno and Borhegyi tend to the speculative and often rely on “flimsy evidence.” Since there is little data (archaeological or otherwise) to 5. The translation here is from Fowler (1989:34), which also provides a discussion of these dates. Contact- ­E ra Pacific N icaragua I ndigenous G roups and T heir O rigins

33

support the “high degree of chronological precision pretended by Jiménez Moreno and Borhegyi,” Fowler believes that their hypotheses are of dubious value in dating the migrations. Furthermore, most of these hypotheses explain the complex phenomena of migration (Anthony 1990, 1992, 1997, 2007; Burmeister 2000; Chapman and Hamerow 1997) using a single monocausal formula (i.e., political expansion/collapse = diaspora). It is questionable whether this approach, with its default assumption that the migrations invariably represented mechanical responses to major Central Mexican events, provides much new insight into Greater Nicoyan prehistory, since it fails to recognize the potential importance of numerous alternative push-­and-­pull factors that might have brought Mesoamericans into Lower Central America. A partial list of these factors (discussed in Niemel 2003) might include comparatively minor regional conflicts, factional disputes, environmental factors (like Motolinía’s drought), previous knowledge of the destination area or trade routes facilitating movement, and/or a desire to expand trading networks or exploit new resources.6 At any rate, since direct connections between the expansion and/or downfall of Central Mexican political entities and the various migrations that brought Mesoamericans into Nicaragua cannot be taken for granted, especially considering the previously discussed lack of ethnohistoric evidence supporting the hypothesis that the migrants arrived directly from Central Mexico, we cannot presume to use major Central Mexican events to confidently date these migrations—­even if the dating of these events themselves were not subject to ongoing debates (cf. Evans 2013). Linguistic studies have not only provided potential dates for the Chorotega, Nahua, and Maribios migrations but have also offered clues as to the origins of these groups that are probably of even greater value.7 With respect 6. Various scholars (e.g., Ibarra Rojas 1995; Niemel 2003; Salgado González 1996:36) have suggested that populations of Mesoamerican origin originally settled in strategically important locations that allowed them to access or control trade routes linking Greater Nicoya to interaction spheres on the Southeastern Mesoamerican Periphery and facilitating the region’s integration into the periphery of the Mesoamerican world-­system (Carmack and Salgado González 2006). The original territories of the Chorotega (including the Masaya, Granada, and Rivas areas) would have served this purpose admirably (Carmack and Salgado González 2006), and the lands of the northern Nahuas and Maribios also appear to have been adjacent to ancient trading routes (Werner 2000). In Rivas, the Nicarao more specifically might have been able to control trade between the northern and southern sectors of Greater Nicoya, since only a single road passed through the isthmus (Abel-­Vidor 1986:392). 7. The linguistic evidence relevant to Greater Nicoya is summarized in Constenla Umaña (1994).

34

S T EI N B R EN N ER

to the Chorotega, there appears to be a consensus that Chiapanec-­Mangue originated in the Valley of Puebla (Hopkins 1984:52; Kaufman 1990:97; Lastra 2001:128). The dominant center in this valley for much of Mesoamerican prehistory was Cholula, and scholars, following Torquemada’s lead, have long assumed that “Chorotega” points to a Cholulan origin (Kaufman 1990:97; Lothrop 1926:20–­21). However, Samuel Lothrop believed it was a mistake to link “Chorotega” to Cholula and cites Daniel Brinton (1883) as observing that both terms could derive from “Aztec” root words referring to the act of running away—­that is, to words that the Nicarao might have used to describe defeated enemies. Lothrop also observed that “Chorotega” was the name of the first cacique encountered by the Spanish in Costa Rica and that the Spanish typically named groups after their leaders (Lothrop 1926:21). Still, it is not inconceivable that Chorotega’s name might have derived from the name of either his group or his group’s homeland. The common assumption (e.g., Hopkins 1984:52) is that a group from Cholula traveled from the area of Puebla to Chiapas toward the end of the Mesoamerican Classic period (a hypothesis obviously consistent with Torquemada), where the group then split into the Chiapanecs (who remained) and the Chorotega (who continued on to Nicaragua). Late twentieth-­century linguistic studies (e.g., Campbell 1988; Constenla Umaña 1994:200; Hopkins 1984:43; Kaufman 1974) generally date this split to around ad 600–­700 based on Morris Swadesh’s (1967:97) original glottochronological estimate that the two languages show a minimum of thirteen centuries of divergence, and it is therefore often taken for granted that the Chorotega arrived in Greater Nicoya shortly after this time. There is less consensus about the origins of Central American Nahua groups in the linguistic literature. Earlier scholars (e.g., Campbell 1988:280; Fowler 1989; Jiménez Moreno 1966:65; Kaufmann 1974:49) proposed a homeland in Veracruz predating the appearance of the Pipil-­Nicarao in the Soconusco and a split between the Nicarao and the Pipil dating to around ad 800–­1000. However, more recent work generally treats Pipil as an “Eastern Nahuatl” language with a history separate from the Nicarao’s Nahuat, a “Western Nahuatl” dialect now considered to be more closely related to the Nahuatl of Central Mexico (Constenla Umaña 1994:201–­204; Dakin 2003:261). This work therefore does not appear to support the hypothesis that the Pipil and Nicarao arrived in Central America together; in fact, Constenla suggests that the Nicarao arrived in Central America circa ad 1200, centuries later than the Pipil. This work also makes it more feasible that the Nicarao did indeed arrive in the Soconusco from the Valley of Mexico, as recorded in Torquemada, even while disputing Torquemada’s linking of this Nahua group Contact- ­E ra Pacific N icaragua I ndigenous G roups and T heir O rigins

35

with Pipil. Nahua speakers began to become prominent in Central Mexico during the late Classic period (Evans 2013:367–­368) and eventually occupied vast areas that included the valleys of Puebla and Tehuacán earlier occupied by Otomangueans (Lastra 2001:128). With respect to the Maribios, linguistic studies (e.g., Constenla Umaña 1994:204) usually link the group to the Guerrero homeland of the Tlapanecs. Suárez (1983; cited in Niemel 2003:21) suggests that the Tlapanecs anciently migrated to Guerrero from Malinche, Puebla. Invoking Swadesh (1967:97), who argued that Tlapanec and Maribios show eight minimum centuries of differentiation, Constenla suggests that the Maribios separated from the Tlapanecs and came to Nicaragua sometime after ad 1200. While glottochronological estimates for the divergence of related languages are often treated as almost sacrosanct by archaeologists and ethnohistorians (perhaps because they seem to provide a “high degree of chronological precision” that is attractive to scholars working with a frustratingly imprecise database), these dates should not be accepted uncritically. Most of the dates summarized above are based on a half-­century–­old glottochronological model that assumed a constant rate of linguistic change—­a position rejected even by modern linguists who otherwise continue to endorse this controversial approach (e.g., Kaufman and Justeson 2009:229–­30).8 Given this weakness, it would be a mistake to assume that any of these dates are inherently superior to other lines of evidence, such as ethnohistorical records or archaeological evidence for migration that might point to very different dates for the arrival of Nicaragua’s Mesoamerican groups. DATING THE MIGRATIONS: ARCHAEOLOGICAL EVIDENCE

The estimated dates provided by linguistic studies for the appearance in Greater Nicoya of Chorotega (ad 600–­700) and Nahuat (ad 1200) are approximately similar to the dates traditionally assigned to the beginnings of the Middle Polychrome / Sapoá period (ca. ad 750–­800) and the Late Polychrome / Ometepe period (originally thought to begin ca. ad 1000, but now more commonly assumed to begin ca. ad 1300–­1350; see chapter 4, this volume). Given that these periods are marked by significant changes 8. While Kaufman and Justeson have called for the application of improved glottochronological methods in future work, to my knowledge this has not yet been done for most of the languages of interest here. It goes without saying that an updated glottochronology for Central American languages would be valuable to archaeologists working in Greater Nicoya.

36

S T EI N B R EN N ER

in the archaeological record—­such as the appearance of new ceramic types and shifts in settlement patterns—­it has been generally assumed that the changes marking the onset of the Middle Polychrome / Sapoá period (especially the first appearance of Papagayo Polychrome) reflect the arrival of the Chorotega, while those changes traditionally believed to signal the onset of the Late Polychrome / Ometepe period (especially the appearance of Vallejo Polychrome) mark the arrival of the Nicarao (Baudez 1967; Day 1984; Healy 1980; Lange 1971; Niemel 2003; Salgado González 1996; Sweeney 1975). However, as discussed in chapters 9 and 10 of this volume, most of the polychrome ceramic types previously believed to be diagnostic of the Ometepe period are in fact centuries older and date to the latter part of the preceding Sapoá period, which makes it unlikely that they were associated with the Nicarao unless this group arrived in Nicaragua far earlier than any of the previously summarized evidence indicates. If the “Ometepe period” diagnostic types are, in fact, “late Sapoá” diagnostics and cannot be associated with the Nicarao, then whom, if anyone, could they be associated with? My own interpretation, based on the analysis of Nicaraguan potting traditions summarized in chapter 9 of this volume, involves a nontraditional reading of the archaeological evidence for change. Since I see the early Sapoá-­period ceramic types that have been traditionally associated with Chorotega speakers as being products of a widespread Indigenous Central American tradition clearly connecting Nicaragua and Costa Rica to El Salvador and Honduras (see Steinbrenner 2016), rather than as definitive evidence of a new wave of Mesoamerican influence without precedent in Lower Central America (the de facto traditional assumption), I am hesitant to unquestioningly associate these ceramic types with the Chorotega. To do so would seem to imply invoking a Salvadoran/Honduran Chorotegan presence beyond the boundaries of Greater Nicoya in areas with no ethnohistoric record of the Chorotega (e.g., in Honduras’ Comayagua Valley).9 On the other hand, since I, like virtually all Greater Nicoya scholars, do see a very strong Mesoamerican stylistic influence in the late Sapoá 9. Admittedly, the absence of ethnohistoric evidence for Chorotega speakers in El Salvador and Honduras (notwithstanding the Choluteca of southern Honduras) is not conclusive evidence of their absence in these countries. As Stone (who actually did believe that the Chorotega were a presence in Central Honduras) observed, “there is a small amount of documentation concerning the [Central American] extension of the Chorotega-­Mangue [among other groups]. Outside of this, however, however, the territory of the peoples associated with the interior of Honduras . . . is practically nondefined in historical sources” (Stone 1957:45). Contact- ­E ra Pacific N icaragua I ndigenous G roups and T heir O rigins

37

polychromes—­Vallejo Polychrome in particular is generally acknowledged to be the most “Mexican” of Greater Nicoya’s ceramic types (cf. Day 1984; Bonilla Vargas et al. 1990; Lothrop 1926; Healy 1980)—­I would argue that the archaeological evidence of these ceramics potentially does reflect the arrival of a group with Mesoamerican origins.10 By a process of elimination, this group would seem to be either a later-­arriving population of Chorotega (an interpretation that would conflict with the rather-­suspect date given by glottochronology for the Chorotega-­Chiapanec split) or, alternatively (and perhaps more likely, especially if the Chorotega did arrive at the beginning of the Sapoá period as is commonly believed, if not earlier11), the Maribios, whose origins and potential impact on Nicaragua’s archaeological record have been almost completely ignored.12 Additional, and probably less likely, options might include a non-­Nicarao Nahua group, like the Nahuatlatos (whose origins, as previously noted, were not necessarily the same as the Rivas Nahuas), or Werner’s “Mayoid” (or Lencan) Chondales of Chinandega. What all of these groups have in common is a presence in northwestern Pacific Nicaragua, the region in which Vallejo Polychrome was most likely produced (Dennett 2016:217; Steinbrenner 2010:271; see also chapter 10, this volume). As for the Nicarao themselves, the abandonment circa ad 1250–­1300 of Santa Isabel, the largest occupation site located in the contact-­era heartland of the Nicarao (McCafferty and Steinbrenner 2003; Steinbrenner 2010; see also chapter 6, this volume), seems to infer a dramatic reorganization of the Rivas area that could represent the arrival of a new power in the region at that 10. As discussed in chapter 9 of this volume, I do not see the late Sapoá ceramics as being “introduced” by a new Mesoamerican group but rather as reflecting the impact of the arrival of a new group upon an already establishing potting tradition. There is, in fact, considerable continuity between the ceramics of the early and late Sapoá periods.

11. It is possible that the arrival of Chorotega speakers into Greater Nicoya may have been even earlier than previously believed. In an intriguing recent paper, Dennett and Platz (2017) have argued that the ancestors of the contact-­era Chorotega arrived (possibly from Oaxaca) into a Pacific Nicaragua depopulated by recent volcanic activity in the Tempisque period, a thousand years before the beginning of the Sapoá period! This hypothesis effectively positions Chorotega speakers as being as “autochthonous” to Greater Nicoya as any Chibchan or Chibchan-­affiliated groups and would suggest a likely connection between the Chorotega and early Sapoá ceramic types—­and by extension, between the Chorotega and the “indigenous Central American potting tradition” discussed immediately above.

12. Note that a late Sapoá arrival of the Maribios would be roughly consistent with Constenla Umaña’s (1994:204) suggestion that the group arrived relatively late in Nicaragua.

38

S T EI N B R EN N ER

time. This being the case, the estimated date of arrival for the Nicarao (two or three centuries before contact) remains much the same as that suggested by Constenla and would seem to provide sufficient time for contact-­period Nicarao to have forgotten exactly when their ancestors arrived—­assuming that Bobadilla’s informants were not simply dissembling in order to establish a stronger claim on their territory. QUESTIONS FOR FUTURE RESEARCH

While this new reading (or readings) of the archaeological evidence for change suggests several possible arrival dates for linguistic groups originating in Mesoamerica, inevitably it also raises many new questions that cannot be answered without more data. If, for example, Chorotega speakers were not associated with the new ceramic types and settlement pattern changes that mark the beginning of the Sapoá period, then how are these changes to be explained? Is it possible, in fact, to rule out a Chorotega presence in central Honduras and El Salvador that might explain the apparent connections between certain ceramic types in these countries and Greater Nicoya? Conversely, if Chorotega speakers actually were associated with early Sapoá types while the Maribios were associated with late Sapoá types, then how could this be demonstrated and what sort of interaction between these groups could have led to the development of the latter types out of their earlier antecedents? If the arrival of the Nicarao is not marked by the appearance of late Sapoá ceramic types as previously believed, then what does it look like in the archaeological record, and why have true Nicarao occupation sites not been previously identified?13 13. With respect to the question of the “missing” Ometepe-­period Nicarao archaeological sites in Rivas, I have argued elsewhere (Steinbrenner 2010:68, 529) that the Nicarao, rather than being associated with Vallejo and other former “Ometepe period” ceramics, may have been alternatively linked to the Cuapa Complex (ad 1400–­1600), an intrusive complex that has to date been identified only in Chontales, where it was originally attributed—­mistakenly, I believe—­to a hitherto undocumented “Matagalpa” group (Gorin 1990:33, 221–­222). The Cuapa Complex, which features high frequencies of comals—­a signature vessel form that we would most definitely expect to find in an archaeological assemblage associated with late-­arriving immigrants of Mexican origin—­comprises unremarkable, poor-­quality monochromes and bichromes that might easily be overlooked by archaeologists working in Pacific Nicaragua with archaeological assemblages in which Nicoya Polychromes are also present, which was not the case in the Chontales site where the complex was first identified. Contact- ­E ra Pacific N icaragua I ndigenous G roups and T heir O rigins

39

Underlying all of these questions is an even larger theoretical concern: is it ultimately even possible to link the diverse Indigenous groups—­linguistic or otherwise—­identified in Greater Nicoya’s patchwork ethnohistoric record to material culture remains such as ceramics? While frontiers between languages often are defined by distinctive material culture assemblages, they do not have to be (see Anthony 2007:103ff ). Given how very little we know about some of the groups discussed here, such as the Maribios and Chondales—­to say nothing of the various groups routinely lumped into the presumably “better known” Chorotega and Nicarao groups—­we should face up to the possibility that the answer to our larger question is not necessarily an unqualified “yes.” Alternatively, we should consider that it might ultimately be more productive to attempt to identify other types of social entities in the archaeological record, such as communities or constellations of practice (Bourdieu 1977; Roddick 2009; Sassaman and Rudolphi 2001; Wenger 1998), without presupposing (as Greater Nicoyan scholars have been wont to do) that the identification of specific cultural groups is the be-­all and end-­all of archaeological inquiry. Beyond these new questions, many of our original questions relating to the origins of Nicaragua’s Mesoamerican immigrants also remain unanswered, the most fundamental of which are: Where did these groups really come from, and how did they integrate into Greater Nicoya’s preexisting cultural fabric upon their arrival?14 While the evidence from oral traditions and linguistics, as summarized here, seems frustratingly incomplete and imprecise, it remains possible that questions about origins might still be answerable for at least some of the groups. However, to date no archaeologically oriented study has specifically and systematically attempted to identify linkages between the material culture of Pacific Nicaragua and that of any of the regions that have been tentatively identified as the origin points of any of Greater Nicoya’s Mesoamerican populations, such as the Soconusco. Studies focusing on bioarchaeological evidence have been similarly lacking, though work on ancient DNA is now beginning (see chapter 14, this volume). Studies such as these would seem to be essential to the investigation of Greater Nicoya’s Mesoamerican connections and are therefore urgently needed. The time has come for scholars studying these connections to start looking for answers outside the ethnohistoric record and beyond the northern borders of Nicaragua. 14. For a speculative discussion of how the Chorotega in particular might have integrated with Indigenous Chibchan populations upon their arrival in Greater Nicoya, see Steinbrenner (2010:123–­141). For a discussion of the archaeological evidence for Chorotega “ethnogenesis,” see McCafferty and Dennett (2013).

40

S T EI N B R EN N ER

REFERENCES

Abel-­V idor, Suzanne. 1980. “The Historical Sources for the Greater Nicoya Archaeological Subarea.” Vínculos: Revista de Antropología del Museo Nacional de Costa Rica 6(1–­2):155–­176. Abel-­V idor, Suzanne. 1981. “Ethnohistorical Approaches to the Archaeology of Greater Nicoya.” In Between Continents / Between Seas: Precolumbian Art of Costa Rica, edited by Elizabeth P. Benson, 85–­92. New York: Harry N. Abrams. Abel-­V idor, Suzanne. 1986. “Early Sixteenth Century Evidence for the Settlement Archaeology of Greater Nicoya.” In Prehistoric Settlement Patterns in Costa Rica, edited by Frederick W. Lange and Lynette C. Norr, 387–­405. Journal of the Steward Anthropological Society 14 (1–­2 [1982–­1983]). Urbana: University of Illinois. Anthony, David W. 1990. “Migration in Archaeology: The Baby and the Bathwater.” American Anthropologist 92(4):895–­914. Anthony, David W. 1992. “The Bath Refilled: Migration in Archaeology Again.” American Anthropologist 94(1):174–­176. Anthony, David W. 1997. “Prehistoric Migration as Social Process: Material and Social Constraints.” In Migrations and Invasions in Archaeological Explanation: Long-­Term Perspectives, edited by John Chapman and Helena Hamerow, 21–­32. BAR International Series 664. Oxford: British Archaeological Reports. Anthony, David W. 2007. The Horse, the Wheel, and Language. Princeton, NJ: Prince­ ton University Press. Baudez, Claude F. 1967. Recherches archéologiques dans la vallée du Tempisque, Guanacaste, Costa Rica. Travaux et mémoires 18. Paris: Institut des Hautes Études de l’Amérique Latine, Université de Paris. Bonilla Vargas, Leidy, Marlin Calvo Mora, Juan Vicente Guerrero Miranda, Silvia Salgado González, and Frederick W. Lange. 1990. “La cerámica de la Gran Nicoya.” Vínculos: Revista de Antropología del Museo Nacional de Costa Rica 13(1–­2 [1987]):1–­327. Borhegyi, Stephan F. de. 1965. “Archaeological Synthesis of the Guatemalan Highlands.” In Archaeology of Southern Mesoamerica, edited by Gordon R. Willey, 3–­58. Handbook of Middle American Indians 2, pt. 1, Robert Wauchope, general editor. Austin: University of Texas Press. Bourdieu, Pierre. 1977. Outline of a Theory of Practice. Cambridge Studies in Social Anthropology 16. Cambridge: Cambridge University Press. Brinton, Daniel G., ed. 1883. The Güegüence: A Comedy Ballet in the Nahuatl-­Spanish Dialect of Nicaragua. Translated by Daniel G. Brinton. Brinton’s Library of Aboriginal American Literature 3. Philadelphia: Daniel G. Brinton.

Contact- ­E ra Pacific N icaragua I ndigenous G roups and T heir O rigins

41

Burmeister, Stefan. 2000. “Archaeology and Migration: Approaches to an Archaeological Proof of Migration.” Current Anthropology 41(4):539–­567. Campbell, Lyle. 1988. The Linguistics of Southeast Chiapas, Mexico. Papers of the New World Archaeological Foundation 50. Provo, UT: New World Archaeological Foundation, Brigham Young University. Campbell, Lyle. 1997. American Indian Languages: The Historical Linguistics of Native America. Oxford Studies in Anthropological Linguistics. Oxford: Oxford University Press. Chapman, Anne M. 1974. Los nicarao y los chorotega según las fuentes históricas. 2nd ed. San José: Editorial de la Universidad de Costa Rica. Chapman, John, and Helena Hamerow. 1997. “On the Move Again: Migrations and Invasions in Archaeological Explanation.” In Migrations and Invasions in Archaeological Explanation: Long-­Term Perspectives, edited by John Chapman and Helena Hamerow, 1–­11. BAR International Series 664. Oxford: British Archaeological Reports. Cleaveland, Malcolm K., David W. Stahle, Matthew D. Therrell, José Villanueva-­ Diaz, and Barney T. Burn. 2003. “Tree-­Ring Reconstructed Winter Precipitation and Tropical Teleconnections in Durango, Mexico.” Climatic Change 59(3):369–­388. Constenla Umaña, Adolfo. 1991. Las lenguas del área intermedia: Introducción a su estudio areal. San José: Editorial de la Universidad de Costa Rica. Constenla Umaña, Adolfo. 1994. “Las lenguas de la Gran Nicoya.” Vínculos: Revista de Antropología del Museo Nacional de Costa Rica 18–­19 (1–­2 [1992–­1993]):191–­208. Dakin, Karen. 2003. “Uto-­Aztecan in the Linguistic Stratigraphy of Mesoamerican Prehistory.” In Language Contacts in Prehistory: Studies in Stratigraphy, edited by Henning Andersen, 259–­288. Amsterdam Studies in the Theory and History of Linguistic Science, Current Issues in Linguistic Theory 239. Amsterdam: John Benjamins Publishing. Day, Jane S. 1984. “New Approaches in Stylistic Analysis: The Late Polychrome Period Ceramics from Hacienda Tempisque, Guanacaste Province, Costa Rica.” PhD diss., University of Colorado, Boulder. Dennett, Carrie L. 2016. “The Ceramic Economy of Pre-­Columbian Pacific Nicaragua (ad 1–­1250).” PhD diss., University of Calgary. Dennett, Carrie L., and Lorelei A. Platz. 2017. “Chorotega Ancestry and Religious Cult in Greater Nicoya.” Paper presented at the XI Congreso de la Red Centroamericana de Antropología, University of Costa Rica, San José, February 27–­­March 2. Evans, Susan Toby. 2013. Ancient Mexico and Central America: Archaeology and Culture History. 3rd ed. London: Thames and Hudson. 42

S T EI N B R EN N ER

Fernández de Oviedo y Valdés, Gonzalo. 1976. Nicaragua en los cronistas de Indias: Oviedo. Colección Cultural de Centroamérica. Serie Cronistas 2. Managua: Fondo de Promoción Cultural, Banco de América. Fonseca Zamora, Oscar M. 1994. “El concepto de Area de Tradición Chibchoide y su pertinencia para entender Gran Nicoya.” Vínculos: Revista de Antropología del Museo Nacional de Costa Rica 18–­19(1–­2):209–­227. Fowler, William R., Jr. 1981. “The Pipil-­Nicarao of Central America.” PhD diss., University of Calgary. Fowler, William R., Jr. 1989. The Cultural Evolution of Ancient Nahua Civilizations: The Pipil-­Nicarao of Central America. Civilization of the American Indian Series 194. Norman: University of Oklahoma Press. Haberland, Wolfgang. 1986. “Settlement Patterns and Cultural History of Ometepe Island, Nicaragua: A Preliminary Sketch.” In Prehistoric Settlement Patterns in Costa Rica, edited by Frederick W. Lange and Lynette C. Norr, 369–­386. Journal of the Steward Anthropological Society 14 (1–­2 [1982–­1983]). Urbana: University of Illinois. Healy, Paul F. 1980. Archaeology of the Rivas Region, Nicaragua. Waterloo, ON: Wilfrid Laurier University Press. Hoopes, John W. 1987. “Early Ceramics and the Origins of Village Life in Lower Central America.” PhD diss., Harvard University, Cambridge, MA. Hoopes, John W. 1994. “The Tronadora Complex: Early Formative Ceramics in Northwestern Costa Rica.” Latin American Antiquity 5(1):3–­30. Hoopes, John W. 2005. “The Emergence of Social Complexity in the Chibchan World of Southern Central America and Northern Colombia, ad 300–­600.” Journal of Archaeological Research 13(1):1–­47. Hoopes, John W., and Geoffrey G. McCafferty. 1989. “Out of Mexico: An Archaeological Evaluation of the Migration Legends of Greater Nicoya.” Paper presented at the 54th Annual Meeting of the Society for American Archaeology, Atlanta, Georgia, April 5–­­9. Hopkins, Nicholas A. 1984. “Otomanguean Linguistic Prehistory.” In Essays in Otomanguean Culture History, edited by J. Kathryn Josserand, Marcus C. Winter, and Nicholas A. Hopkins, 25–­64. Vanderbilt University Publications in Anthropology 31. Nashville: Vanderbilt University Press. Ibarra Rojas, Eugenia. 1990. Las sociedades cacicales de Costa Rica (Siglo XVI). Colección Historia de Costa Rica 3. San José: Editorial de la Universidad de Costa Rica. Ibarra Rojas, Eugenia. 1994. “Los matagalpas a principios del siglo XVI: Una aproximación a las relaciones interétnicas en Nicaragua (1522–­1581).” Vínculos: Revista de Antropología del Museo Nacional de Costa Rica 18–­19(1–­2):229–­244. Contact- ­E ra Pacific N icaragua I ndigenous G roups and T heir O rigins

43

Ibarra Rojas, Eugenia. 1995. “Historia de Nicaragua y Nicoya durante la conquista española: Una perspectiva desde la dinámica interétnica (800 d.C.–­1544).” MA thesis, Universidad de Costa Rica. Ibarra Rojas, Eugenia. 2001. Fronteras étnicas en la conquista de Nicaragua y Nicoya: Entre la solidaridad y el conflicto, 800 d.C.–­1544. San José: Editorial de la Universidad de Costa Rica. Incer Barquero, Jaime. 1993. Nicaragua, viajes, rutas y encuentros, 1502–­1838: Historia de las exploraciones y descubrimientos, antes de ser Estado independiente, con observaciones sobre su geografía, etnia y naturaleza. 2nd ed., Colección Quinto Centenario, Serie Raíces. San José: Libro Libre. Incer Barquero, Jaime, ed. 2002. Descubrimiento, conquista y exploración de Nicaragua: Crónicas de fuentes originales seleccionadas y comentadas por Jaime Incer Barquero, Colección Cultural de Centro América, Serie Cronistas 6. Managua: Fundación Vida. Jiménez Moreno, Wigberto. 1959. “Síntesis de la historia pretolteca de Mesoamérica.” In Esplendor del México antiguo. Vol. 2, edited by Carmen Cook de Leonard, 1019–­1108. Mexico City: Centro de Investigaciones Antropológicas de México, Editorial del Valle de México. Jiménez Moreno, Wigberto. 1966. “Mesoamerica before the Toltecs.” In Ancient Oaxaca: Discoveries in Mexican Archaeology and History, edited by John Paddock, 3–­82. Stanford, CA: Stanford University Press. Kaufman, Terrance. 1974. Idiomas de Mesoamérica. Seminario de Integración Social Guatemalteca, Publicación 33. Guatemala: Editorial José de Pineda Ibarra, Ministerio de Educación. Kaufman, Terrance. 1990. “Early Oto-­Manguean Homelands and Cultures: Some Premature Hypotheses.” University of Pittsburgh Working Papers in Linguistics 1(1):91–­136. Kaufman, Terrance, and John Justeson. 2009. “Historical Linguistics and Pre-­ Columbian Mesoamerica.” Ancient Mesoamerica 20(2):221–­231. Lange, Frederick W. 1971. “Culture History of the Sapoá River Valley, Costa Rica.” PhD diss., University of Wisconsin, Madison. Lange, Frederick W., Payson D. Sheets, Aníbal Martínez, and Suzanne Abel-­V idor. 1992. The Archaeology of Pacific Nicaragua. Albuquerque: University of New Mexico Press. Lastra, Yolanda. 2001. “Linguistics.” In The Oxford Encyclopedia of Mesoamerican Cultures, edited by Davíd Carrasco, 123–­131. Vol. 2, The Civilizations of Mexico and Central America. Oxford: Oxford University Press. Lehmann, Walter. 1920. Zentral-­Amerika. Teil I. Die Sprachen Zentral-­Amerikas in ihren Beziehungen zueinander sowie zu Süd-­Amerika und Mexiko. 2 vols. Berlin: Verlag Dietrich Reimer. 44

S T EI N B R EN N ER

Lewis, M. Paul, ed. 2009. Ethnologue: Languages of the World. 16th ed. Dallas: SIL International. Lothrop, Samuel K. 1926. Pottery of Costa Rica and Nicaragua. 2 vols. Memoir 8. New York: Museum of the American Indian, Heye Foundation. McCafferty, Geoffrey G., and Carrie L. Dennett. 2013. “Ethnogenesis and Hybridity in Proto-­historic Nicaragua.” Archaeological Review from Cambridge 28(1):191–­215. McCafferty, Geoffrey G., and Larry Steinbrenner. 2003. “Informe N-RI-­44-­00: Santa Isabel, Rivas.” Manuscript on file, Instituto Nicaragüense de Cultura (INC), Managua. Metcalfe, Sarah, and Sarah Davies. 2007. “Deciphering Recent Climate Change in Central Mexican Lake Records.” Climatic Change 83(1–­2):169–­186. Motolinía, Fray Toribio (Toribio de Benavente). (1540) 1950. Motolinía’s History of the Indians of New Spain. Translated and edited by Elizabeth Andros Foster. Documents and Narratives Concerning the Discovery and Conquest of Latin America 4. Berkeley: Cortes Society. Navarrete, Carlos. 1966. The Chiapanec History and Culture. Edited by John Alden Mason. Papers of the New World Archaeological Foundation 21, Publication 16. Provo, UT: Brigham Young University Press. Newson, Linda A. 1987. Indian Survival in Colonial Nicaragua. Civilization of the American Indian Series 175. Norman: University of Oklahoma Press. Nicholson, Henry B. 1973. “Middle American Ethnohistory: An Overview.” In Guide to Ethnohistoric Sources (Part Four), edited by Howard F. Cline, Charles Gibson, and Henry B. Nicholson, 487–­505. Handbook of Middle American Indians 15, Robert Wauchope, general editor. Austin: University of Texas Press. Niemel, Karen S. 2003. “Social Change and Migration in the Rivas Region, Pacific Nicaragua (1000 bc–­a d 1522).” PhD diss., State University of New York, Buffalo. Roddick, Andrew P. 2009. “Communities of Pottery Production and Consumption on the Taraco Peninsula, Bolivia, 200 bc–­300 ad.” PhD diss., University of California, Berkeley. Salgado González, Silvia. 1996. “Social Change in a Region of Granada, Pacific Nicaragua (1000 B.C.–­1522 A.D.).” PhD diss., State University of New York, Albany. Sassaman, Kenneth E., and Wictoria Rudolphi. 2001. “Communities of Practice in the Early Ceramic Traditions of the American Southeast.” Journal of Anthropological Research 57(4):407–­425. Steinbrenner, Larry. 2010. “Potting Traditions and Cultural Continuity in Pacific Nicaragua, ad 800–­1350.” PhD diss., University of Calgary.

Contact- ­E ra Pacific N icaragua I ndigenous G roups and T heir O rigins

45

Steinbrenner, Larry. 2016. “ ‘Nicoya Polychromes’ beyond Greater Nicoya: Las Vegas Polychromes in El Salvador and Honduras.” Paper presented at the 81st Annual Meeting of the Society for American Archaeology, Orlando, Florida, April 6–­­10. Stone, Doris Z. 1957. The Archaeology of Central and Southern Honduras. Papers of the Peabody Museum of Archaeology and Ethnology, vol. 49, no. 3. Cambridge, MA: Peabody Museum of Archaeology and Ethnology, Harvard University. Suárez, Jorge A. 1983. La lengua tlapaneca de Malinaltepec. Mexico City: Universidad Nacional Autónoma de México, Instituto de Investigaciones Filológicas. Swadesh, Morris. 1967. “Lexicostatistic Classification.” In Linguistics, edited by Norman A. McQuown, 79–­115. Handbook of Middle American Indians 5, Robert Wauchope, general editor. Austin: University of Texas Press. Sweeney, Jeanne W. 1975. “Guanacaste, Costa Rica: An Analysis of Precolumbian Ceramics from the Northwest Coast.” PhD diss., University of Pennsylvania. Torquemada, Juan de. (1615) 1975–­1983. Monarquía Indiana. De los veinte y un libros rituales y monarquía indiana, con el origen y guerras de los indios occidentales, de sus poblazones, descubrimiento, conquista, conversión y otras cosas maravillosas de la mesma tierra. 3rd ed. 7 vols., Historiadores y Cronistas de Indias 5. Mexico City: Instituto de Investigaciones Históricas, Universidad Nacional Autónoma de México. Van Broekhoven, Laura N. K. 2002. Conquistando lo invencible: Fuentes históricas sobre las culturas indígenas de la región central de Nicaragua. Mededelingen van het Rijksmuseum voor Volkenkunde 31. Leiden, Netherlands: CNWS Publications, Leiden University. Ximénez, Fray Francisco. 1929. Historia de San Vincente de Chiapa y Guatemala. Biblioteca “Goathemala” de la Sociedad de Geografía e Historia, Vol. 1. Guatemala. Wenger, Étienne C. 1998. Communities of Practice: Learning, Meaning and Identity. Learning in Doing: Social, Cognitive and Computational Perspectives. Cambridge: Cambridge University Press. Werner, Patrick S. 2000. Ethnohistory of Early Colonial Nicaragua: Demography and Encomiendas of the Indian Communities. Institute for Mesoamerican Studies Occasional Publication 4. Albany: State University of New York.

46

S T EI N B R EN N ER

3 A researcher faces a challenge when he or she seeks a deeper understanding of the daily interactions—­both between the different ethnies friendly and hostile—­ (ethnic groups) (Ibarra Rojas 2001) that lived in the region of Pacific Nicaragua and Nicoya in the sixteenth century. Increasingly, it is apparent that the area that has been known for more than a half century as “Greater Nicoya” can no longer be studied in the same way as it has been for all those years. From a historical perspective, it is now apparent that the concept of a “cultural area” that is rigidly delimited in both time and space is too limited to facilitate a detailed examination of dynamic human relationships. I believe that it is now both appropriate and convenient to abandon the concept of the Greater Nicoya cultural area along with its conceptual implications and to focus instead on a broader area that can be more simply designated as the Pacific Nicaragua and Nicoya region and its surroundings. This is an extensive territory wherein the history of Chibchan-­ affiliated groups (e.g., the Corobici and Chontals) as well as Mesoamerican-­ linked groups (e.g., the Chorotega, Nicarao, and Maribios) can be better reconstructed and understood, for the sixteenth century as well as for some time backward and forward. It must be considered that the Spanish conquest and colonization together were not a monolithic process but rather a multifaceted process that unfolded at different times and in different ways throughout the area. To reconstruct and explain

Indigenous Peoples of Pacific Nicaragua and Nicoya in the Sixteenth Century A Historical Approach Eugenia Ibarra Rojas

DOI: 10.5876/9781646421510.c003

47

the Indigenous substratum that received the impact of the Spanish presence in the area and how this substratum was changed by this presence and the differentiated processes of colonization, our examination of the process must be broken down to focus on actors, time and space. As is well known, the historical sources for addressing this subject are scarce. Only a couple of “new” documents have been included in this chapter, but their content is not so significant as to dramatically change most of the “known” history that has been documented previously. The published material found in the compilations of León Fernández Bonilla, Andrés Vega Bolaños, Manuel María Peralta, Gonzalo Fernández de  Oviedo y Valdés, and Juan de  Torquemada, to name just a few classic authors, remains as valid today as it was several decades ago. The significant information found in such historical texts is useful for demonstrating the existence and continuity of some Indigenous practices and the interruption of others, due to changes generated by different newcomers arriving at different times. One must also consider the important bibliography bequeathed to the social sciences by ethnohistorians, colonial historians, art historians, archaeologists, linguists, anthropologists, biologists, and geographers, and these are just the main disciplines that have been involved in specific studies in the area. Getting closer to the people of the Pacific Nicaragua and Nicoya region and its surroundings necessarily begins with an interdisciplinary approach. As a historian, I believe that the continuation of some Indigenous cultural practices into early colonial times facilitates the identification of precontact cultural customs that were probably practiced on the verge of the Spanish presence. The reconstruction of sociopolitical structures based on the sixteenth-­century materials can be used to understand how these structures may have merged with later colonial institutions, such as the cofradías, for example, which continue to thrive today and which appear to interweave Indigenous sociopolitical and religious practices with Christian religious and political practices. It is the questions that we ask of documentary sources and the theoretical-­ methodological paths that we follow to find the answers that open possibilities to different ways of reinterpreting “old” information. Our approach can make visible some of the cultural variables that had been made invisible by employing the concept of the Greater Nicoya cultural area. This theoretical-­ methodological approximation begins by considering Pacific Nicaragua, Nicoya, and the surrounding region as a “pluricultural confluence space.” The available sources must be read once again and contrasted continually with new evidence. They must be researched with a powerful magnifying glass and 48

I BA R R A RO JA S

probed with the finest tweezers possible. I will explore these ideas in more depth in the following pages. THE PACIFIC REGION OF NICARAGUA AND NICOYA AND ITS SURROUNDINGS AS A PLURICULTURAL CONFLUENCE SPACE

Twenty years ago, Silvia Salgado González (1996) commented that the concept of Greater Nicoya had already been in use in archaeology, ethnohistory, and other disciplines, for almost thirty-­five years. As she observed, the customary use of the concept has introduced it into the minds of researchers and members of the general public alike. However, this status quo must not preclude the possibility of changing the way the area and its studies are perceived in the present. While archaeology has generally been successful in classifying the artifacts that are characteristically found in Greater Nicoya, that scientific discipline still falls short of explaining the social dynamics that kept the different ethnies occupying the territory, one next to the other, through changes in time. Instead, Greater Nicoya has been typically described as an area inserted between other cultural areas—­that is, Mesoamerica and the Intermediate Area—­or has been treated as a buffer zone (Lange 1979, 1994). Recently, Frederick Lange (2015) has proposed the methodological use of the wedge concept (i.e., the Southern Mesoamerican Wedge, or SMW) as a possible tool for dismissing frontiers and moving further away from migrations and conquests toward an understanding of the Mesoamerican presence in Greater Nicoya. Still, this concept relies heavily on the classification of objects and cultural characteristics rather than the reconstruction of social dynamics that is of interest to current social sciences. Greater Nicoya has also been conceptualized as a product of core-­periphery relationships and dynamics (Berdan and Smith 2004; Carmack and Salgado González 2006). This perspective has generated some important ideas and has helped scholars to better understand some processes, including economic and political issues. As well, since Greater Nicoya has been studied as being framed within lines or borders, questions arise regarding the definition of the area. Is the same Greater Nicoya delimitation valid all the way from 10,000  bc to the sixteenth century? Salgado González and Ricardo Vázquez Leiva (2006) and Salgado González and Elisa Fernández León (2011) have demonstrated otherwise, showing that after ad 900 the delimitation of the area undoubtedly changed. Other authors, myself included (Ibarra Rojas 2011a, 2012; Ibarra Rojas and Salgado González 2010; Sheets and Mueller 1984; Sheets 2008; I ndigenous P eoples of Pacific N icaragua and N icoya in the S ixteenth C entury

49

Sheets and Sever 1991), have been able to discover and trace changes made by people who lived and exchanged goods with Mesoamerican-­originating peoples who occupied nearby territories. Such research results provide strong arguments for reconsidering the concept of Greater Nicoya. It is essential that research on Greater Nicoya consider the regional context, as this will be of enormous aid in defining local contexts. Put simply, this means that researchers must look for interactions beyond the bounds of the delimited area, including warfare. Contexts undoubtedly play important roles in interaction spheres, and comparisons are also needed. Anthropologists are often surprised with the many unique ways human beings behave under diverse circumstances, and the similarities of their responses as well as their differences! A pluricultural confluence space is a concept that incorporates not only a given area occupied by members of different cultural groups but also the chronological and spatial reconstruction of the history of that area. Implicit in this concept is the need to study the development and the dynamics of cultural practices and social relations between those members, at different moments and in different places. Archaeological, linguistic, colonial-­ historical, and ethno­historical perspectives collectively confirm that the Pacific Nicaragua and Nicoya region and its surroundings (figure 3.1) can be classified as a pluricultural confluence space from about ad 800 to 1550 (see McCafferty, Salgado González, and Dennett 2009). The various results provided by each of these disciplines in response to specific research questions represent pieces of a puzzle that, when assembled, can aid in the reconstruction of human activities in this region across time and space. This reconstruction will be dynamic, not static, and should see people and objects as interacting; humans, for example, can be perceived as creating and using tools and resources, participating in cultural practices such as religious ceremonies or political elections, managing production and exchange, or engaging in warfare. In the case of Pacific Nicaragua and Nicoya, the territories where different cultures coincided were not originally empty, since members of the Misumalpan and Chibchan language families (collectively described as the “Macro-­Chibchan” family) lived there before the arrival of migrants with Mesoamerican origins, particularly the Chorotega (who arrived close to ad 800) and the Nicarao and Maribios (who made their way closer to ad 1200).1 The area must therefore be explored with these first inhabitants in mind, since pluricultural confluence spaces do not begin in empty territories (figure 3.2). 1. Editor’s note: the challenges inherent in dating the migrations of various Mesoamerican-­ originating groups to Greater Nicoya is also discussed in chapter 2, this volume.

50

I BA R R A RO JA S

Figure 3.1. Indigenous occupation of the Pacific region of Nicaragua and Nicoya and its surroundings in the sixteenth century. Map by Paula M. Pérez.

The actors in this pluricultural confluence area are therefore the Corobici (or Rama), the Matagalpa (or Chontals), the Chorotega (or Mangue), the Nicarao (or Nicaragua), and the Maribios (or Subtiaba). Corobici and Matagalpa are Macro-­Chibchan languages, Chorotega and Subtiaba belong I ndigenous P eoples of Pacific N icaragua and N icoya in the S ixteenth C entury

51

Figure 3.2. Reconstructed distribution of Greater Nicoya languages. Map redrawn by Larry Steinbrenner after Constenla Umaña (1994).

to the Otomanguean language family, and the Nahuat spoken by the Nicarao was an Uto-­Aztecan dialect closely related to the Nahuatl of Central Mexico (Constenla Umaña 1994). 52

I BA R R A RO JA S

This historical context must necessarily be considered in order to wholly understand the processes governing the interactions between these different groups. Archaeology is capable of detecting some of the activities of the older inhabitants and can even perhaps synthesize some of their reactions to the migrants’ presence. For example, Payson Sheets and Thomas Sever (1991) detected that people from the Arenal area made a cultural turn toward other Indigenous societies in the Caribbean region of Costa Rica, distancing themselves from Mesoamerican-­ originating people. Such findings can provide potential answers to social-­oriented questions, suggesting that Chibchans had their own attitudes and made their own decisions. Results of this sort can also help to clarify that territories were already occupied and delimited by the time Mesoamericans arrived, immersed in their own problems. The migrants not only generated changes in the precise areas where they arrived, but also caused changes to societies further away. History has also constructed hypotheses about how the presence of Mesoamerican migrants could have caused changes to the Indigenous societies of northeastern Honduras, Caribbean Nicaragua, and the central valley and northern plains of Costa Rica (Ibarra Rojas 2011b). Their arrival in the Pacific area seems to have acted like a stone thrown at the middle of a pond, triggering waves that rippled away from the center to cause movements and changes on shores farther away. The area of interest is full of bodies of water such as the ocean, gulfs, lakes, and rivers. It must be stated that the islands in the Gulf of Fonseca were also pluricultural, as were the Solentiname Islands in Lake Nicaragua. For archaeological and historical research, the main questions should focus on the identification and explanation of the processes that made it possible for so many different cultures to live in the same territory. Another main subject, close to the first one mentioned, is related to processes linked to identity and ethnicity, which go hand in hand with sociopolitical hierarchization and power. The cultures in Nicaragua and Nicoya and surrounding areas were still diverse in the sixteenth century, implying that the various ethnies must have developed social, political, and economic interactions that kept them consciously differentiated from each other. This suggests that a Frederick Barth–­inspired examination of how political variables pertained to the problem of ethnicity in this region could therefore be helpful here (Ibarra Rojas 2001). Ethnicity is closely related to the languages spoken. It should be noted that the Corobici of the Nicoya Peninsula, immersed within the Chorotegan groups that dominated the peninsula, still spoke their own language in the sixteenth century. In northern Pacific Nicaragua, the remaining Chontals or Matagalpa also still spoke their own language, some eight centuries after the arrival of the first Mesoamerican-­originating groups. I ndigenous P eoples of Pacific N icaragua and N icoya in the S ixteenth C entury

53

The persistence of these languages representing the oldest inhabitants among the various other languages spoken in the area was witnessed and described by Oviedo and other Spanish officers. What processes made those languages so enduring? One conclusion seems clear: the Matagalpa/Chontals and the Corobici/Rama survived as minorities on the eve of the Spanish conquest, facts that in themselves suggest a connection to political processes associated with the domination, through time, of the Chorotega, possibly including cooperation or even alliances. When we undertake a study of the Spanish arrival and colonial practices in Pacific Nicaragua and Nicoya in the sixteenth century, the resulting historical perspective leads us to understand that not all processes happened at the same time everywhere (see Ibarra Rojas 2014). Neither were all the changes imposed upon Indigenous societies made in the same way, nor was the impact upon these societies always identical. We also see that the diverse ethnies were also organized differently. These findings are explored in the following section. THE INDIGENOUS PEOPLE AS REVEALED IN SPANISH ACCOUNTS

The major Spanish conquest and colonization visits to Nicaragua and Nicoya occurred in 1519 ( Juan de  Castañeda and Hernán Ponce de  León), 1523 (Francisco Hernández de Córdoba), 1526 (Pedrarias Dávila [Pedro Arias de Dávila]), 1528 (Gonzalo Fernández de Oviedo y Valdés), and 1534 (Bartolomé de Las Casas). These visits are important because the documentary accounts that they generated (discussed below) offer essential information concerning different sociopolitical aspects of the Indigenous peoples who are mentioned. Bartolomé de  Las Casas (1986) comments on how, in 1519, when Juan de Castañeda and Hernán Ponce de León first visited the Gulf of Nicoya, the inhabitants received them with their weapons ready in a very hostile and menacing manner—­definitely not a passive response but a very active one indeed. A first question arises, then: who were these inhabitants? As it was multicultural territory, did the Chorotega and Corobici/Rama unite in defense of their lives, women, and territories? Was it only the Corobici/Rama under the command of Chorotega warrior leaders who went to the beach to defend themselves, or members of both groups? While the question is challenging to answer, it could have been members of both ethnies in an alliance. I shall return to this point after following the Spanish footsteps into the area. Our examination of the discovery and conquest of the area discussed here must begin in Panama, for that is where the process was initiated. 54

I BA R R A RO JA S

It must be kept in mind that in 1519 Juan de  Castañeda captured three or four Indigenous persons from the Nicoya Peninsula and took them to Panama to use as guides and interpreters (Solórzano Fonseca and Quirós Solórzano 2006). Three years later, on January 21, 1522, Gil González Dávila and Andrés de Cereceda headed west from Panama. Soon after their departure, González Dávila claimed he was having problems with his vessel. The plan was that Cereceda should keep navigating west while González Dávila would leave the ship and travel overland to the Gulf of Nicoya, where they would meet. As soon as he was on land, González Dávila started exchanging with the local inhabitants, looking for gold. The conquerors reunited as planned in the Gulf of Nicoya, then visited Gurutiña and other places in the peninsula, including Paro, Cangen, and Nicoya. Cacique Nicoya, the supreme ruler in the region, invited them to stay for ten days, after which they headed north, visiting Zapandí, Corobici, Diriá, Namiapí, Orosí, and Papagayo. From Papagayo they moved to a small bay close to present-­day San Juan del Sur in Nicaragua. A close examination of the route followed, settlement by settlement, suggests these Spaniards knew from the beginning where they were going, how to get there, and where the gold and the major populated areas were located, information they perhaps acquired from captured locals in 1519. In April 1523, González Dávila and his men arrived at Quauhcapolca, close to present-­day San Jorge in Nicaragua, where Cacique Nicaragua lived. From there González Dávila moved to Nochari, on a northeast route. In Diriá he met Cacique Diriangén, whom he tried to convert to Christianity. Diriangén promised he would be back with an answer in three days. He came back, but to fight González Dávila and his people, who had to return to Nicoya, from whence they sailed back to Panama in June of 1523 (Solórzano and Quirós 2006). The accounts of this expedition and its happenings present us with an image of the native inhabitants as possessing organized sociopolitical structures and caciques who were ready to wage war upon menacing strangers. Settlement patterns also emerge from the aforementioned documentary sources. Indigenous people chose diverse areas in which to build their towns. It is clear that coastal areas, inland zones close to navigable rivers and/or the gulf coast, and islands were prominent choices as preferred areas. As well, the foothills of volcanoes stand out as important and convenient places upon which to live. One can also discern the existence of pre-­Hispanic routes that joined the places mentioned. Nowhere does either González Dávila or Cereceda mention any difficulties whatsoever involving moving from one place to another. The I ndigenous P eoples of Pacific N icaragua and N icoya in the S ixteenth C entury

55

paths had to be wide enough to support horses, a problem that was partially mitigated by the dryness and the flatness the terrain presented. Spaniards in this region did not claim the heavy tasks of “opening” paths, as they usually did when they spoke of the difficulties of traveling in Talamanca or in other areas in southern Costa Rica. The evidence summarized above points toward organized complex societies with chiefs/caciques or hierarchized cacicazgos (see Salgado González 1996:137–­149) who were immersed in trade and exchange with people both near and far away using routes that included lakes, rivers, land routes, and oceans. Isla del Caño in the Pacific Ocean off Costa Rica provides archaeological evidence of contact with Nicoya (Finch and Honetschlager 1986), and in 1514 Pascual de Andagoya mentioned how people “from Nicaragua” navigated to Panama to attack cacique Paris for his gold (Álvarez Rubiano 1944). Trade and exchange and warfare with other central Costa Rican cacicazgos were also frequent (Ibarra Rojas 1988, 1990, 2012). Lower Central America, from the Pacific region of Nicaragua and Nicoya and extending south into other areas of Costa Rica and Panama, was a busy commercial area at the time of the Spanish arrival. The trade networks were surely constructed centuries earlier, and some modifications likely happened to them over time, due to changing socioeconomic interests, or perhaps as a result of warfare. In the sixteenth century, however, they were complicated routes that had to be controlled by strong, cunning, political decisions that most likely involved many cacicazgos—­a variable that must be considered in any study of the Pacific region of Nicaragua and Nicoya. For example, in the sixteenth century, the cacao trade not only was important to caciques in Pacific Costa Rica and the Central Valley but also involved the caciques who occupied the lower lands close to the Desaguadero or San Juan River basin—­that is, the Huetares and Ramas (or Votos), who may even have participated in the cacao network in earlier times (Ibarra Rojas 2011a; Steinbrenner 2006). In 1523, Hernández de Córdoba arrived at the Gulf of Nicoya and went north, following the same route González Dávila had used previously. Hernández de Córdoba knew about the la mar dulce (the sweet sea, a.k.a. Cocibolca, modern Lake Nicaragua) and for that reason, his people (200 men plus African slaves) carried an unassembled vessel to sail on the lake. There is not much information about this conquistador’s doings, but there was violence and fighting. In a letter to the king, Castañeda spoke negatively of the Chorotega, stating some of them fled while others let themselves die, a statement that is difficult to understand. He also added that they were “bad and vicious,” for they consumed human flesh (Peralta 1883:49). 56

I BA R R A RO JA S

Facing strong resistance and the constant threat of war, Hernández de Córdoba was convinced of the need to fortify himself against Indigenous groups as well as his rival González Dávila. He therefore founded the original city of León in the Indigenous territory of Nagrando close to the northern end of Lake Managua (or Xolotlán), which also served as a strategic location against Pedro de Alvarado in Nequepio (modern El Salvador). The inhabitants of Nagrando were described as speakers of mexicana corrupta (i.e., Nahuat); this identifies them as the Nicarao, who totaled around 15,000 and who also fought against the conquistadors. When Oviedo lived in León (1528–­1529), there were already more than 200 Spanish residents living in wooden homes with thatched roofs. Fernández de Oviedo y Valdés (1959) also mentions the presence of Matagalpa/Chontals and Maribios/Subtiaba living amongst the Nicarao. We can therefore see the city of León and its surroundings as a clear example of a pluricultural confluence area at this time. In 1526, Hernández de  Córdoba was executed in the plaza of León as a declared enemy of Pedrarias, who had originally sent him to conquer Nicaragua and who was subsequently named governor of Nicaragua (and Nicoya) in 1527. The later-­arriving Pedrarias visited the island of Chira in the Gulf of Nicoya in March of 1526, after the abandonment of the Spanish settlement of Villa Bruselas on the northern coast of the gulf. His visit was in response to the menacing presence of his rival Hernández de Córdoba in this territory. Pedrarias’s intention was to demonstrate his power over other conquerors and (of course) to impress the Indigenous inhabitants. While his stay on Chira Island provides an example of how the Spanish tried to impose Christianity on local peoples, his account also tells us who were the Indigenous inhabitants of the island and provides some information on how they lived (Ibarra Rojas 2001:95–­98). Pedrarias and his officers stayed in a bohío, a (typically) round building built of pole and thatch. He brought soldiers and priests with him and insisted on carrying out a religious ceremony, with a cross he would plant on the most important port on the island. This place is perhaps the one located toward the peninsula, nearest Nicoya. The cross was carved by Indigenous artisans, and it is said they did an extraordinary job, which suggests that wood carving was a common practice for some of the men of the island. The existence of another bohío is mentioned, where inhabitants supposedly held ceremonies. In it, on a pole or wooden plank that hung from the beams, they had what sources describe as “idols.” Through an interpreter of the (presumably) Chorotega language named Lucas, the Spanish understood the locals also made sacrifices in this structure. Pedrarias broke the “idols”—­ceramic vessels—­with a stick and I ndigenous P eoples of Pacific N icaragua and N icoya in the S ixteenth C entury

57

destroyed the wooden plank. A church was built on that same spot with the aid of the Indigenous inhabitants, though this is not described in all sources. Pedrarias also tried to introduce religious paraphernalia, songs, flags, and other Christian symbols. How successful this introduction was is unknown, though he does confirm the local inhabitants’ participation in the religious ceremonies. In sum, while much of Pedrarias’s own account does not ring totally true, it does appear to provide interesting details of Indigenous life in the island. Oviedo, named officially as chronicler by the Spanish crown, was in Nicaragua and Nicoya in 1528 and 1529, as mentioned above. His descriptions of life in this region have greatly enriched ethnohistorical syntheses of this Central American Pacific area, and different interpretations of his writings have contributed significantly to our understanding of life here in the sixteenth century and, for some time before, providing a useful context for interpreting the results of archaeological research (for example, Abel-­Vidor 1986; Creamer 1983; Salgado González 1996). While in Nicaragua, Oviedo primarily involved himself in asking questions about (and seeking to understand) topics ranging from geography and volcanoes to flora and fauna to the Indigenous groups themselves, though he also wrote about the conflicts between the conquistadors. A keen observer, Oviedo describes numerous customs and ceremonial practices of the inhabitants of Nicaragua and reports their use of several distinct languages. His information is also indispensable for the identification and reconstruction of settlement patterns and for the understanding of sociopolitical hierarchies and the exercise of political power that I emphasize here. Oviedo’s descriptions of what he tried hard to understand—­that is, the Indigenous political organization—­ allows us to interpret a sociopolitical structure for Tecoatega, a Nicarao center close to León. The teyte was the principal chief or cacique and was assisted by an elders’ council and a captain’s body composed of the smaller chiefs of the towns, or pueblos, within the cacicazgo. The relationships between these leaders were characterized by subordination relationships, with the smaller caciques of the lesser settlements occupying the lowest positions in the hierarchy. In terms of tribute, the lesser caciques, as the heads of diverse hierarchized settlements, intermediated between their production, the captains, and the teyte. In this way, production arising from families—­that is, calpultin—­would be channeled to the teyte to be exchanged in local and foreign trade networks (Ibarra Rojas 2014). Oviedo reports that the Spanish did not fully understand the Indigenous sociopolitical structure and found it difficult to communicate with numerous caciques for the extraction of tribute. They therefore dissolved the Indigenous 58

I BA R R A RO JA S

political structure, leaving only one of the caciques to take care of what had been previously attended by many. Such political decisions contributed to modify and eventually destroy Tecoatega’s original ways of handling politics, society, and the economy. Oviedo also reports that there were elections for the members of the elders’ council, a common practice in the region, as I shall discuss next. In 1534, Bartolomé de Las Casas visited Nicoya, a fact that has not often been emphasized in the historiography of some Central American countries, though I consider his writings to be basic to understanding the Indigenous sociopolitical systems of the time. Besides defending Indigenous peoples from the hardships imposed upon them by the conquerors—­including cruelty, abuse, and slavery—­Las Casas writes about Indigenous political organizations in Honduras and Nicaragua, including Nicoya, and mentions that caciques were elected by the people. In Honduras, he reports, caciques were elected by all the members of the pueblo and governed for sixteen months. Once in Nicoya, he reports another election but states that because he did not understand the language—­Chorotega—­he could not ask how long the elected one would remain in his post (Las Casas 1966:179). A regional overview, then, demonstrates that the election of officers was a common pre-­Columbian political practice in Honduras, Pacific Nicaragua, and Nicoya. This way of choosing leaders continues to be practiced by many modern Indigenous peoples, for instance, the Cunas in Panama and the Ngobes in Costa Rica. It is also true that today, in various towns of Nicaragua and in Nicoya, members of the colonial institution known as the cofradía are still elected popularly. Las Casas’s information not only provides an invaluable context for the study of old Indigenous practices that still persist, but also helps us to understand how older political structures functioned in the past. THE SOCIOPOLITICAL STRUCTURE IN NICOYA

The sociopolitical structure in Nicoya was different from that in Tecoatega, as described above. All the evidence reviewed points consistently to the possible presence of a dual organization there. This type of organization in southern Central America should be viewed in the context of pueblos juntos that have been previously recognized in other areas of Costa Rica, Panama, and even Colombia. In earlier publications (Ibarra Rojas 1984), I proposed that this type of sociopolitical organization was present in different parts of Costa Rica, such as in Atirro, where Atirro el de arriba and Atirro el de abajo are located, and in Turrialba, where Turrialba la grande and Turrialba la chica I ndigenous P eoples of Pacific N icaragua and N icoya in the S ixteenth C entury

59

are found. It is also notable that the fortress at Couto had two palenques, one larger than the other, and that Térraba is described in the literature as Térrebi el grande and Térrebi el chico. This dualist organization was also mentioned by Henri Pittier and Doris Stone and was studied by María Eugenia Bozzoli de Wille (1979:41–­57). All of this work provides a solid basis for arguing that a dual organization might also be found in Nicoya, especially considering that Chibchan speakers who might have favored this type of organization occupied that territory before the Mesoamerican migrations. Jeffrey Quilter and Jeffrey Frost (Frost 2009; Quilter and Frost 2007) have also found archaeological evidence supporting some form of societal dualism at the Panteón de la Reina Cemetery at Sitio Rivas in southern Costa Rica. Frost suggests this dualism was also present in Chibchan societies in Gran Chiriquí, as well as among the Chibchan-­speaking Kogi of the Sierra Nevada de Santa Marta in Colombia. What might a dualist society look like? Claude Lévi-­Strauss (1969) famously explained how the two halves in such a society might behave with one another and how they can be very intimate even when hostility may also be present. Women might be exchanged, and economic, social, and religious practices might also be shared. Dualist societies are usually exogamous and matrilineal, and persons typically define themselves as members of one half or the other. In the case of Nicoya, the evidence for dual organization does not come from either archaeological research or from clear descriptions of practices provided in the earliest Spanish accounts. Here the evidence derives primarily from the historical sources discussing the organization of tribute exaction in Nicoya, which make repeated mention of Nicoya having two caciques, from early through colonial times. In 1522, even though Gil González Dávila himself did not mention two caciques, his treasurer, Andrés de Cereceda, did: Caciques Nicoya and Mateo, both of whom gave him gold. In 1541, two caciques are also mentioned for Nicoya when it was given as an encomienda to María de Peñalosa (Vega Bolaños XIV, 1956:172). Moreover, the tasaciones de tributos (tribute assessment) of 1548 (Vega Bolaños XIV, 1956:402, 404) and of 1677 (Cea Cuadra 1948) both mention two Nicoyas, one larger than the other, and two caciques. In 1554 Nicoya is included with its two parcialidades as part of the Corregimiento (Solórzano Fonseca and Quirós Vargas 2006:155). In 1564 it is noted that the people of Nicoya do not live in one town but are spread out and separated by rivers or streams (Fernández Bonilla 1907, V.VII:130), in an arrangement resembling the dual settlement pattern found in Turrialba. Duality is also indicated in documentary sources from 1585, which report that tributes were extracted from both 60

I BA R R A RO JA S

Nicoyas, and in 1643, when the first cofradía (i.e., the Cofradía del Santísimo Sacramento) was founded, two caciques of Nicoya (Don Blas de Contreras and Don Diego de Mendoza) offered the first herd of cattle for its functioning (Quirós Vargas 1997). Later mentions of Nicoya’s parcialidades de abajo y de arriba are found in 1687 and 1772 (ANCR 1772) as well as in 1751, when Obispo Morel de Santa Cruz visited Nicoya and found two parcialidades: la de arriba, with seventy-­one houses with thatched roofs, and la de abajo, with forty-­four (Quirós Vargas 1997:67). It is therefore clear that a dualist settlement pattern could have characterized Nicoya. The above information also supports the various hypotheses that this dualism was pre-­Columbian, that it may have originated before the Mesoamerican migrants arrived, and that it did not disappear with them. One suspects that these migrants probably accommodated parts of their own lifestyles, production, and exchange to this existing pattern (Ibarra Rojas 2014), though this possibility still needs to be researched more fully, especially by archaeologists. The possible dualist pattern in Nicoya suggests a consideration of the following. As previously noted, Nicoya in the sixteenth century was a pluricultural confluence space, occupied mainly by Chorotegan speakers who shared just a portion of the peninsula with Corobici or Rama speakers. All the evidence suggests that the Chorotega were politically and economically dominant, in the sixteenth century and perhaps even earlier. This begs the question: what occurred between the original inhabitants and the Chorotega newcomers through the years to produce the sixteenth-­century state of affairs? As I have noted elsewhere (Ibarra Rojas 2012), the Corobici/Rama seem to have controlled a trade route that connected with the Solentiname Islands and the mouth of the Desaguadero River. Could there have been an alliance between the Chorotega and the Corobici with respect to that precise route? Alternatively, were the Chorotega simply forced to cope with existing settlement patterns and trade systems and networks? Supposing that the dualist organization in Nicoya predated the arrival of the migrants, is its continued presence in the sixteenth century a clue as to how the Chorotega related themselves with respect to the older inhabitants (Ibarra Rojas 2011a, 2012)? It should be considered that this type of organization might have seemed familiar to the Mesoamerican migrants, since dualism was also an important part of other Mesoamerican cultures, such as Cacaxtla and Xochicalco between ad 600 and ad 900 (Palavicini Beltrán 2011). It should also be remembered that Torquemada wrote concerning the migrants into Nicaragua that “su intento no era de matar sino de rendir gentes y sujetarlas al imperio mexicano” (i.e., the I ndigenous P eoples of Pacific N icaragua and N icoya in the S ixteenth C entury

61

intention was not to kill people but to subject them to the Mexican empire) (Torquemada, cited in Incer 2002:462–­464). There could be a relationship here between history and myth, for the sources do not speak of the destruction of the older inhabitants’ ways or homes, nor of their replacement by entirely new forms. Modification, yes; destruction, doubtful. As is now obvious, political organization in the northern area of the Pacific region was different from that in Nicoya. What underlay the cultural specificities of both areas was a general baseline sociopolitical organization based on cacicazgos, similar systems of production and exchange, and (perhaps) similar resources. However, while the Spanish conquerors made changes to the political structure in Tecoatega to create a more efficient (by their own standards) tribute-­collection system, this did not happen in Nicoya, and therefore the destructuring of the Indigenous sociopolitical and economic institutions in both areas did not occur at the same time nor in the same way. This fact can help us to understand why, even today, there yet remains a permanent cultural divide between the Pacific regions of Nicaragua and Nicoya that continues to be expressed through the persistence of old Indigenous ways in daily cultural practices. CONCLUSIONS

In this chapter I have briefly discussed and analyzed how a dynamic, historical approach to the study of the Indigenous peoples who inhabited a part of Central America in the sixteenth century can provide deeper and richer insights into how these groups related to one another. The regional perspective and comparative analysis presented here allows us to distinguish, with precision, differences between the sociopolitical organizations of the cacicazgos in northern Pacific Nicaragua and in Nicoya prior to the colonial era. It is clear that much of what was associated with the older inhabitants of the region—­languages, customs, cultural practices, important aspects of trade networks, practices surrounding resources and agriculture, even settlement patterns—­was not destroyed by the Mesoamerican migrants upon their arrival. Torquemada’s words provide at least a partial clue as to how and why this happened. One conclusion my research suggests is that the Chorotega tried hard, and with considerable success, to control the production of the older Chibchan inhabitants. Their process of establishing dominance in the area involved letting their predecessors keep some of their old customs, as well as their language and religion. This approach also provided the original inhabitants with a space for resistance, since even if they were 62

I BA R R A RO JA S

moved to change in some aspects, it gave them a way to remember their unique identity. Examining the processes associated with the occupation of this region, first by the Mesoamerican-­originating groups and then by the Spanish conquistadors and colonizers, allows for a dynamic reconstruction of its history. Abandoning the culture area concept and employing an alternative approach that uses a pluricultural confluence space-­based scenario for studying the development of human relationships across space and time sheds light upon complex processes and opens doors to additional future research that will be oriented to think about this region using this historically rooted approach. The evidence that the sharing of some cultural traits was possible in Nicoya suggests that other traits that might be of interest to future researchers—­techniques for producing ceramics, for example—­might also have been shared or combined. So, getting back to the earlier question concerning who went out to fight the first Spaniards who approached the Gulf of Nicoya in 1519 with this perspective in mind, I suggest that it quite likely was a combined force of Chorotega and Corobici/Rama, who very well could have allied to fight together against a common enemy, just as they likely worked together in other endeavors, even if under unequal conditions. REFERENCES

Archivo Nacional de Costa Rica, San José, Costa Rica (ANCRF). 1772. Orden de Juan Antonio de la Peña y Medrano, Corregidor de Nicoya, para que no se trasladen los indios. CC 321, f. 1. Abel-­V idor, Suzanne. 1986. “Early Sixteenth Century Evidence for the Settlement Archaeology of Greater Nicoya.” In Prehistoric Settlement Patterns in Costa Rica, edited by Frederick W. Lange and Lynette C. Norr, 387–­405. Journal of the Steward Anthropological Society 14 (1–­2 [1982–­1983]). Urbana: University of Illinois. Álvarez Rubiano, Pablo. 1944. Pedrarias Dávila. Contribución al estudio de la figura del “Gran Justador,” gobernador de Castilla del Oro y de Nicaragua. Madrid: Consejo superior de investigaciones científicas, Instituto Gonzalo Fernández de Oviedo. Berdan, Frances F., and Michael E. Smith. 2004. “El sistema mundial mesoamericano postclásico.” Relaciones: Estudio de Historia y Sociedad 25(99):19–­7 7. Bozzoli de Wille, and María Eugenia. 1979. El nacimiento y la muerte entre los bribris. San José: Editorial Universidad de Costa Rica. Carmack, Robert M., and Silvia Salgado González. 2006. “A World-­Systems Perspective on the Archaeology and Ethnohistory of the Mesoamerican / Lower Central American Border.” Ancient Mesoamerica 17(2):219–­229. I ndigenous P eoples of Pacific N icaragua and N icoya in the S ixteenth C entury

63

Cea Cuadra, Luis, ed. 1948. El libro de tributos de la Provincia de Nicaragua en los años 1662 y 1692. Manuscript on file, Biblioteca del Banco Central de Nicaragua, Managua. Constenla Umaña, Adolfo. 1994. “Las lenguas de la Gran Nicoya.” Vínculos: Revista de Antropología del Museo Nacional de Costa Rica 18–­19 (1–­2 [1992–­1993]):191–­208. Creamer, Winifred. 1983. “Production and Exchange on Two Islands in the Gulf of Nicoya, Costa Rica, ad 1200–­1550.” PhD diss., Tulane University, New Orleans. Fernández Bonilla, León, ed. 1907. Colección de documentos para la historia de Costa Rica. Vol. 7. Edited by Ricardo Fernández Guardia. Barcelona: Imprenta Viuda de Luis Tasso. Fernández de Oviedo y Valdés, Gonzalo. (1520–­1541) 1959. Historia general y natural de las Indias. Edited by Juan Pérez de Tudela Bueso. Biblioteca de Autores Españoles, vol. 4, bk. 120. Madrid: Ediciones Atlas. Finch, William O., and Kim Honetschlager. 1986. “Preliminary Archaeological Research on Isla del Caño.” In Prehistoric Settlement Patterns in Costa Rica, edited by Frederick W. Lange and Lynette C. Norr, 189–­206. Journal of the Steward Anthropological Society 14 (1–­2 [1982–­1983]). Urbana: University of Illinois. Frost, R. Jeffrey. 2009. “The Ancestors Above, the People Below: Cemeteries, Landscape and Dual Organization in Late Pre-­Columbian Costa Rica.” PhD diss., University of Wisconsin, Madison. Ibarra Rojas, Eugenia. 1984. “Los cacicazgos indígenas en la Vertiente Atlántica y Valle Central de Costa Rica: Un intento de construcción etnohistórica.” Licentiate’s thesis, Universidad de Costa Rica, San Jóse. Ibarra Rojas, Eugenia. 1988. “El intercambio y la navegación en el Golfo de Huetares (o de Nicoya) durante el siglo XVI.” Revista de Historia 17:35–­67. Ibarra Rojas, Eugenia. 1990. Las sociedades cacicales de Costa Rica (Siglo XVI). Colección Historia de Costa Rica 3. San José: Editorial de la Universidad de Costa Rica. Ibarra Rojas, Eugenia. 2001. Fronteras étnicas en la conquista de Nicaragua y Nicoya: Entre la solidaridad y el conflicto, 800 d.C.–­1544. San José: Editorial de la Universidad de Costa Rica. Ibarra Rojas, Eugenia. 2011a. “Los nicaraos, los indios votos y los huetares en escenarios conflictivos en el siglo XVI.” Cuadernos de Antropología 21:1–­24. Ibarra Rojas, Eugenia. 2011b. Del arco y la flecha a las armas de fuego: Los indios mosquitos y la historia centroamericana 1633–­1786. San José: Editorial Universidad de Costa Rica. Ibarra Rojas, Eugenia. 2012. “Tras los pasos de los corobicíes en el Siglo XVI, del Golfo de Nicoya a las Islas de Solentiname en el Lago de Nicaragua.” In

64

I BA R R A RO JA S

Proceedings of the 1st Intercontinental Conference of the Society for American Archaeology, 79–­88. Panama City, January 13–­­15. Ibarra Rojas, Eugenia. 2014. Entre el dominio y la resistencia: Los pueblos indígenas del Pacífico de Nicaragua y Nicoya en el siglo XVI. San José: Editorial de la Universidad de Costa Rica. Ibarra Rojas, Eugenia, and Silvia Salgado González. 2010. “Áreas culturales o regiones históricas en la explicación de relaciones sociales de pueblos indígenas de Nicaragua y Costa Rica de los siglos XV y XVI.” Anuario de Estudios Centro­ americanos 35–­36(2009–­2010):37–­60. Lange, Frederick W. 1979. “Theoretical and Descriptive Aspects of Frontier Studies.” Latin American Research Review 14(1):221–­227. Lange, Frederick W. 1994. “Evaluación histórica del concepto Gran Nicoya.” Vínculos: Revista de Antropología del Museo Nacional de Costa Rica 18–­19(1–­2 [1992–­1993]):1–­8. Lange, Frederick W. 2015. “La cuña (‘WEDGE’) histórico-­cultural entre el sur de Mesoamérica y la Gran Nicoya: Una propuesta.” In Contribuciones del Dr. Michael J. Snarskis a la arqueología costarricense, edited by María del Carmen Araya Jiménez and Silvia Salgado González, 123–­143. San José: Universidad de Costa Rica. Las Casas, Bartolomé de. 1996. Los indios de México y Nueva España. Antología. Edited by Edmundo O’Gorman and Jorge Alberto Manrique. Colección “Sepan cuantos” 57. Mexico City: Editorial Porrúa. Lévi-­Strauss, Claude. 1969. Las estructuras elementales del parentesco. Translated by Marie Thérèse Cevasco and Paidós Básica. Buenos Aires: Editorial Paidós. McCafferty, Geoffrey G., Silvia Salgado González, and Carrie L. Dennett. 2009. “¿Cuándo llegaron los mexicanos? La transición entre los periodos Bagaces y Sapoá en Granada, Nicaragua.” In Proceedings of the Third Congreso Centroamericano de Arqueología en El Salvador, October 2009. San Salvador: Museo Nacional de Antropología “Dr. David J. Guzmán.” Palavicini Beltrán, Beatriz. 2011. “Instituciones políticas y gobiernos duales en la transición del epiclásico al posclásico.” Estudios Mesoamericanos n.s., 10(1):63–­68. Peralta, Manuel María de. 1883. Costa Rica, Nicaragua y Panamá en el siglo XVI, su historia y sus límites según los documentos de archivo de Indias de Sevilla, del de Simancas etc., recogidos y publicados con notas y aclaraciones históricas y geográficas. Madrid: Editorial Manuel Ginés Hernández. Quilter, Jeffrey, and R. Jeffrey Frost. 2007. “Investigaciones en el complejo arqueo­ lógico Rivas—­Panteón de la Reina en el suroeste de Costa Rica.” Vínculos: Revista de Antropología del Museo Nacional de Costa Rica 30(1–­2):23–­56.

I ndigenous P eoples of Pacific N icaragua and N icoya in the S ixteenth C entury

65

Quirós Vargas, Claudia. 1997. “Las cofradías indígenas en Nicoya.” Revista de Historia 36:37–­77. Salgado González, Silvia. 1996. “Social Change in a Region of Granada, Pacific Nicaragua (1000 bc–­1522 ad).” PhD diss., State University of New York, Albany. Salgado González, Silvia, and Elisa Fernández León. 2011. “Elementos para el estudio de una migración antigua: El caso de los Chorotega-­Mangue.” Cuadernos de Antropología 23:1–­30. Salgado González, Silvia, and Ricardo Vázquez Leiva. 2006. “Was There a Greater Nicoya Subarea during the Postclassic?” Vínculos: Revista de Antropología del Museo Nacional de Costa Rica 29(1–­2):1–­16. Sheets, Payson D. 2008. “Memoria social perdurable a pesar de desastres volcánicos en el área de Arenal.” Vínculos: Revista de Antropología del Museo Nacional de Costa Rica 31(1–­2):1–­26. Sheets, Payson D., and Marilyn Mueller. 1984. “Investigaciones arqueológicas en la Cordillera de Tilarán Costa Rica 1984.” Vínculos: Revista de Antropología del Museo Nacional de Costa Rica 10(1–­2):3–­5. Sheets, Payson D., and Thomas L. Sever. 1991. “Prehistoric Footpaths in Costa Rica: Transportation and Communication in a Tropical Rainforest.” In Ancient Road Networks and Settlement Hierarchies in the New World, edited by Charles D. Trombold, 53–­65. New Directions in Archaeology. Cambridge: Cambridge University Press. Solórzano Fonseca, Juan Carlos, and Claudia Quirós Vargas. 2006. Costa Rica en el siglo XVI: Descubrimiento, exploración y conquista. San José: Editorial de la Universidad de Costa Rica. Steinbrenner, Larry. 2006. “Cacao in Greater Nicoya: Ethnohistory and a Unique Tradition.” In Chocolate in Mesoamerica: A Cultural History of Cacao, edited by Cameron L. McNeil, 253–­270. Maya Studies Series, general editors Diane Z. Chase and Arlen F. Chase. Gainesville: University Press of Florida. Torquemada, Juan de. 2002. “Sobre la Provincia de Nicaragua y sus habitantes.” In Descubrimiento, conquista y exploración de Nicaragua: Crónicas de fuentes originales seleccionadas y comentadas por Jaime Incer Barquero, edited by Jaime Incer Barquero, 462–­473. Colección Cultural de Centro América, Serie Cronistas 6. Managua: Fundación Vida. Vega Bolaños, Andrés, ed. 1956. Documentos para la Historia de Nicaragua. Edited by Andrés Vega Bolaños. Colección Somoza 14. Madrid: Imprenta y Litografía Juan Bravo.

66

I BA R R A RO JA S

4 The dating of archaeological components in Pacific Nicaragua rests, to a very great extent, on the archaeologist’s ability to identify the various ceramic types associated with these components that have been previously assigned to an established chronological sequence and that can therefore be treated as de facto chronological markers. While this approach initially served archaeologists well following the development of the first chronological sequence for all of Greater Nicoya based on uncalibrated radiocarbon dates in the 1960s, the collection of numerous additional radiocarbon dates in the intervening half-­century has resulted in an unusual situation in which the modern iteration of the Greater Nicoya chronological sequence no longer appears to be entirely consistent with the cumulative database of calibrated radiocarbon dates upon which it is ostensibly based. This is particularly true with respect to the later, post–­a d 800 periods in this sequence that have traditionally been most closely associated with the arrival of Mesoamerican influence and/or groups in southern Central America. As a result, many ceramic types once thought to be temporally diagnostic of the period immediately preceding European contact in the early 1500s now appear to be much older—­a finding that has significant implications for the interpretation of late-­period archaeological sites in Pacific Nicaragua and their connections to Mesoamerica.1 1. See chapters 9 and 10 in this volume for more discussion of changing views regarding these key types.

A Critical Reevaluation of Pacific Nicaragua’s Late Period Chronology Larry Steinbrenner and Geoffrey G. McCafferty

DOI: 10.5876/9781646421510.c004

67

In this chapter, we review the historical development of the entire Greater Nicoya chronological sequence with an eye toward understanding the current inconsistency between the modern iteration of the sequence that has been in widespread use since the early 1990s and the actual radiocarbon database. We then discuss how calibrated 14C dates produced by more recent archaeological research focusing primarily upon Pacific Nicaragua and Guanacaste, including more than two dozen dates reported by our own research in the departments of Rivas and Granada, make a very strong case for reevaluating the temporal significance attributed to certain key ceramic types. We conclude with a discussion of some of the challenges posed by this reevaluation with respect to attempts to identify and understand archaeological contexts associated with the last phase of Greater Nicoya’s prehistory immediately prior to European contact. THE GREATER NICOYA CHRONOLOGICAL SEQUENCE, 1960 s –­1 990 s

The first modern archaeological sequence to be proposed for Greater Nicoya was tentative and subject to frequent adjustment by its developers, Michael Coe and Claude Baudez. Working separately, Coe and Baudez carried out stratigraphic excavations at a number of sites in Costa Rica’s Nicoya Peninsula and Guanacaste in the late 1950s and early 1960s (Baudez 1967; Baudez and Coe 1962; Coe 1962; Coe and Baudez 1961). In their earliest joint publication (Coe and Baudez 1961), the researchers named four chronological periods that could be individually characterized by diagnostic ceramic types: the Zoned Bichrome and Early, Middle, and Late Polychrome periods. No specific dates were given for the individual periods. A subsequent publication included dates (table 4.1) and tentatively split the Early Polychrome period into “Early Polychrome A” and “Early Polychrome B” periods (Baudez and Coe 1962), though this refinement was abandoned by most later researchers (cf. Lange et al. 1984:201), as was Baudez’s (1967) later suggestion for a “Linear Decorated period” intermediate between the Zoned Bichrome and Early Polychrome periods.2 The original four periods were explicitly intended to correlate roughly with the Maya sequence of late Formative, early–­late Classic, late Classic–­early Postclassic, and late Postclassic (Coe and Baudez 1961:505) and can therefore be said to reflect a perception that Greater Nicoya should be 2. The Linear Decorated period is generally treated as a local phase in the Tempisque Valley sequence (Lange et al. 1984:201).

68

S T EI N B R EN N ER A N D M CCA F F ERT Y

considered part of Mesoamerica (Haberland 1978:403; cf. Coe 1962). This four-­ period sequence was commonly used by archaeologists until the early 1990s and is still used in many Nicaraguan museums. In this present study, we take a particular interest in the two later periods in this sequence and their diagnostic decorated ceramic types: (1) the Middle Polychrome or Sapoá period, characterized especially in Pacific Nicaragua by Papagayo, Pataky, and Granada polychromes; and (2) the Late Polychrome or Ometepe period, marked by polychromes such as Vallejo, Madeira, Luna, El Menco, and Bramadero and monochromes such as Castillo Engraved and Ometepe Red Slipped-­Incised (see figure 4.1). As table 4.1 illustrates, the earliest published incarnations of the four-­period sequence were only broadly consistent with respect to the actual dating of the individual periods, presumably reflecting both the limited and uncalibrated radiocarbon database that was used to ascribe calendar date ranges to these periods as well as Coe and Baudez’s evolving thinking about their sites and the connection between their own periodization and the Mesoamerican chronologies upon which they were modeled. In the 1960s, the embryonic radiocarbon database for Greater Nicoya comprised fewer than ten radiocarbon dates recovered by Coe and Baudez—­the first ever reported for the region—­ and seven dates from Nicaraguan sites based on contemporary work by Gordon Willey and Alfred Norweb (who excavated several sites in Pacific Nicaragua between 1959 and 1961) and Wolfgang Haberland and Peter Schmidt (who focused on Ometepe Island in Lake Nicaragua in the early 1960s).3 No single period in Coe and Baudez’s periodization was defined using more than four uncalibrated radiocarbon dates, and in Costa Rica proper, the Early and Middle Polychrome periods were each associated with single dates. Additionally, some of these dates were less than ideal for chronology building. For example, we believe that the seemingly “late” dates from the La Bocana cave site (Baudez 1967:24)—­samples GsY-­98 and GsY-­99, two of the total four dates used to help define the “Late Polychrome” period—­should 3. While most of this Nicaragua-­based research was not published in detail until later (e.g., Haberland 1978, 1992; Healy 1974, 1980;), the dates that derived from this work appear to have been freely circulated and were occasionally cited by other researchers despite their not having been previously published. Examples of the early version of the Greater Nicoyan 14C database can be found in Lange (1971:143) and Sweeney (1975:36). The only 14C dates from Nicaragua that were formally published in the 1960s (see Norweb 1964:555–­556) comprised three Early Polychrome–­period dates: one from the Cruz Site on Ometepe Island and two from the “Granada region” (i.e., the Ayala site; see Vázquez Leiva et al. 1994:249). A C ritical R eevaluation of Pacific N icaragua’ s L ate P eriod C hronology

69

Figure 4.1. Pacific Nicaraguan ceramic types commonly treated as being “diagnostic” of the Sapoá (or Middle Polychrome) period (a–­c) and the Ometepe (Late Polychrome) period (d–­j). (a) Papagayo Polychrome. (b) Pataky Polychrome. (c) Granada Polychrome. (d) Vallejo Polychrome. (e) El Menco Polychrome. (f ) Luna Polychrome. (g) Castillo Engraved. (h) Madeira Polychrome. (i) Bramadero Polychrome. (j) Ometepe Red Slipped-­ Incised. Figures a, e (bottom), g, h, i, j: Mi Museo, Granada. Figure b: National Museum, Managua. Figures c, f: Rivas Museum, Rivas. Figure d: private collection. Figure e (top): Acahualinca Museum, Managua. All photos by Larry Steinbrenner except h (Mi Museo).

be treated as problematic because the only “Late Polychrome” diagnostic ceramics recovered from the site comprised two sherds in the surface collection, while all ceramic material from actual excavations of the site dated to the 70

S T EI N B R EN N ER A N D M CCA F F ERT Y

T able 4.1. Variations on the four-­period system originally developed by Coe and Baudez based on Maya periodization Period

Zoned Bichrome

“Standard” Maya Sequence

Late Formative

Baudez and Coe Coe (1962:369) (1962:174)

Norweb (1964:553) a

Healy (1980:343–­345) b

ad 400–­ 800

ad 300/400–­800

300 bc– ad 200

ad 0 [sic]–­ 300

ad 0 [sic]–­ 400

500/350 bc–­ ad 300/400

Early Polychrome

Early–­Late Classic

ad 200– ­800c

ad 800– ­1175

ad 750–­ 1000

ad 800–­ 1000

ad 800–­1200

Late Polychrome

Late Postclassic

ad 1175–­ conquest

ad 1000– ­conquest

ad 1000–­ 1600

ad 1200–­1522

Middle Polychrome

Late Classic–­Early Postclassic

ad 300–­ 750

Based on Coe (1962). Modified from Baudez and Coe (1962). c Subdivided into EPA (ad 200–­450) and EPB (ad 450–­800); not subsequently used.

a

b

centuries-­older Zoned Bichrome period.4 Given the limitations of the available database, it is therefore not surprising to find that the ranges assigned to each period could vary considerably—­even between two papers published in the same year, as a comparison of the dates from Baudez and Coe (1962) and Coe (1962) provided in table 4.1 shows. The dates in Coe (1962) provided the model for the first chronological sequence for Pacific Nicaragua proposed by Norweb (1964:553), which gave the dates for the four periods as, respectively, ad 0–­400 [sic], ad 400–­800, ad 800–­1000, and ad 1000–­1600. However, most subsequent studies in the following decade (e.g., Haberland 1966; Healy 1974; Lange 1971; Sweeney 1975), followed the lead of the seemingly more refined Baudez and Coe (1962) sequence in extending the range of the Zoned Bichrome period several centuries into the past, dating the transition between the Early and Middle Polychrome periods to about ad 800 (as did Norweb) and pushing the Middle to Late Polychrome transition two centuries later to approximately ad 1200. For example, Paul Healy’s (1980:343–­345) chronological sequence for the 4. The four original uncalibrated dates used to define the Late Polychrome period were from La Bocana (GsY-­98, ad 1484 ± 85, and GsY-­99, ad 1408 ± 85) and Chahuite Escondido (Y-­816, ad 1120 ± 70) in Costa Rica and San Lázaro on Ometepe Island (Hv-­2692, ad 1445 ± 30) (cf. and Lange 1971:143; Sweeney 1975:36). The La Bocana samples are discounted here; the relatively early Y-816 date and the much later Hv-­2692 date are discussed subsequently in this chapter. A C ritical R eevaluation of Pacific N icaragua’ s L ate P eriod C hronology

71

Rivas area of Pacific Nicaragua (which supplanted Norweb’s as the “standard” sequence for this area for the next several decades) suggested the following period ranges: 500/350 bc–­a d 300/400, ad 300/400–­800, ad 800–­1200, and ad 1200–­1522. Eventually, the adoption of Baudez and Coe’s (1962) sequence by more and more scholars contributed to its widespread acceptance and use in the ensuing decades (and perhaps even a sort of de facto reification), in spite of its originally tentative nature and despite the fact that later-­reported 14C dates provided only partial support for the sequence and were especially inconsistent with respect to the dating of supposed “Late Polychrome–­period” contexts (a situation mostly recognizable in hindsight). A contemporary observation made by Jean Sweeney (1975:33) concerning her era’s 14C database—­that is, “the mere fact that everyone working in Lower Central America reports the same set of radiocarbon dates has given them an aura of sanctity”—­might have been equally well applied to the chronological sequence, which seems to have acquired, for better or worse, a similar aura of its own. The radiocarbon database for Greater Nicoya continued to grow throughout the 1970s and 1980s, though initially this expansion did not directly reflect important new investigations in Greater Nicoya in the late 1960s and early 1970s, such as Fred Lange’s (1971) survey of Costa Rica’s Río Sapoá Valley. In fact, almost all “new” dates reported prior to the 1980s derived directly from earlier work by Coe and Haberland. Without question, the most significant addition to the 14C database in the 1970s was provided by Sweeney’s 1975 dissertation, which added seventeen previously unreported dates from sites in northwestern Costa Rica excavated by Coe. Most of these new dates were associated by Sweeney with the Middle Polychrome period, and they helped to better define this critical period (see Sweeney 1975:380ff ).5 However, two of the new dates from the Chahuite Escondido site on the Santa Elena Peninsula associated with high frequencies of supposed “Late Polychrome-­ period” diagnostic ceramic types such as Vallejo and Bramadero polychromes (i.e., samples P-2168 and P-2169) were both earlier than the ad 1200 date that was generally used to mark the beginning of the Late Polychrome period, with uncalibrated ranges from the ninth to the eleventh century that actually—­and 5. Felipe Del Vecchio Solís (2017) has recently observed that about a half-­dozen of the samples that generated these “Middle Polychrome” dates were actually recovered from contexts that also contained “Late Polychrome–­period” diagnostic ceramics, albeit in very small quantities. The low frequency of these diagnostics in contexts rich in Middle Polychrome period material likely explains why Sweeney chose not to associate these dates with the Late Polychrome period.

72

S T EI N B R EN N ER A N D M CCA F F ERT Y

intriguingly—­fell within the span of time indicated by Sweeney’s more numerous Middle Polychrome–­period dates (Sweeney 1975:380, 394ff ). Sweeney’s study makes it clear that the same could be said for a third date from the site (Y-­816) that had been the earliest of the handful of dates used by Coe and Baudez to establish the Late Polychrome period (see note 4), despite its likely “pre-­a d 1200” date (table 4.2, figure 4.2).6 Collectively, these three seemingly “pre–­Late Polychrome period” dates were the only dates Sweeney associated with the period and its characteristic ceramics.7 Given that the Late Polychrome period had been defined in the original Baudez-­Coe sequence using a similarly small number of uncalibrated dates—­one of which, Y-816, probably predated ad 1200 and at least two of which, as previously noted, derived from a single problematic context—­in retrospect it is somewhat surprising that Sweeney’s new evidence associating supposed Late Polychrome ceramics with pre-­rather than post-­a d 1200 contexts did not raise a red flag and spur a reappraisal of the accepted sequence (or at very least, the position of Late Polychrome diagnostics within this sequence). However, while Sweeney (1975:394) herself cautiously interpreted her uncalibrated dates as suggesting that the Late Polychrome period may have started earlier at Chahuite Escondido (i.e., at the “end of the eleventh century”), subsequent work did not follow her lead and appears to have uncritically treated Sweeney’s dates as simple outliers rather than as being potentially “typical” for their contexts, which later work suggests was actually the case. Discounting the first “official” publication of several previously circulated Nicaraguan dates (Haberland 1978:405), the only truly new dates beyond Sweeney’s that were added to the 14C database during the 1970s included an additional date (Hv-­2669) from Ometepe Island (based on Haberland and 6. Baudez (1967:210) himself rejected Y-816 as being too early, presumably because it diverged quite widely from his own La Bocana dates. This appears to have led Healy (1980:309) to also reject P-2168 and P-2169 as being “too early for a Late Polychrome context.” However, if we alternatively accept Y-816 and reject Baudez’s La Bocana dates, there seems little reason to discount Sweeney’s early dates beyond their failure to confirm preconceptions that the Late Polychrome period began circa ad 1200. 7. It is notable (and somewhat ironic) that the calibrated 2σ ranges for P-2169 and Y-816 (which were unavailable to Sweeney at a time when the calibration of 14C dates had not yet been established as a standard practice) generally do extend past ad 1200 (as indicated in table 4.2), as do many other subsequently published dates from Pacific Nicaragua associated with similar diagnostic ceramics (see discussion later in this chapter). However, this observation does not change the fact that all three dates that Sweeney associated with the Late Polychrome period would have appeared to predate the Late Polychrome period as it was understood at the time.

A C ritical R eevaluation of Pacific N icaragua’ s L ate P eriod C hronology

73

T able 4.2. Previously reported Greater Nicoya radiocarbon dates associated with high frequencies of “Late Polychrome / Ometepe-­period” diagnostic ceramic types Sample

Site

P-­2168

Chahuite Escondido, Santa Elena Peninsula

Costa Rica

C Age (bp ) 2σ Interval (ad )

Original Reference

1070 ± 50

Sweeney (1975:35)

14

778 (1.0%) 790 809 (0.2%) 813 826 (0.9%) 841

863 (93.0%) 1040 P-­2169

1109 (0.4%) 1116

Chahuite Escondido, Santa Elena Peninsula

870 ± 40

Chahuite Escondido, Santa Elena Peninsula

840 ± 70

1040 (95.4%) 1276 Sweeney (1975:35)

San Francisco, Granada

747 ± 135

1020 (95.4%) 1430 Wyckoff (1976)

San Lázaro, Ometepe Island

505 ± 30

1330 (2.6%) 1340

GIF-­ 7228

La Pachona (I 43), Chontales

430 ± 60

Beta-­ 66956

Ayala (Ni-­Gr-­2), Granada

840 ± 60

Beta-­ 66957

Ayala (Ni-­Gr-­2), Granada

810 ± 60

1045 (8.5%) 1095

Beta-­ 140586

Santa Isabel (Ni-­Ri-­44), Rivas

1030 ± 90

776 (95.2%) 1190

Y-­816

Nicaragua

WSU-­? Hv-­ 2692

1042 (23.4%) 1106 Sweeney (1975:35) 1117 (72.0%) 1254

1396 (92.8%) 1448

Haberland (1978:405)

1406 (68.5%) 1530 Gorin (1990:257–­259)

1538 (26.9%) 1635

1041 (17.6%) 1107 Salgado González (1996:437) 1116 (77.8%) 1275

1120 (86.9%) 1288

Salgado González (1996:437) Niemel (2003:350)

1199 (0.2%) 1202

Note: 2-­sigma (2σ) intervals calibrated using OxCal v4.3.2 (Bronk Ramsey 2017) and the IntCal13 atmospheric curve (Reimer et al. 2013).

Schmidt’s earlier work there, but rejected by Haberland himself as being “much too late; probably contaminated”; see 1978:405) and a single date from the San Francisco site in Granada (Wyckoff 1976). Interestingly, the San 74

S T EI N B R EN N ER A N D M CCA F F ERT Y

Figure 4.2. Previously reported Greater Nicoya radiocarbon dates associated with high frequencies of “Late Polychrome / Ometepe-­period” diagnostic ceramic types.

Francisco date, though associated with “Late Polychrome–­period” diagnostic ceramics, overlapped comfortably with the three pre-­a d 1200 Chahuite Escondido dates that were associated with similar ceramics (cf. table 4.2). The evidence that there was a problem with the dating of the Late Polychrome period was slowly accumulating, albeit unnoticed. Archaeological research in Pacific Nicaragua was significantly disrupted by the 1972 Managua earthquake and the country’s late 1970s–­early 1980s political turmoil, and all new dates added to the Greater Nicoya 14C database in the 1980s came exclusively from work in northwestern Costa Rica, which experienced a florescence in research during this period.8 The biggest contributions 8. While the Sandinista revolution did not stop archaeological research in Nicaragua entirely, most new work carried out during the 1970s and 1980s did not actually contribute to the 14C database (e.g., Baker and Smith 1987; Bruhns 1974; Hughes 1980; Wyss 1983; etc.) and/or remained unpublished until the 1990s (e.g., Espinoza Pérez and Rigat 1994; Gorin 1990, 1992; Lange et al. 1992; Navarro Genie 1993; Rigat 1992; Rigat and González Rivas 1996; Rigat and Gorin 1993). In Costa Rica, notable work in the 1970s and 1980s that also did not contribute specifically to the 14C database (and which is therefore not discussed here) included projects focused on the Gulf of Nicoya (e.g., A C ritical R eevaluation of Pacific N icaragua’ s L ate P eriod C hronology

75

T able 4.3. Revised regional chronological sequence for Greater Nicoya Period

Date

Tempisque (formerly Zoned Bichrome)

500 bc–­a d 300

Orosí (not previously identified)

Bagaces (formerly Early Polychrome)

Sapoá (formerly Middle Polychrome)

Ometepe (formerly Late Polychrome)

2000–­500 bc ad 300–­800

ad 800–­1350

ad 1350–­1550

Source: Adapted from Vázquez Leiva et al. (1994:248, table 1).

to the database came from two doctoral dissertations by John Hoopes (1987, 19 dates) and Lynette Norr (1986, 6 dates), continued work directed or supervised by Fred Lange (Lange 1980, 2 dates; Lange and Stone 1984, 7 dates), and new work by Costa Rican archaeologist Juan V. Guerrero Miranda (Guerrero Miranda and Blanco Vargas 1987, 4 dates; Guerrero Miranda, Vázquez Leiva, and Solano 1992, 1 date). Various other projects (e.g., Aguilar Piedra 1984; Hurtado de  Mendoza and Alvarado Induni 1990; Ryder 1986; Sheets 1986) also contributed new samples. Cumulatively, this new work more than doubled the size of the entire 14C database, which now comprised more than ninety dates (now calibrated) and was neatly summarized in a special edition of the Costa Rican journal Vínculos dedicated to “Taller sobre el futuro de las investigaciones arqueológicas y etnohistóricas en Gran Nicoya” (Vázquez Leiva et al. 1994:248–­250, table 2).9 The Vínculos volume largely comprised the proceedings of a 1993 archaeological conference in Guanacaste (the “Cuajiniquil workshop”) that introduced a revamped chronological sequence (Vázquez Leiva et al. 1994:248, table 1) to replace the Zoned Bichrome–­Early / Middle / Late Polychrome period system (table 4.3). Prior to this conference, a series of earlier meetings on Greater Nicoyan ceramics held in Denver (1982 and 1983), San José (1984), and Washington, DC (1985), and involving many of the leading archaeologists working in the area had already led to modifications in the Coe-­Baudez sequence (Lange 1990:1–­2). Following these meetings, the beginnings of the Early Polychrome and Late Creamer 1983, 1986); the Bay of Culebra (e.g., Abel-­Vidor 1980; Benson 1981; Lange, Ryder, and Accola 1986; Moreau 1983, 1984) and the Tempisque Valley (e.g., Day 1984).

9. An earlier (and somewhat more detailed) summary of about three dozen dates from Greater Nicoya had previously appeared in Lange and Stone (1984:383–­384). Greater Nicoya scholars generally treat the Vínculos 14C database as authoritative, though a small number of 14C dates collected before its publication (e.g., dates reported in Gorin 1990:255–­259) are excluded.

76

S T EI N B R EN N ER A N D M CCA F F ERT Y

Polychrome periods had both been pushed later, respectively, to ad 500 and ad 1350. However, after the Cuajiniquil workshop the beginning of the Early Polychrome period (now known as the Bagaces period) reverted to ad 300 (based primarily on a study of ceramic styles, radiocarbon dates, settlement patterns, and mortuary customs by Costa Rican archaeologists [Guerrero Miranda, Solís Del Vecchio, and Vázquez Leiva 1994]), while the beginning of the Late Polychrome period (now the Ometepe period) remained at ad 1350. This new chronological sequence remains in widespread use. Significantly, the suggested later date of ad 1350 for the beginning of the Ometepe period appears to have owed to the presence of several rather late (i.e., post-­a d 1200) radiocarbon dates in the 14C database associated with diagnostic Middle Polychrome (now Sapoá) period ceramics—­in other words, dates that suggested that the Middle Polychrome / Sapoá period actually endured for longer than previously believed10—­rather than to any new evidence supporting the idea that the ad 1200 date previously assigned to the beginning of the Late Polychrome period might actually be too early (e.g., a block of calibrated radiocarbon dates that clearly indicated that diagnostic “Late Polychrome / Ometepe-­period” ceramics could be more strongly associated with post-­a d 1350 contexts than with ad 1200–­1350 contexts). In fact, the situation was almost exactly opposite with respect to new evidence for “Ometepe-­period” diagnostic ceramics, which had never been particularly well associated with post-­a d 1200 contexts at any time, as indicated in the preceding discussion. In spite of the fact that the 14C database had doubled in the 1980s, only a half-­dozen distinct Ometepe-­period dates are listed in the Vínculos database (table 4.4), and only one of these dates—­a date from the Arenal region (Tx-­5079) that was not associated with any diagnostic ceramics from Greater Nicoya (Hoopes 1987:560)—­could be considered “new,” since the other late dates listed in the database had been long available. Almost all of these other Ometepe-­period dates are either problematic or cannot be firmly associated with “Ometepe-­period” diagnostic ceramics. These include the aforementioned problematic La Bocana dates and a date from a possible salt-­making station on 10. W hile not every post-­a d 1200 date assigned to the Sapoá period in the Vínculos 14C database (Vázquez Leiva et al. 1994:248–­250, table 2) can be clearly associated with Sapoá-­period diagnostic ceramics, most can. Post-­a d 1200 radiocarbon dates that are clearly associated with Sapoá-­period diagnostic ceramics include at least three dates from Guanacaste (P-­2180, P-2175, and P-2171; Sweeney 1975:380), two dates from Ometepe Island (Hv-­2690, Hv-­2691; Haberland 1978:405), and one date from the Arenal region, a highland area of Costa Rica adjacent to but somewhat outside the boundaries of Greater Nicoya as traditionally defined (Tx-­5077; Hoopes 1987:557). A C ritical R eevaluation of Pacific N icaragua’ s L ate P eriod C hronology

77

T able 4.4. Dates indicated to be “Ometepe-­period” in the Vínculos 14C database Sample

Site

GsY-­98(2)

La Bocana

Tx-­5079

C Age (bp) Issues and Comments

14

Armadillo

570 ± 30

No diagnostic ceramics from Greater Nicoya.

GsY-­89

La Bocana

516 ± 61

Sample represents an “average” of Gsy-­98 (1 and 2).

GsY-­98(1)

La Bocana

478 ± 85

Y-­814

Miramar

220 ± 100

Hv-­2692

San Lázaro

554 ± 85

505 ± 30

Diagnostic ceramics found in surface collection only.

Only sample actually associated with diagnostic “Late Polychrome period” ceramics (i.e., Madeira Polychrome). Diagnostic ceramics found in surface collection only. Salt-­making site; no diagnostic ceramics.

Source: Vázquez Leiva et al. (1994:248–­250, table 2). Note: Radiocarbon dates presented here are uncalibrated. All sites except San Lázaro are located in NW Costa Rica.

Costa Rica’s Santa Elena peninsula (Y-­814), where no diagnostic decorated ceramics of any type were found (Stuiver and Deevey 1961:132). In fact, only one designated Ometepe-­period 14C date in the Vínculos database can be definitively associated with a ceramic type (Madeira Polychrome) generally considered to be diagnostic of the period: i.e., a sample from the San Lázaro site on Ometepe Island (Hv-­2692; Haberland 1978:405, 1992:109–­110) that was one of the original four 14C dates used by Baudez and Coe to first establish the Late Polychrome period. But what of the other dates previously associated with the Late Polychrome period? Ironically, the three dates used by Sweeney to define the “Late Polychrome period” in her dissertation—­dates that were clearly associated with “Late Polychrome” diagnostic ceramics and which, recall, included one of the other dates originally used to define the Late Polychrome period—­were not listed in the Vínculos database as belonging to the final precontact (Ometepe) period but were rather reassigned to the preceding Sapoá period, presumably because these dates, even following calibration, were now considered to be “too old” to be associated with a period beginning in ad 1350!11 Inadvertently, it would appear, the compilers of the 14C database had almost completely separated the supposed ceramic evidence of the late period from the period itself—­a separation that, we believe, had actually been a long time coming. And yet this separation went unrecognized for at least another decade, as archaeologists in the 1990s and into the twenty-­first 11. Wyckoff ’s Late Polychrome period date from San Francisco (also associated with “Late Polychrome” diagnostic ceramics) was also reassigned to the Sapoá period in the database.

78

S T EI N B R EN N ER A N D M CCA F F ERT Y

century (including the present authors at the beginning of our own work in Nicaragua) paradoxically continued to treat these ceramics as being diagnostic of the “final” period in the Greater Nicoya sequence (and what is more, as being markers of an even later period than previously believed). For better or worse, this practice has remained standard in Greater Nicoya archaeology, despite the fact that more recent radiocarbon dates from Pacific Nicaragua published after the Vínculos database—­ including roughly two dozen Sapoá-­and Ometepe-­period dates reported in the last twenty years (e.g., G. McCafferty 2008; G. McCafferty and Steinbrenner 2005a; Niemel 2003; Salgado González 1996; Steinbrenner 2002, 2010)—­have generally indicated that Sweeney’s earlier dates were actually representative of contexts containing “Ometepe-­period” diagnostic ceramics and now provide extremely compelling evidence for associating most “Ometepe-­period” diagnostics with the latter half of the earlier Sapoá period rather than with post-­a d 1350 proveniences.12 This more recent Nicaraguan evidence will be discussed in detail in the following section.13 12. Why the practice of treating certain ceramics as Ometepe-­period diagnostics in spite of mounting evidence to the contrary has endured so long is a question that seems unlikely to have a simple answer. The aforementioned tendency to endow the established chronological sequence with an “aura of sanctity” and treat it as a fait accompli has probably been an important contributing factor, and the fact that “Ometepe-­period” diagnostics are often found in the uppermost strata in archaeological contexts has also likely influenced perceptions of their chronological position. Furthermore, we should observe that most projects prior to the University of Calgary projects in the twenty-­ first century produced only small numbers of radiocarbon dates, and that since 14C dates represent probable ranges of dates it is unlikely that one or two “outlier” dates for “Ometepe-­period” material closer to the earlier end of an expected range would have suggested a serious problem with the established chronological sequence for most archaeologists. In fact, we would go so far as to say that it would not be particularly good archaeological practice for an archaeologist to call for a revamping of the established sequence based solely on the evidence provided by one or two seemingly divergent dates, as opposed to numerous dates.

13. Although archaeological research has also continued in northwestern Costa Rica since the publication of the Vínculos database—­e.g., see chapters 8 and 16 in this volume as well as the summary of Solís Del Vecchio’s (2017) work at the end of the next section in this current chapter—­and although research elsewhere in Costa Rica (e.g., Corrales 2000) has added considerably to the 14C database for regions beyond Greater Nicoya—­to the best of the authors’ knowledge none of this work has produced additional 14C dates that significantly challenge our conclusions concerning the Sapoá-­ period provenience of supposed “Ometepe-­period” diagnostic ceramics. In light of this, narrowing our primary focus for the remainder of this chapter to the Pacific Nicaragua 14 C database seems appropriate.

A C ritical R eevaluation of Pacific N icaragua’ s L ate P eriod C hronology

79

Before continuing, it should be noted that at least one other 14C date from the Ometepe period collected during the 1980s but excluded by oversight from the Vínculos database can be associated, like the San Lázaro date, with the supposed diagnostic ceramics of the period. This sample (GIF-­7228), reported by Franck Gorin (1990:206–­216, 257–­259) from the La Pachona site in Chontales on the eastern coast of Lake Nicaragua (table 4.2), ascribes a potentially very late date (ad 1406–­1635, calibrated, 2σ) to a context that included significant quantities of “Ometepe-­period” diagnostics (including Ometepe Red Slipped-­ Incised and Vallejo and Madeira polychromes) but that also included considerable quantities of Papagayo Polychrome, generally viewed as being characteristic of the Sapoá period.14 While this single date would seem to support the standard perception that “Ometepe-­period” diagnostic ceramics actually are characteristic of the Ometepe period, it is worth noting that while Gorin (1990:269) himself viewed most supposed “Ometepe-­period” polychromes (e.g., Vallejo, Madeira, and Luna) as being typical of a rather “late” phase (i.e., the Monota phase, ad 1200–­1500) in Chontales that was more or less contemporary with the Late Polychrome / Ometepe period as it was understood elsewhere, he was also unique inasmuch as he did not see these ceramics as being diagnostic of the final phase in his own precontact chronological sequence. This view is actually consistent with our own view that supposed “Ometepe-­ period” ceramics are probably not diagnostic of the final period in the broader chronological sequence. Instead, Gorin (1992:33) postulated that an even later phase characterized mostly by unimpressive monochrome and bichrome types—­that is, the Cuapa phase, ad 1400–­1600—­followed his Monota phase and suggested that the appearance of the “Cuapa Complex” at several sites in Chontales pointed to a late invasion that provoked a local retraction of Greater Nicoya.15 Gorin speculated (based on little actual evidence) that this “invasion” might have come from the little-­studied area of the central cordillera to the north. However, if his single La Pachona date is actually too late (i.e., an outlier, as we are inclined to believe) and his “Ometepe-­period” 14. Gorin (1990:259) reports eight additional 14C dates in his dissertation, though he rejects four as being contaminated or intrusive. Three of the four “good” dates are associated with the Bagaces period; the other date is from early in the Sapoá period and is associated with Papagayo Polychrome but not with “Ometepe-­period” diagnostic ceramics (1990:201). 15. Gorin (1992:33) provides no 14C dates for the Cuapa phase, dating it instead by the association of Cuapa Complex ceramics with colonial ceramics and by the phase’s stratigraphic relationship to the Monota phase, which was itself dated primarily by the single La Pachona date.

80

S T EI N B R EN N ER A N D M CCA F F ERT Y

diagnostic polychromes are instead contemporary with the much earlier time frame suggested by Sweeney’s dates, then it would follow that the intrusive monochrome/bichrome complex that replaced these polychromes might also be earlier than Gorin believed and represent a true Ometepe-­period ceramic complex. We will return to this point later. OMETEPE-­P ERIOD DIAGNOSTICS IN SAPOÁ-­P ERIOD NICARAGUA

In the decade following the publication of the Vínculos database, two dissertations based on archaeological surveys of adjacent departments of Pacific Nicaragua added almost a dozen new dates to the Greater Nicoya 14C database. Silvia Salgado González’s survey of the department of Granada (Salgado González 1996:437) reported seven new dates from the site of Ayala (Ni-­Gr-­2), including five dates associated with the Bagaces period and two dates (Beta-­ 66956 and Beta-­66957) associated with “Ometepe-­period” diagnostics. The most likely calibrated 2σ ranges for these latter dates (table 4.2) fall comfortably in the latter half of the Sapoá period, between ad 1100 and ad 1300. Using the same survey methodology as Salgado, Karen Niemel’s survey of the department of Rivas (Niemel 2003:350; see chapter 7, this volume, for a summary) reported an additional Sapoá-­period date (Beta-­140586; table 4.2) from Santa Isabel (Ni-­Ri-­44) associated with “Ometepe-­period” diagnostics, in addition to three new Bagaces-­period dates from a different site. All three of Salgado’s and Niemel’s new dates associated with “Ometepe-­period” diagnostics are generally consistent with Sweeney’s dates for “Late Polychrome-­ period” ceramics.16 Salgado and Niemel’s dates for “Ometepe-­period” diagnostics are also consistent with numerous dates for similar ceramics reported by two major University of Calgary archaeological projects directed by Geoffrey McCafferty starting in 2000 and focusing on sites in the same departments of Rivas and Granada surveyed by Niemel and Salgado. These two projects, the first (2000–­2005) focusing on Santa Isabel (Debert 2005; G. McCafferty 2008; S. McCafferty 16. These relatively “early” dates help to explain why the local phases to which these two scholars assigned their ceramics—­Salgado González’s (1996:248–­251) Xalteva phase (ad 1150–­1522) and Niemel’s (2003:154) Las Lajas phase (ad 1200–­1522)—­have date ranges that are more akin to the dates traditionally assigned to the earlier-­starting Late Polychrome period than to the dates assigned to the later-­starting Ometepe period. Note that Niemel’s Las Lajas phase actually collapses two phases described by Healy (1980) for the Rivas area (see following discussion). A C ritical R eevaluation of Pacific N icaragua’ s L ate P eriod C hronology

81

and G. McCafferty 2008; G. McCafferty and S. McCafferty 2009, 2011; G. McCafferty and Steinbrenner 2005a; Steinbrenner 2002, 2010; see also chapters 6, 9, and 14, this volume) and the second (2008–­2010) on the sites of Tepetate (Ni-­Gr-­10) and El Rayo (Ni-­Gr-­39) in Granada (Dennett 2016; Manion 2016; G. McCafferty and Dennett 2013; Wilke 2012; see also chapters 6, 10, 11, 14, and 15, this volume), have collectively added twenty-­six new dates (mostly from the Sapoá period) to the Greater Nicoya 14C database (table 4.5, figure 4.3). These 14 C dates now comprise the bulk of the Nicaraguan dates in the total database. The Santa Isabel site, first visited by the University of Calgary in 2000, was identified by Niemel’s survey as the largest archaeological site in the department of Rivas and was originally presumed to have been occupied during the Ometepe period until Spanish contact in ad 1522 based on the abundant evidence of ceramic types that had been associated with this period in previous work. However, a dozen dates collected during the 2003 and 2004 field seasons provided consistent—­and quite unexpected—­evidence that this site actually predated the Ometepe period by several hundred years, spurring our initial investigation (G. McCafferty and Steinbrenner 2005a) into potential problems with the accepted chronological sequence. All twelve dates, which clustered between ad 890 and 1280 (calibrated, 2σ; G. McCafferty and Steinbrenner 2005a:138),17 were recovered from stratified contexts that were either directly associated with “Ometepe-­period” diagnostics (including Vallejo and Madeira polychromes, Castillo Engraved, and Ometepe Red Slipped-­Incised) or were positioned stratigraphically above contexts containing these same diagnostics. No contexts with Ometepe-­period dates (i.e., postdating ad 1350) were identified, suggesting an occupation of the site primarily limited to the Sapoá period, circa ad 800 to 1250/1300. Five additional Sapoá-­period dates recovered during the 2005 field season (G. McCafferty 2008; Steinbrenner 2010) reaffirmed the site’s primary occupation during this period, though most of these dates were not recovered from contexts containing diagnostic ceramics. The second University of Calgary project focusing on sites in Granada produced nine dates. The first field season in 2008 investigated Tepetate, a large site adjacent to the modern city of Granada that has been mostly destroyed by urban development and that had been previously identified with the contact-­ era town of Xalteva (see Dennett 2016:143ff; Salgado González 1996:358). The site produced incised whiteware ceramics similar to those found at Santa 17. Interestingly, a recent recalibration of these originally reported dates now pushes the earlier end of this range back by more than a century, from ad 890 to ad 780 (see table 4.5).

82

S T EI N B R EN N ER A N D M CCA F F ERT Y

Figure 4.3. 14C dates reported by University of Calgary archaeological projects in Rivas and Granada, Pacific Nicaragua, 2000–­2012.

Operation

Unit

Level/Context

2005

Beta-­217128

1

2/10

Beta-­196658

Beta-­196659

2/10

5

Beta-­196660

Beta-­196661

Locus 2

2005

Beta-­217130

2005

2005

Beta-­217127

Beta-­217129

E

4

C

Beta-­196656

Beta-­196655

Beta-­196654

Locus 1

S70E65

S62E52

S63E51.5

S60E41

N21E16

N21E13

N21E13

N21E13

N30E10

N21E8

N20E30

9 (hearth)

F43

7 (hearth)

4

930 ± 60

1010 ± 70

970 ± 60

820 ± 50

1180 ± 70

860 ± 70

F18 (175–­200 cm)

7 (122–­142 cm)

1090 ± 60

1010 ± 40

980 ± 50

870 ± 60

920 ± 50

C Age (bp)

14

12 (208–­216 cm)

F20 (202–­208 cm)

10

3

9

Santa Isabel (NI-­R I-­4 4) (All contexts are Sapoá period)

Sample #

−27.4 o/oo

−25.8 o/oo

−27.2 o/oo

−27.2 o/oo

−24.2 o/oo

−24.6 o/oo

−24.9 o/oo

−24.2 o/oo

−26.8 o/oo

−25.7 o/oo

−25.9 o/oo

C/12C Ratio

13

continued on next page

996 (95.4%) 1222

1198 (0.4%) 1204

886 (95.0%) 1191

972 (95.4%) 1212

1150 (87.7%) 1280

1124 (1.5%) 1136

1048 (6.1%) 1084

686 (95.4%) 986

1032 (95.4%) 1268

774 (95.4%) 1030

1065 (25.2%) 1154

966 (68.3%) 1059

903 (1.9%) 918

1176 (0.5%) 1182

972 (94.9%) 1170

1034 (95.4%) 1260

1022 (95.4%) 1214

2σ Interval (ad )

T able 4.5. 14C dates reported by University of Calgary archaeological projects in Rivas and Granada, Pacific Nicaragua, 2000–­2012

9

2

Beta-­196664

Beta-­196665

7

2005

Beta-­217131

N503E480

N142W123

S10E50

S82W123

S73E68

S73E61

S72E60

Unit

4

4

N504E480

VII

XV

Feature 2

Beta-­265450

Beta-­265463

1

1

(AMS sample)

N505E499

Level XII

Level VI

1230 ± 40

1220 ± 40

1360 ± 40

1430 ± 40

910 ± 50

900 ± 60

1020 ± 70

F3 (80–­90 cm)

4 (urn burial)

860 ± 60

990 ± 60

940 ± 80

C Age (bp)

14

7

6

4

Level/Context

Locus 2, Area 1 (transitional Bagaces-­Sapoá–­period contexts)

Beta-­265462

Beta-­265454

Locus 2, Area 2 (Bagaces-­period contexts)

El Rayo (Ni-­G r-­3 9)

Locus 7

Beta-­196657

Locus 5

Locus 4

9

9

Operation

Beta-­196663

Beta-­196662

Sample #

T able 4.5—continued

−24.8 o/oo

−25.0 o/oo

−25.8 o/oo

−10.5 o/oo

−26.3 o/oo

−24.8 o/oo

−25.5 o/oo

−26.7 o/oo

−25.2 o/oo

−25.7 o/oo

C/12C Ratio

13

continued on next page

683 (95.4%) 886

684 (95.4%) 892

742 (8.1%) 766

606 (87.3%) 716

558 (95.4%) 662

1024 (95.4%) 1218

1022 (95.4%) 1248

1198 (0.2%) 1202

875 (94.8%) 1190

780 (0.4%) 788

1038 (95.4%) 1264

948 (92.7%) 1185

900 (2.7%) 922

969 (95.4%) 1260

2σ Interval (ad)

1

Operation

N504E499

Unit

3

3

N485E460

N485E460 Level X

Level IX

Level IX

Level/Context

Locale 2

Beta-­257988

S144E244

N23E27

Level 4

Level 4 930 + 40

820 + 40

1030 ± 50

960 ± 40

1220 ± 60

C Age (bp)

14

−24.3 o/oo

−24.7 o/oo

−26.0 o/oo

−25.3 o/oo

−24.9 o/oo

C/12C Ratio

13

1022 (95.4%) 1189

1154 (94.4%) 1277

1058 (1.0%) 1072

1080 (14.8%) 1152

892 (80.6%) 1052

996 (95.4%) 1164

920 (5.9%) 961

668 (89.5%) 902

2σ Interval (ad)

Note: 2-­sigma (2σ) intervals calibrated using OxCal v4.3.2 (Bronk Ramsey 2017) and the IntCal13 atmospheric curve (Reimer et al. 2013).

Locale 1

Beta-­257985

Tepetate (NI-­G R-­1 0) (All contexts are Sapoá period)

Beta-­265453

Beta-­265459

Locus 2, Area 4 (Sapoá-­period contexts)

Beta-­265451

Sample #

T able 4.5—continued

Isabel (albeit more poorly preserved) that suggested a Sapoá-­period occupation contemporary with that of the Rivas site. Two 14C dates clustering between ad 1022 and 1277 (calibrated, 2σ) would seem to support this interpretation, though these dates unfortunately could not be directly associated with any diagnostic ceramics. Fieldwork in 2009 and 2010 focused on El Rayo, a secondary center on the Asese Peninsula south of the city of Granada characterized by much better preservation of archaeological material than that of Tepetate. El Rayo produced seven new 14C dates and, unlike Santa Isabel and Tepetate, contained a well-­defined Bagaces-­period component underlying Sapoá-­period deposits. Three of the earliest dates from this site were associated exclusively with Bagaces-­period ceramics. These included two dates (Beta-­265454, Beta-­265462) predating ad 700 (calibrated, 2σ) and a third date (Beta-­265463) with a calibrated (2σ) range of ad 684–­892 that spans the transition from the Bagaces to the Sapoá period circa ad 800. This date shares an intercept of ad 780 with two other dates (Beta-­265450, Beta-­265451, discussed in more detail below) associated with diagnostic ceramics postdating the Bagaces period, suggesting that the ad 800 date conventionally assigned to the end of the Bagaces period is appropriate. The two latest dates from the site (Beta-­265459, Beta-­ 265453), which derived from a context containing no underlying Bagaces component, were consistent with the Santa Isabel evidence, ranging from ad 892 to ad 1164 (calibrated, 2σ) and being associated with a similar assortment of “Ometepe-­period” diagnostics that included Vallejo and Madeira polychromes and Castillo Engraved. The Beta-­265450 and Beta-­265451 dates, which range from ad 668 to ad 961 (calibrated, 2σ), are provocative inasmuch as they represent the earliest dates securely associated with “Ometepe-­period” diagnostics anywhere in Pacific Nicaragua (and possibly all of Greater Nicoya as well) and were derived from contexts in which late-­period types such as Madeira Polychrome and Castillo Engraved were associated with Bagaces-­period diagnostic ceramics before ultimately supplanting the earlier types. Hopefully, future work will clarify whether these dates represent simple outliers, disturbed contexts, or, alternatively, if they reflect a precocious appearance of “Ometepe-­period” diagnostics at El Rayo during the transitional phase following the end of the Bagaces period.18 While the latter scenario remains an intriguing possibility—­particularly 18. One of these samples (Beta-­265451) is only tenuously associated with “Ometepe-­ period” diagnostics via two Banda polychrome sherds found in a deep stratigraphic context (Level IX) dominated mostly by Bagaces-­period material. The other sample (Beta-­265450)—­which derived from a stratigraphic context in which numerous A C ritical R eevaluation of Pacific N icaragua’ s L ate P eriod C hronology

87

since it would seem to be out of step with most of the other data pointing to the appearance of most “Ometepe-­period” diagnostics in the late Sapoá period—­it clearly demands further investigation; as this chapter has demonstrated, broad reinterpretations of chronology based on a handful of dates have rarely served the best interests of Greater Nicoyan archaeology. To summarize the implications of both University of Calgary projects with respect to the dating of “Ometepe-­period” diagnostic ceramics: even discounting those dates from Santa Isabel, Tepetate, and El Rayo deriving from contexts that are not specifically associated with “Ometepe-­period” diagnostic ceramics, there still remain a total of eighteen dates from Santa Isabel and El Rayo that clearly associate “Ometepe-­ period” diagnostics with Sapoá-­period contexts. Almost all of these dates (excepting the pair of El Rayo samples discussed immediately above) more specifically link these diagnostics to the latter half of the Sapoá period. These dates combine with other Pacific Nicaragua dates associated with “Ometepe-­period” diagnostics previously reported by Wyckoff, Salgado, and Niemel for a total of twenty-­ two dates associating “Ometepe-­period” diagnostic ceramics with the earlier Sapoá period that far exceeds the meager two dates reported by Haberland and Gorin for the same area associating these diagnostics with post-­a d 1350 proveniences. In short, Pacific Nicaragua’s 14C database provides overwhelming evidence for associating most “Ometepe-­period” diagnostic ceramics in Pacific Nicaragua with the late Sapoá period. Furthermore, the fact that none of these dates (nor the dates reported earlier by Salgado and Niemel) range past ad 1300 strongly suggests that this date, rather than ad 1350, might represent a more appropriate date for the Sapoá-­Ometepe period transition in Pacific Nicaragua. These two findings, which are consistent with the evidence from northwestern Costa Rica discussed above, obviously demands that the chronology of the southern half of Greater Nicoya—­built as it is on many of these same types—­must also be revisited. And indeed, recent work reported by Felipe Solís (2017) of Costa Rica’s National Museum makes it abundantly clear that the association of “Ometepe-­period” diagnostic ceramics with contexts predating ad 1350 is not a phenomenon that is restricted to Pacific Nicaragua. Solís provides sixteen radiocarbon dates (all with intercepts of ad 990 or later) obtained from six different sites around the Bay of Culebra excavated sherds from Sapoá-­period serving wares (including Papagayo, Pataky, and Madeira polychromes and Castillo Engraved) outnumber Bagaces-­period diagnostic sherds by a 2:1 ratio—­provides better evidence for the possible early appearances of types like Madeira and Castillo Engraved at El Rayo.

88

S T EI N B R EN N ER A N D M CCA F F ERT Y

between the 1990s and the early twenty-­first century that consistently link “Ometepe-­period” diagnostics to late Sapoá-­period contexts—­and in all but three cases, with contexts predating ad 1300.19 Additionally, Solís reports that “Ometepe-­period” diagnostics can also be associated with other late Sapoá-­ period contexts elsewhere in Guanacaste beyond the previously discussed site of Chahuite Escondido, including the sites of Huerta del  Aguacate on Tamarindo Bay and La Ceiba in the Tempisque River Valley. Clearly, the pattern revealed in the archaeological record of Guanacaste parallels the pattern revealed in our own work in Pacific Nicaragua. THE TRUE OMETEPE-­P ERIOD CERAMIC COMPLEX

The wealth of radiocarbon evidence that we have presented here in support of our new hypothesis—­that is, the “Ometepe-­period” diagnostic ceramics of Pacific Nicaragua (and, most likely, the rest of Greater Nicoya as well) were actually being produced several centuries earlier than is widely believed—­is difficult to dismiss. In fact, it is hard to imagine what might constitute more compelling archaeological evidence for this hypothesis, short of indisputable proof of a distinctive ceramic complex postdating the late Sapoá period and supplanting, in stratified contexts, the ceramic types conventionally treated as being diagnostic of the last phase of Pacific Nicaragua’s precontact history. The seeming absence of an identifiable ceramic complex in the archaeological record of Nicaragua that can be definitively dated to the Ometepe period poses an obvious problem for our hypothesis, though not one that provides an excuse for simply ignoring the radiocarbon evidence presented in this chapter. It should be recognized that this apparent absence has likely contributed to the widely held view that “Ometepe-­period” diagnostics must have endured until European contact,20 even though this interpretation has never been supported by more than a handful of disparate dates (several problematic, as shown here) and in spite of the fact that most dates suggest an earlier appearance (as also shown here) that seems hard to reconcile with the idea that these ceramics were still being produced, almost unchanged, five or six centuries later. However, while the general dearth of well-­dated archaeological contexts 19. Eleven of Solís’s new dates are from the Jícaro site (discussed in more detail in chapter 16, this volume); the remainder (single dates from each site) are from El Conchal, Hunter Robinson, Llano La Molonga, Manzanillo, and Nacascolo.

20. For example, Salgado González (1996:251) extended her Xalteva phase until the Spanish conquest because she found “no evidence of a differentiated ceramic complex later than Xalteva” at the sites she excavated in Granada.

A C ritical R eevaluation of Pacific N icaragua’ s L ate P eriod C hronology

89

from the Ometepe period means that we cannot presently rule out the possibility that the various diagnostic types that first appeared in the Sapoá period did indeed endure into the subsequent period and became its characteristic, “typical” types, we consider this unlikely and would alternatively call attention to several late-­appearing ceramic types—­all poorly understood at present—­which suggest that distinctive Ometepe-­period ceramic complexes probably did exist. These types include • Luna Polychrome, a type produced on Ometepe Island (Bransford 1881:21; Knowlton 1992,1996; Lothrop 1926:194–­213) that has been previously associated with European trade wares in archaeological contexts (Bovallius 1886:9). Although potentially first appearing in late-­Sapoá contexts, Luna is often found in upper strata and surface collections at sites across Greater Nicoya. • El Menco Polychrome (Knowlton 1992, 1996; Steinbrenner 2010:616–­339), an easily-­confused-­but-­distinct cousin of Luna that was frequently found in Santa Isabel surface collections and burials. • Managua Polychrome, an atypical, easily recognizable polychrome type with clear analogues (Steinbrenner 2014, 2015) to late Postclassic Mesoamerican types found outside Greater Nicoya (cf. Babcock 2012; Haberland 1975; Henderson 1977; Henderson et al. 1979; Navarrete 1966; Rands and Smith 1965; Schortman and Urban 2011; Stone 1957; Strong, Kidder and Paul 1938; Wonderley 1981, 1986). While Managua Polychrome is best known from unprovenienced collections, it has been linked to colonial contexts around the Lake Managua Basin (Cornavaca 2003:113, 141–­146) and, more recently, in Nicoya (Camacho Mora et al. 2016; Camacho Mora and Vargas Madrigal 2017).21 • Various finely made monochrome types characterized by highly polished black surfaces, possibly associated with the widely traded and historically documented ceramics produced on Chira Island in the Gulf of Nicoya into colonial times (Abel-­Vidor 1980:161; Creamer 1986:212). • Gorin’s (1992:33) aforementioned “Cuapa Complex,” which supplanted “Ometepe-­period” diagnostics in Chontales and which was also associated with colonial-­era ceramics. 21. As discussed elsewhere by Steinbrenner (2014, 2015), Managua Polychrome bears an extraordinary resemblance to—­and might easily be mistaken for—­the ceramic types Nolasco Bichrome and Vagando Polychrome from Naco, Honduras (Wonderley 1986), as well as to Nimbalari Trichrome from Chiapa de Corzo, Mexico (Navarrete 1966). A variant of the type is sometimes referred to as “Managua Black-­on-­Red.”

90

S T EI N B R EN N ER A N D M CCA F F ERT Y

We are not the first scholars to suggest that the traditional “Ometepe-­ period” diagnostic ceramics of Greater Nicoya were eventually superseded by new types. For example, Coe, following his preliminary work at Chahuite Escondido, suggested that the Late Polychrome period might be divided into two phases: the earlier (La Cruz A) characterized by Vallejo Polychrome (which we see as being primarily a Sapoá type) and the later (La Cruz B) by Luna Polychrome and various fine monochromes (Baudez and Coe 1962:368). Healy (1980:305) likewise suggested two late phases for Rivas characterized by ceramic types similar to Coe’s: Las Lajas (ad 1200–­1350)—­ characterized by Vallejo, Madeira, and Ometepe Red Slipped-­ Incised—­ and Altagracia (ad 1350–­1522)—­characterized by Luna and black monochromes. Haberland (1986:381) also saw Luna as distinguishing a late phase on Ometepe Island beginning circa ad 1400. And, finally, Gorin’s work, as already noted, also proposed that the traditional “Ometepe-­period” diagnostics were replaced by new types in Chontales around ad 1400. While our own work obviously challenges the dating of the various phases identified in previous attempts to subdivide the Late Polychrome / Ometepe period, and while such attempts have often been rejected by previous researchers (e.g., Niemel, Román Lacayo, and Salgado González 1998:679; Sweeney 1975:398), in part owing to the difficulty of differentiating contexts containing “late” material from earlier contexts containing “Ometepe-­period” diagnostics,22 we believe that the evidence presented here for the late-­Sapoá provenience of “Ometepe-­period” types makes a strong case for reconsidering these arguments. So why is it so difficult to define archaeological contexts associated with the late prehistory of Pacific Nicaragua? While there is probably not a simple answer to this question, several contributing factors are worth considering. For sites that have been previously explored, it is possible that some locales once thought to be “late” (mostly based on misdated ceramic rather than ethnohistoric evidence) may actually have been largely abandoned by the end of the Sapoá period (possibly in response to a significant cultural change, such as the arrival of new migrants from Mesoamerica; see chapter 2, this volume) and therefore may never have featured significant Ometepe-­period components. 22. Niemel, Román Lacayo, and Salgado González (1998:679), for example, argued that Healy’s Altagracia phase should be absorbed into the Las Lajas phase because survey data from the adjacent departments of Rivas, Granada, and Masaya indicate that the phase is quite difficult to identify anywhere other than on Ometepe Island and that to accept the existence of such a phase is therefore tantamount to accepting that there were no pre-­Columbian settlements in Pacific Nicaragua for two centuries prior to Spanish contact. A C ritical R eevaluation of Pacific N icaragua’ s L ate P eriod C hronology

91

For example, we believe this to be the case for Santa Isabel, as previously noted. Unfortunately, while such sites therefore cannot really tell us much about the final stage of Nicaragua’s precontact history, their continued misidentification as Ometepe-­period sites will likely discourage the adoption of our new hypothesis that the Ometepe period remains poorly defined and will likewise discourage acceptance of the possibility that later precontact contexts—­what we might call “true” Ometepe-­period sites—­need to be identified and explored. An alternative possibility that must be considered for previously investigated sites located in urban areas (e.g., Tepetate) or in areas devoted to modern agriculture (e.g., most of the sites surveyed by Salgado and Niemel) is that some Ometepe-­period components may have simply been destroyed by plowing or construction, or lie buried beneath modern developments. With respect to this latter point, it is important to note the near absence of scientific archaeological investigation within modern urban areas that were also known primary centers of pre-­Hispanic occupation in Pacific Nicaragua at the end of the Ometepe period. These include the cities of Nandaime, Masaya/Nindirí, Granada (formerly Xalteva), Managua, León (formerly Subtiaba), Chinandega, El Viejo (formerly Tecoatega), and Rivas–­San Jorge (about 10 km south of Santa Isabel).23 With the exception of Managua (Lange 1995, 1996; see also chapter 5 in this volume) and our own limited work in Granada, these potentially valuable “sites” have been largely ignored because most major archaeological work in Pacific Nicaragua has opportunistically focused on more accessible “abandoned” sites in nonurban areas scattered across the countryside. Another factor that may hamper efforts to define true Ometepe-­period archaeological contexts is that some components from this period (particularly those associated with newly arrived groups as opposed to preestablished Sapoá-­era groups) might simply not be recognized for what they are (and therefore not be adequately explored) if the ceramics associated with these components are not perceived as clearly reflecting the “Mesoamerican” influence that appears to have played a critical role in the late prehistory of Pacific Nicaragua. The strong expectation that pottery linked to newly arrived Mesoamerican groups should also look Mesoamerican, combined with a comparative lack of interest in supposedly nondiagnostic utilitarian wares as opposed to decorated wares, has contributed to a general neglect of most

23. The city of Nicoya in Guanacaste provides another such site in Greater Nicoya’s southern sector.

92

S T EI N B R EN N ER A N D M CCA F F ERT Y

nondecorated ceramics in Pacific Nicaragua.24 But there is no logical reason to assume that the beautiful polychromes of the Sapoá period would “naturally” have been superseded by equally attractive decorated ceramics correlating with the arrival of a new migrant population in the Ometepe period. In fact, the ceramic complexes of the Ometepe period—­including the yet-­to-­be-­defined complex of the Nicarao, who likely arrived during this period (see chapter 2, this volume)—­may have been unremarkable compared to the complexes that preceded them or even largely indistinguishable from Indigenously produced utilitarian wares dating to the colonial period, which might be dismissed by archaeologists searching for precontact occupations (and perhaps especially in middens located in continuously occupied urban centers). Established preconceptions about Ometepe-­period ceramics likely explain why Gorin, for example, refused to link his seemingly unimpressive Cuapa Complex (largely comprising undecorated utilitarian wares) to a supposedly “Mesoamerican” Greater Nicoya, despite the fact that this complex appears to have included comals (Gorin 1990:477ff ), an especially Mesoamerican utilitarian form that was absent from our own pre-­Ometepe period sites in Rivas and Granada but that likely remained common in colonial-­era contexts (and could therefore easily contribute to possible confusion between pre-­and postcontact components).25 Whatever the explanation for the apparent lack of evidence for archaeological components that can be truly associated with the Ometepe period, we find it understandable that past researchers were not motivated to identify alternative Ometepe-­period ceramic complexes or define discrete Ometepe-­period components in archaeological contexts given that the accepted chronological sequence gave them no reason to suspect that there might be a need to do so—­that is, to suspect that the characteristic archaeological evidence of the 24. The preconception that the pottery of Mesoamerican groups should strongly reflect their origins largely explains why Vallejo Polychrome in particular has been linked to “Mexican” influence since at least the early twentieth century (cf. Lothrop 1926). Decorated in a style highly reminiscent of the codex books of Central Mexico and Oaxaca, Vallejo Polychrome is broadly recognized as representing the southernmost manifestation of the Postclassic Mixteca–­Puebla stylistic tradition (cf. Day 1984; Healy 1980; McCafferty and Steinbrenner 2005b). Vallejo is commonly treated as a marker of the Nicarao, one of the last Mesoamerican migrant groups to arrive in Pacific Nicaragua (see chapters 2 and 9); however, this assumption now seems very problematic, given the late Sapoá provenience of the type established here.

25. Gorin does not appear to have recognized that two of the types he associates with the Cuapa Complex, Miragua Common and Coronado Red, contained comals. Our identification of these forms is based on rim profile drawings in his dissertation (1990:477, 480). A C ritical R eevaluation of Pacific N icaragua’ s L ate P eriod C hronology

93

Ometepe period had not been sufficiently well established. However, as we have demonstrated in this chapter, there is now very good reason to question the supposed archaeological evidence of the Ometepe period and to embrace the alternative hypothesis that this primarily ceramic evidence instead represents a considerably older period in Pacific Nicaraguan prehistory. We recognize that this hypothesis represents a paradigm shift that raises many new questions and has potentially profound cultural implications that many of our colleagues may, at first, find difficult to accept. However, we remain convinced that making this shift represents the way forward to a better understanding of the archaeology of Pacific Nicaragua in particular, and Greater Nicoya in general. REFERENCES

Abel-­V idor, Suzanne. 1980. “The Historical Sources for the Greater Nicoya Archaeological Subarea.” Vínculos: Revista de Antropología del Museo Nacional de Costa Rica 6(1–­2):155–­176. Aguilar Piedra, Carlos H. 1984. “Introducción a la arqueología de la región del Volcán Arenal: Tefraestratigrafía y secuencia cultural.” Anales de la Academia de Geografía e Historia de Costa Rica 1979–­1982:55–­87. Babcock, Thomas F. 2012. Utatlán: The Constituted Community of the K’iche’ Maya of Q’umarkaj. Boulder: University Press of Colorado. Baker, Suzanne M., and Michael Smith. 1987. “Archaeological Reconnaissance on Isla Zapatera, 1986 Field Season.” Manuscript on file, Archivo Informes, Departamento de Arqueología, Dirección de Patrimonio Cultural (DPC), Instituto Nicaragüense de Cultura (INC), Managua. Baudez, Claude F. 1967. Recherches archéologiques dans la vallée du Tempisque, Guanacaste, Costa Rica, Travaux et mémoires 18. Paris: Institut des Hautes Études de l’Amérique Latine, Université de Paris. Baudez, Claude F., and Michael D. Coe. 1962. “Archaeological Sequences in Northwestern Costa Rica.” In Proceedings of the 34th International Congress of Americanists, 366–­373. Vienna: Verlag Ferdinand Berger. Benson, Elizabeth, ed. 1981. Between Continents / Between Seas: Precolumbian Art of Costa Rica. New York: Harry N. Abrams. Bovallius, Carl. 1886. Nicaraguan Antiquities. Stockholm: Swedish Society of Anthropology and Geography. Bransford, John F. 1881. Archaeological Researches in Nicaragua. Smithsonian Institution Contributions to Knowledge 383. Washington, DC: Smithsonian Institution. 94

S T EI N B R EN N ER A N D M CCA F F ERT Y

Bronk Ramsey, Cristopher. 2009. “Bayesian Analysis of Radiocarbon Dates.” Radiocarbon 51(1):337–­360. Bruhns, Karen Olsen. 1974. “Field Notes Taken on Zapatera Island. August 13–­­15, 1974.” Manuscript on file, Adán E. Treganza Anthropology Museum, Department of Anthropology, San Francisco State University. Camacho Mora, Fernando, and Geissel Vargas Madrigal. 2017. “Informe final evaluación arqueológica ‘Restauración estructural de la iglesia de San Blas de Nicoya’ (segunda temporada) Guanacaste, Nicoya.” Manuscript on file, Archivo Informes, Archivo de Investigación, Comisión Arqueológica Nacional and Departamento de Antropología e Historia, Museo Nacional de Costa Rica, San José, Costa Rica. Camacho Mora, Fernando, Geissel Vargas Madrigal, Josebec Ureña Pérez, and Zeidy Mora Aguilar. 2016. “Informe trabajos realizados en la evaluación arqueológica del proyecto ‘Restauración estructural de la iglesia de San Blas de Nicoya.’ ” Manuscript on file, Archivo Informes, Archivo de Investigación, Comisión Arqueo­ lógica Nacional and Departamento de Antropología e Historia, Museo Nacional de Costa Rica, San José. Coe, Michael D. 1962. “Costa Rican Archaeology and Mesoamerica.” Southwestern Journal of Anthropology 18(2):170–­183. Coe, Michael D., and Claude F. Baudez. 1961. “The Zoned Bichrome Period in Northwestern Costa Rica.” American Antiquity 26(4):505–­515. Cornavaca, Deborah Ellen Erdman. 2003. “Leon Viejo, Nicaragua: A Community of Contact.” PhD diss., Department of Anthropology, University of California, Los Angeles. Corrales Ulloa, Francisco. 2000. “An Evaluation of Long Term Cultural Change in Southern Central America: The Ceramic Record of the Diquís Archaeological Subregion, Southern Costa Rica.” PhD diss., University of Kansas. Creamer, Winifred. 1983. “Production and Exchange on Two Islands in the Gulf of Nicoya, Costa Rica, A.D. 1200–­1550.” PhD diss., Tulane University. Creamer, Winifred. 1986. “Archaeological Reconnaissance in the Gulf of Nicoya.” In Prehistoric Settlement Patterns in Costa Rica, edited by Frederick W. Lange and Lynette C. Norr, 207–­220. Journal of the Steward Anthropological Society 14 (1–­2 [1982–­1983]). Urbana: University of Illinois. Day, Jane S. 1984. “New Approaches in Stylistic Analysis: The Late Polychrome Period Ceramics from Hacienda Tempisque, Guanacaste Province, Costa Rica.” PhD diss., University of Colorado, Boulder. Debert, Jolene. 2005. “Raspadita: A New Lithic Tool Type from Santa Isabel, Nicaragua.” MA thesis, University of Manitoba. Dennett, Carrie L. 2016. “The Ceramic Economy of Pre-­Columbian Pacific Nicaragua (ad 1–­1250).” PhD diss., University of Calgary. A C ritical R eevaluation of Pacific N icaragua’ s L ate P eriod C hronology

95

Espinoza Pérez, Edgard, and Dominique Rigat. 1994. “Gran Nicoya y la región de Chontales, Nicaragua.” Vínculos: Revista de Antropología del Museo Nacional de Costa Rica 18–­19(1–­2 [1992–­1993]):139–­156. Gorin, Franck. 1990. “Archéologie de Chontales, Nicaragua.” PhD diss., Université de Paris I, Panthéon-­Sorbonne. Gorin, Franck. 1992. “Arqueología de Nicaragua.” Revista TRACE (Travaux et Recherches dans les Amériques du Centre) 21:22–­35. Guerrero Miranda, Juan Vicente, and Aida María Blanco Vargas. 1987. “La Ceiba: Un asentamiento policromo medio en el Valle del Tempisque con actividades funerarias (G-­60-­LC).” Licentiate’s thesis, Universidad de Costa Rica, San José. Guerrero Miranda, Juan Vicente, Ricardo Vázquez Leiva, and Federico Solano B. 1992. “Entierros secundarios y restos orgánicos de ca. 500 AC preservados en un área de inundación marina, Golfo de Nicoya, Costa Rica.” Vínculos: Revista de Antropología del Museo Nacional de Costa Rica 17 (1991):17–­51. Guerrero Miranda, Juan Vicente, Luis Felipe Solís Del Vecchio, and Ricardo Vázquez Leiva. 1994. “El período Bagaces (300–­800 d.C.) en la cronología arqueo­ lógica del noroeste de Costa Rica.” Vínculos: Revista de Antropología del Museo Nacional de Costa Rica 18–­19(1–­2 [1992–­1993]):91–­109. Haberland, Wolfgang. 1966. “Early Phases on Ometepe Island, Nicaragua.” In Proceedings of the 36th International Congress of Americanists 1, 399–­403. Seville: Editorial Católica Española. Haberland, Wolfgang. 1975. “Further Archaeological Evidence for the Nicarao and Pipil Migrations in Central America.” In Proceedings of the 41st International Congress of Americanists 1, 551–­559, Mexico City, September 2–­­7, 1974. Haberland, Wolfgang. 1978. “Lower Central America.” In Chronologies in New World Archaeology, edited by Royal Ervin Taylor and Clement Woodward Meighan, 395–­430. Studies in Archaeology. New York: Academic Press. Haberland, Wolfgang. 1986. “Settlement Patterns and Cultural History of Ometepe Island, Nicaragua: A Preliminary Sketch.” In Prehistoric Settlement Patterns in Costa Rica, edited by Frederick W. Lange and Lynette C. Norr, 369–­386. Journal of the Steward Anthropological Society 14 (1–­2 [1982–­1983]). Urbana: University of Illinois. Haberland, Wolfgang. 1992. “The Culture History of Ometepe Island: Preliminary Sketch (Survey and Excavations, 1962–­1963).” In The Archaeology of Pacific Nicaragua, edited by Frederick W. Lange, Payson D. Sheets, Aníbal Martínez, and Suzanne Abel-­Vidor, 63–­117. Albuquerque: University of New Mexico Press. Healy, Paul F. 1974. “Archaeological Survey of the Rivas Region, Nicaragua.” PhD diss., Harvard University, Cambridge, MA. 96

S T EI N B R EN N ER A N D M CCA F F ERT Y

Healy, Paul F. 1980. Archaeology of the Rivas Region, Nicaragua. Waterloo, ON: Wilfrid Laurier University Press. Henderson, John S. 1977. “The Valley de Naco: Ethnohistory and Archaeology in Northwestern Honduras.” Ethnohistory 24(4):363–­377. Henderson, John S., Ilene Sterns, Anthony Wonderley, and Patricia A. Urban. 1979. “Archaeological Investigations in the Valle de Naco, Northwestern Honduras: A Preliminary Report.” Journal of Field Archaeology 6(2):169–­192. Hoopes, John W. 1987. “Early Ceramics and the Origins of Village Life in Lower Central America.” PhD diss., Harvard University, Cambridge, MA. Hughes, Neil Cameron 1980. “Urn Burial in Prehistoric Nicaragua.” MA thesis, George Washington University, Washington, DC. Hurtado de Mendoza, Luis, and Guillermo E. Alvarado Induni. 1990. “Datos arqueo­ lógicos y vulcanológicos de la región del Volcán Miravalles, Costa Rica.” Vínculos: Revista de Antropología del Museo Nacional de Costa Rica 14(1–­2 [1988]):77–­89. Knowlton, Norma E. 1992. “Ancient Mortuary Ceramics from Nicaragua: A Study of Luna Polychrome.” MA thesis, Trent University. Knowlton, Norma E. 1996. “Luna Polychrome.” In Paths to Central American Pre­history, edited by Frederick W. Lange, 143–­176. Niwot: University Press of Colorado. Lange, Frederick W. 1971. “Culture History of the Sapoá River Valley, Costa Rica.” PhD diss., University of Wisconsin, Madison. Lange, Frederick W. 1980. “The Formative Zoned Bichrome Period in Northwestern Costa Rica (800 B.C. to A.D. 500), Based on Excavations at the Vidor Site, Bay of Culebra.” Vínculos: Revista de Antropología del Museo Nacional de Costa Rica 6(1–­2):33–­42. Lange, Frederick W. 1990. “Breve resumen de las conferencias sobre la cerámica de la Gran Nicoya.” Vínculos: Revista de Antropología del Museo Nacional de Costa Rica 13(1–­2 [1987]):1–­5. Lange, Frederick W., ed. 1995. Descubriendo las huellas de nuestros antepasados: El proyecto “Arqueología de la Zona Metropolitana de Managua.” Managua: Alcaldía de Managua. Lange, Frederick W., ed. 1996. Abundante cooperación vecinal: La segunda temporada del proyecto “Arqueología de la Zona Metropolitana de Managua.” Managua: Alcaldía de Managua. Lange, Frederick W., Suzanne Abel-­Vidor, Claude F. Baudez, Ronald L. Bishop, Winifred Creamer, Jane S. Day, Juan Vicente Guerrero Miranda, Paul F. Healy, Silvia Salgado González, Robert Stroessner, and Alice Tillet. 1984. “New Approaches to Greater Nicoya Ceramics.” In Recent Developments in Isthmian Archaeology: Advances in the Prehistory of Lower Central America: Proceedings of the 44th International A C ritical R eevaluation of Pacific N icaragua’ s L ate P eriod C hronology

97

Congress of Americanists, Manchester, 1982, edited by Frederick W. Lange, 199–­214. BAR International Series 212. Oxford: British Archaeological Reports.

Lange, Frederick W., Peter Ryder, and Richard M. Accola. 1986. “Bay of Culebra Survey.” In Prehistoric Settlement Patterns in Costa Rica, edited by Frederick W. Lange and Lynette C. Norr, 25–­36. Journal of the Steward Anthropological Society 14 (1–­2 [1982–­1983]). Urbana: University of Illinois.

Lange, Frederick W., Payson D. Sheets, Aníbal Martínez, and Suzanne Abel-­Vidor. 1992. The Archaeology of Pacific Nicaragua. Albuquerque: University of New Mexico Press. Lange, Frederick W., and Doris Z. Stone, eds. 1984. The Archaeology of Lower Central America. School of American Research Advanced Seminar Series. Albuquerque: University of New Mexico Press.

Lothrop, Samuel K. 1926. Pottery of Costa Rica and Nicaragua. 2 vols. Memoir 8. New York: Museum of the American Indian, Heye Foundation. Manion, Jessica. 2016. “Remembering the Ancestors: Mortuary Practices and Social Memory in Pacific Nicaragua.” MA thesis, University of Calgary.

McCafferty, Geoffrey G. 2008. “Domestic Practice in Postclassic Santa Isabel, Nicaragua.” Latin American Antiquity 19(1):64–­82.

McCafferty, Geoffrey G., and Carrie L. Dennett. 2013. “Ethnogenesis and Hybridity in Proto-­Historic Nicaragua.” Archaeological Review from Cambridge 28(1):191–­215.

McCafferty, Geoffrey G., and Sharisse D. McCafferty. 2009. “Crafting the Body Beautiful: Performing Social Identity at Santa Isabel, Nicaragua.” In Mesoamerican Figurines: Small-­Scale Indices of Large-­Scale Social Phenomena, edited by Christina T. Halperin, Katherine A. Faust, Rhonda Taube, and Aurore Giguet, 183–­204. Gainesville: University Press of Florida.

McCafferty, Geoffrey G., and Sharisse D. McCafferty. 2011. “Bling and Things: Ornamentation and Identity in Pacific Nicaragua.” In Identity Crisis: Archaeological Perspectives on Social Identity. Proceedings of the 42nd Annual Chamool Archaeology Conference, edited by Lindsey Amundsen-­Meyer, Nicole Engel, and Sean Pickering, 243–­252. Calgary: Chacmool Archaeology Association, University of Calgary.

McCafferty, Geoffrey G., and Larry Steinbrenner. 2005a. “Chronological Implications for Greater Nicoya from the Santa Isabel Project, Nicaragua.” Ancient Mesoamerica 16(1):131–­146.

McCafferty, Geoffrey G., and Larry Steinbrenner. 2005b. “The Meaning of Mixteca-­ Puebla Stylistic Tradition: The View from Nicaragua.” In Art for Archaeology’s Sake: Material Culture and Style across the Disciplines. Proceedings of the 33rd Annual Chacmool Archaeology Conference, edited by Andrea Waters-­Rist, Christine Cluney,

98

S T EI N B R EN N ER A N D M CCA F F ERT Y

Calla McNamee, and Larry Steinbrenner, 282–­292. Calgary: Chacmool Archae­ ology Association, University of Calgary.

McCafferty, Sharisse D., and Geoffrey G. McCafferty. 2008. “Spinning and Weaving Tools from Santa Isabel, Nicaragua.” Ancient Mesoamerica 19(1):143–­156. Moreau, Jean-­François. 1983. “L’adaptation maritime préhistorique au site Vidor, Costa Rica.” PhD diss., Université de Montréal. Moreau, Jean-­François. 1984. “Subsistance [sic] Changes on the Coast of the Greater Nicoya Region.” In Inter-­regional Ties in Costa Rican Prehistory: Papers Presented at a Symposium at Carnegie Museum of Natural History, Pittsburgh, April 27, 1983, edited by Esther R. Skirboll and Winifred Creamer, 129–­141. BAR International Series 226. Oxford: British Archaeological Reports. Navarrete, Carlos. 1966. The Chiapanec History and Culture. Edited by John Alden Mason. Papers of the New World Archaeological Foundation 21, Publication 16. Provo, UT: Brigham Young University Press. Navarro Genie, Rigoberto. 1993. “Labor del Departamento de Arqueología (1980–­85).” In 30 años de arqueología en Nicaragua, edited by Jorge Eduardo Arellano, 19–­21. Managua: Museo Nacional de Nicaragua, Instituto Nicaragüense de Cultura. Niemel, Karen S. 2003. “Social Change and Migration in the Rivas Region, Pacific Nicaragua (1000 bc ad 1522).” PhD diss., State University of New York, Buffalo. Niemel, Karen S., Manuel A. Román Lacayo, and Silvia Salgado González. 1998. “Las secuencias cerámicas de los periodos Sapoá (800–­1350 dC) y Ometepe (1350–­1522 dC) en el Pacífico Sur de Nicaragua.” In XI Simposio de Investigaciones Arqueológicas en Guatemala, 1997, edited by Juan Pedro Laporte and Héctor L. Escobedo, 677–­683. Guatemala City: Museo Nacional de Arqueología y Etnología, Instituto de Antología e Historia, Ministerio de Cultura y Deportes y Asociación Tikal. Norr, Lynette C. 1986. “Archaeological Site Survey and Burial Mound Excavations in the Río Naranjo-­Bijagua Valley.” In Prehistoric Settlement Patterns in Costa Rica, edited by Frederick W. Lange and Lynette C. Norr, 135–­156. Journal of the Steward Anthropological Society, 14 (1–­2 [1982–­1983]). Urbana: University of Illinois. Norweb, Albert Holden. 1964. “Ceramic Stratigraphy in Southwestern Nicaragua.” Proceedings of the 35th International Congress of Americanists 1, 551–­561. Mexico City, August 19–­­25, 1962. Rands, Robert L., and Robert E. Smith. 1965. “Pottery of the Guatemalan Highlands.” In Archaeology of Southern Mesoamerica, edited by Gordon R. Willey, 95–­145. Handbook of Middle American Indians 2, pt. 1, Robert Wauchope, general editor. Austin: University of Texas Press. Reimer, Paula J., et al. 2013. “IntCal13 and Marine13 Radiocarbon Age Calibration Curves 0–­50,000 Years cal bp.” Radiocarbon 55(4):1869–­1887. A C ritical R eevaluation of Pacific N icaragua’ s L ate P eriod C hronology

99

Rigat, Dominique. 1992. “Préhistoire au Nicaragua: Région Juigalpa, Département de Chontales.” PhD diss., Université de Paris I, Panthéon-­Sorbonne. Rigat, Dominique, and Rafael González Rivas. 1996. “Preliminary Research in Chontales and the Lake Managua Basin, Nicaragua.” In Paths to Central American Pre­ history, edited by Frederick W. Lange, 177–­190. Niwot: University Press of Colorado. Rigat, Dominique, and Franck Gorin. 1993. “ ‘Proyecto Chontales’: Informe final.” In 30 años de arqueología en Nicaragua, edited by Jorge Eduardo Arellano, 97–­102. Managua: Museo Nacional de Nicaragua, Instituto Nicaragüense de Cultura. Ryder, Peter. 1986. “Hacienda Mojica.” In Prehistoric Settlement Patterns in Costa Rica, edited by Frederick W. Lange and Lynette C. Norr, 105–­120. Journal of the Steward Anthropological Society 14 (1–­2 [1982–­1983]). Urbana: University of Illinois. Salgado González, Silvia. 1996. “Social Change in a Region of Granada, Pacific Nicaragua (1000 B.C.–­1522 A.D.).” PhD diss., State University of New York, Albany. Schortman, Edward M., and Patricia A. Urban. 2011. Networks of Power: Political Relations in the Late Postclassic Naco Valley, Honduras. Edited by Davíd Carrasco and Eduardo Matos Moctezuma, Mesoamerican Worlds: From the Olmecs to the Danzantes. Boulder: University Press of Colorado. Sheets, Payson. 1986. “The Proyecto Prehistórico Arenal: An Introduction.” Vínculos: Revista de Antropología del Museo Nacional de Costa Rica 10 (1984):17–­29. Solís Del Vecchio, Luis Felipe. 2017. “Bahía Culebra en el noroeste de Costa Rica: ¿Tuvo ocupación prehispánica durante el Período Ometepe (1350–­1550 d.C.)?” Paper presented at the XI Congreso de la Red Centroamericana de Antropología, University of Costa Rica, San José, February 27–­­March 3. Steinbrenner, Larry. 2002. “Ethnicity and Ceramics in Rivas, Nicaragua, ad 800–­1550.” MA thesis, University of Calgary. Steinbrenner, Larry. 2010. “Potting Traditions and Cultural Continuity in Pacific Nicaragua, ad 800–­1350.” PhD diss., University of Calgary. Steinbrenner, Larry. 2014. “The Mystery of Managua Polychrome.” Paper presented at the 79th Annual Meeting of the Society for American Archaeology, Austin, Texas, April 23–­­27. Steinbrenner, Larry. 2015. “The Mystery of Managua Polychrome Part II.” Paper presented at the 80th Annual Meeting of the Society for American Archaeology, San Francisco, California, April 15–­­19. Stone, Doris Z. 1957. The Archaeology of Central and Southern Honduras. Papers of the Peabody Museum of Archaeology and Ethnology, vol. 49, no. 3. Cambridge: Peabody Museum of Archaeology and Ethnology, Harvard University. Strong, William Duncan, Alfred V. Kidder, and Anthony Joseph Drexel Paul Jr. 1938. Preliminary Report on the Smithsonian Institution–­Harvard University Archaeology 100

S T EI N B R EN N ER A N D M CCA F F ERT Y

Expedition to Northwestern Honduras, 1936. Smithsonian Miscellaneous Collections, vol. 97, no. 1. Publication 3445. Washington, DC: Smithsonian Institution. Stuiver, Minze, and Edward S. Deevey. 1961. “Yale Natural Radiocarbon Measurements VI.” Radiocarbon 3:126–­140. Sweeney, Jeanne W. 1975. “Guanacaste, Costa Rica: An Analysis of Precolumbian Ceramics from the Northwest Coast.” PhD diss., University of Pennsylvania. Vázquez Leiva, Ricardo, et al. 1994. “Hacia futuras investigaciones en Gran Nicoya.” Vínculos: Revista de Antropología del Museo Nacional de Costa Rica 18–­19(1–­2 [1992–­1993]):245–­277. Wilke, Sacha. 2012. “From Remains to Rituals: Exploring the Changing Mortuary Program at El Rayo, Nicaragua.” MA thesis, University of British Columbia, Vancouver. Wonderley, Anthony W. 1981. Late Postclassic Excavations at Naco, Honduras. Latin American Studies Program. Diss. Series 86. Ithaca, NY: Latin American Studies Program, Cornell University. Wonderley, Anthony W. 1986. “Material Symbolics in Pre-­Columbian Households: The Painted Pottery of Naco, Honduras.” Journal of Anthropological Research 42(4):497–­534. Wyckoff, Lydia L. 1976. “The Role of Northwestern Nicaragua in the Relations between Mesoamerica and the Greater Nicoya Sub-­Area.” MA thesis, Wesleyan University, Middletown, CT. Wyss, Sue Bursey. 1983. “San Cristobal Archaeological Site, Managua, Nicaragua: Site Report and Preliminary Ceramic Analysis.” MA thesis, Texas A&M University, College Station.

A C ritical R eevaluation of Pacific N icaragua’ s L ate P eriod C hronology

101

Part 2

Projects and Surveys

5 During the years 1995–­1997, Clemente Guido Martínez (director of culture and historical patrimony for the Alcaldía of Managua), Leonor Martínez de  Rocha (director of the National Museum of Nicaragua), Mario Molina (director general of cultural patrimony) and the author (then curator of anthropology at the Natural History Museum of the University of Colorado) collaborated on an ambitious project to inventory, investigate, interpret, and promote the conservation of the archaeological heritage of the Metropolitan area of Managua. The monographs from the 1995 and 1996 field seasons (Lange 1995, 1996) were published cooperatively by the Alcaldía of Managua, the National Museum of Nicaragua, the San Marcos Campus of the University of Mobile, and the University of Colorado, Boulder. The monograph for the 1997 field season was prepared but never published, since the project agreed to release publication funds promised by the Alcaldía of Managua for hurricane relief in Puerto Momotombo (near León Viejo) following Hurricane Mitch in 1998. Now, some twenty years later, the present chapter provides the opportunity to incorporate the 1997 data with the 1995 and 1996 data to provide a more complete overview of the spatial and temporal contents of the occupation of prehistoric Managua. Figure 5.1 shows that some research forays were conducted in the surrounding area, outside the Managua Basin, specifically San Marcos (Carazo) and Laguna de Moyuá

The Managua Metropolitan Project A Retrospective Frederick W. Lange

DOI: 10.5876/9781646421510.c005

105

Figure 5.1. Sites discussed in this chapter. Map by Adam Benfer.

(Matagalpa). In addition, this chapter briefly summarizes the papers regarding ceramic type revision, new-­for-­their-­time data recovery techniques such as GPS and flotation, and overviews of ceramic sequences and regional contacts that were important results of the projects, as well as the contributions of the survey and testing efforts. This chapter also highlights the public education efforts of the 1995–­1997 project and in the conclusion briefly touches on advances in Nicaraguan, Greater Nicoya, and Lower Central America studies that were fomented by the Managua Metropolitan Area Project. Additional research has been ongoing in the Managua region, but no efforts have been made to incorporate these data at the present time; hopefully this will be possible in the near future. SITE SURVEY, TESTING, AND RECORDATION

The following sites were surveyed and tested, some for multiple seasons, during the Metropolitan Managua Project. All projects involved Nicaraguan and North American (and sometimes European or Asian) students and young professionals. All of these sites were either previously known to local archaeologists or were brought to the project’s attention by members of the public. 106

LANGE

Los Placeres (N-­M A-­1 )

This site was tested in 1996 (Pichardo Pichardo 1996; Stauber 1996). It was occupied from ad 500–­1520, or until Spanish contact. Two test pits and two controlled surface collections marked the preliminary research. The findings demonstrated the possible presence of wooden post architecture (possibly a stockade line between the main mound and the lakeshore). The testing also revealed an area with dense domestic refuse including evidence for fishing and recovered a high density and wide variety of lithics from nonlocal sources. The site contains large mounds indicative of ceremonial activity or ranked social structure, with the domestic settlement close to the mounds. It is thought that there was a mixture of long-­distance and regional exchange, with an emphasis on local organization. San Cristóbal (N-­M A-­2 )

San Cristóbal is a large site off the eastern end of the runway at Augusto Cesar Sandino International Airport and less than one mile southeast of the shore of Lake Managua. Excavations were conducted here in the late 1970s by Susan Wyss (1983).1 The site was visited by Lange and Sheets (accompanied by Aníbal Martínez) in 1983 (Lange et al. 1992), and sometime shortly thereafter the earthen mounds at the site were demolished by an agricultural reform project. During the 1997 project, Zak Fink et al. (forthcoming) demonstrated the value of continuing to investigate severely impacted sites. Excavations at the site indicated occupation and utilization as a cemetery from 800 bc to ad 1520 (until Spanish contact). The investigations showed evidence of Middle Formative occupation and demonstrated a broad, if not continuous, presence of Formative populations. There was also evidence of large cemeteries with two different interment styles: the upper cemetery level consisted of flexed burials placed in shoe-­shaped funerary urns and the lower level consisted of internments placed in a prone position without the use of shoe-­shaped vessels. Ceramics, lithics, animal bone, and quantities of small clay beads were present. Students studied the range of appliqué decoration on the external surface of the shoe-­shaped funerary vessels (figure 5.2), which suggests that the shoe-­shaped vessels were made at the time of death and that each such vessel was unique in its decoration. There has been some speculation that the bulge in the vessel at the end opposite 1. Editor’s note: some of the ceramic material collected by Wyss from San Cristóbal received a more detailed examination in Steinbrenner’s (2010) doctoral dissertation focusing on the late Sapoá-­period ceramics of Pacific Nicaragua (see chapter 9, this volume). T he M anagua M etropolitan P roject

107

Figure 5.2. Sacasa Striated burial urn, San Cristóbal site. Photo by author.

the mouth of the vessel represented female pregnancy, but this vessel form more likely was made to accommodate the long bones of the postcranial skeleton. Ciudad Sandino (N-­M A-­1 2, N-MA-­3 7)

García Vásquez (1995) presented a summary of previous excavations in Ciudad Sandino. He indicated that the existing data reflected a long period of occupation (Tempisque period to Ometepe period) and both domestic and funerary activities. In addition, two small projects were also carried out in Ciudad Sandino during 1996 (González Rivas, Zambrana Fernández, and Gámez Montenegro 1996; Keller, Lawson, and Sievers 1996). Las Brisas (N-­M A-­2 4)

This was a small salvage project (Zambrana Fernández and García Vásquez 1995) carried out near the Lenin Fonseca Hospital. A single urn was rescued 108

LANGE

after it was found and reported by local residents. Inside were the remains of a small infant (12–­18 months). Given the size of the urn and some associated ceramics, the burial was placed at approximately ad 1100. Jorge Zambrana Fernández (1995b) prepared a type description for a new ceramic type (Las Brisas Stamped) found at the site. El Ferrocarril (N-­M A-­3 5)

This was a small salvage project (Zambrana Fernández and García Vásquez 1995) carried out near a street and drainage where discarded materials from the railroad had been deposited. The portion of the site that was exposed in 1995 was limited to one funerary urn. Associated artifacts included fragmentary human bones, ceramic beads from a necklace or bracelet, and ceramics. The chronological placement of the urn is uncertain. Villa Tiscapa Sector of the Villa Tiscapa Site (N-­M A-­3 6)

Heidi Pullen (1995) described the chronology of this site (figure 5.3), based on surface collections that produced ceramics dating from the Tempisque period to the Ometepe period. She noted (Pullen 1995:41) that the ceramics from Villa Tiscapa supported the division of Greater Nicoya into a northern sector and a southern sector. She also noted the presence of obsidian and suggested that this represented contacts with either Honduras or Guatemala, or both. Quite unexpectedly, excavations during the 1995 field season recorded the oldest ceramics yet found in the Managua area (Espinoza Pérez 1995b). These ceramics dated from approximately 2000  bc (by cross-­dating with similar ceramics from southern Mesoamerica and from the Tronadora area in the Guanacaste highlands in adjacent Costa Rica). Previously, ceramics of the periods prior to 800 bc had been found in Nicaragua only on Ometepe Island (Haberland 1992). This was the first recording of an early complex in the Isthmus of Rivas. The early phases were divided into two complexes by Espinoza Pérez: the Tiscapa phase (predating 2000 bc) and the Pinata phase (2000 bc to ad 500). While the original publication (Espinoza Pérez 1995b:17) stated that both early Nicaraguan phases were found in Costa Rica, further investigation has brought this author to the conclusion that the Tiscapa phase is absent from northern Costa Rica, while the Pinata phase is present. Excavation at Villa Tiscapa continued in 1996 (Marshall Brown, Margaret Krieg, and Christopher Wilmott 1996), in an effort to further determine the span of occupation. The 1996 excavation yielded a much larger sample T he M anagua M etropolitan P roject

109

Figure 5.3. Local school group visiting Villa Tiscapa during 1995 field season. Photo by author.

of Usulután-­style ceramics (Demarest and Sharer 1982; Lange et al. 2003). A large quantity of obsidian was found, of which 50 percent had cortex and 50  percent did not. Partially worked and unworked small obsidian nodules were also found, representing micro-­core and flake technology at the site. The presence of small prismatic blades—­larger than any of the small, obsidian nodules used for flake manufacture—­and the absence of cores of sufficient size for blade production indicate that the prismatic blades were imported already prepared from sources in Guatemala or Honduras. This is not only evidence of trade with southern Mesoamerica, but also of the transfer of ideas and technology (Boyette and Zambrana Fernández 1995). Boyette and Zambrana Fernández also noted the presence of jadeite beads. The architectural features at the site include alignments of stones that were probably walls, with a section of plaster floor, post molds, and a firepit dated to ad 70 (Beta 94419 and Beta 95901). UCA Sector of the Villa Tiscapa Site (N-­M A-­3 6)

In 1997, the third season of the Metropolitan Managua Project, Laurie Rinfret, Ed Fielder, and Kate Armstrong (forthcoming) and González Rivas 110

LANGE

(forthcoming) excavated four 2 × 2 m tests at this site in order to demonstrate the spatial unity of the UCA, UNI, and Villa Tiscapa sites. These excavations helped to confirm the broad extension and dense prehistoric occupation of the shoreline of ancient Lake Managua. Excavations in this sector focused on a concentration of trash dating from the Tempisque period (500 bc to ad 300). This sector has been severely impacted and partially destroyed by the construction of the sports field at the UCA. Rafael González Rivas, Jorge Zambrana Fernández, and Gámez Montenegro (1996) noted that their study of this parcel revealed differences in the density of cultural resources in different areas of the site, as well as differences in the geomorphological formation processes. UNI Sector of the Villa Tiscapa Site (N-­M A-­ 62, renumbered to be part of N-MA-­3 6)

Excavations at this site in 1995 (Pichardo Pichardo and Zambrana Fernández 1995) and 1996 (Bargnesi, Hartman, and Dirksen 1996) resulted in the recovery of ceramics from all five principal chronological periods: Orosí, Tempisque, Bagaces, Sapoá, and Ometepe. An interesting find among the ceramic collections included a number of potsherds that appear to have woven cloth impressions on the interior of the vessels. The excavated lithic remains included basalt, chalcedony, quartz, andesite, jasper, and obsidian materials. All of these lithic materials can be locally obtained in Nicaragua, while the obsidian came from either Honduras or Guatemala (probably the latter). Finally, one mano (grinding stone) suggests the nature of agricultural practices in Managua in the past. Las Torres (N-­M A-­3 8)

This site was reported to the project by local residents. The excavation (García Vásquez, Amanda Diers, and Serena Algozar 1996) recovered part of a cemetery that contained groups of shoe-­shaped urns that appear to have been arranged in family clusters. Small, miniature shoe-­shaped urns accompanied the larger urns. Stone beads and other items of personal adornment were also present. Most of the ceramics were from the Sapoá and Ometepe periods (ad 800–­1520, until Spanish contact). Acahualinca (N-­M A-­6 1)

Acahualinca (figure 5.4) is perhaps the best-­known site in Nicaragua because of its trail of footprints imprinted in a mud flat that was formerly part of T he M anagua M etropolitan P roject

111

the southern shore of Lake Managua (Espinoza Pérez 1995a; González Rivas 1995). Adults, children, birds, and animals were all part of the procession. Part of the 1995 field project was dedicated to a new stratigraphic test at the site. Despite much previous investigation, a documented chronological sequence had never been described. In the process of excavating a stratigraphic test, a human interment was found at a depth of 1.82 m (Zambrana Fernández 1995a). Two ceramic vessels were associated with the interment and suggested that it dated to approximately ad 600–­1350 (García Vásquez 1996). Sitio Ticomo (N-­M A-­6 4)

During the 1997 field season, this was a short-­term project carried out by Estuo Hasegawa (forthcoming, a). It expanded our knowledge of the prehistoric occupation of the Managua metropolitan area during the Tempisque, Bagaces, and Ometepe periods. Obsidian and basalt lithics were present. Sitio La Arenera (N-­M A-­5 )

Edgar Espinoza Pérez, Fumiyo Suganuma, and Luvy Pichardo Pichardo (forthcoming) reported that ceramics from the Tempisque and Bagaces periods were found beneath a 4 m thick layer of volcanic ash during the extraction of building materials. Villa Fontana Norte (N-­M A-­6 6)

During the 1997 field season, this was a short-­term project carried out by Hasegawa (forthcoming, b). It expanded our knowledge of the prehistoric occupation of the Managua metropolitan area during the Tempisque, Bagaces, and Ometepe periods. Obsidian was also present. San Marcos, Carazo (San José Robleto, N-CA-­1 )

Bárbara Piperata (1995) described the survey of the San José Robleto site, some 2.5 km from the town of San Marcos. Diagnostic ceramics dated from 500 bc to ad 1520. The lithic assemblage included a metate leg, a pestle fragment, and a chalcedony projectile point or knife. No obsidian was observed at this site. San Marcos, Carazo (Barrio La Cruz, N-­C A-­3 )

Preliminary research was conducted during the 1995 field season (Michelle Brown 1995) on this site, which is on both sides of the road when entering San 112

LANGE

Figure 5.4. Huellas de Acahualinca site. Photo by author.

Marcos. Silvia Salgado González, Alejandra Bolaños, and Edgar Guerrero (Salgado González, Bolaños, and Guerrero, forthcoming) continued excavations during the 1997 field season. Preliminary testing and surface collections in Barrio La Cruz identified both domestic and mortuary contexts. The T he M anagua M etropolitan P roject

113

occupation occurred over a 2,500-­year period, from the Orosí period through to the Sapoá period (800 bc to ad 1350), and also revealed the presence of obsidian artifacts, which had not been reported from sites in “the pueblos” (i.e., the hilly region southwest of Managua) up to that time. San Marcos, Carazo (Petroglyph Site, N-­C A-­4 )

During the 1995 survey, Piperata and Raymond Kokaly (1995) located a site with multiple petroglyph panels in the area of San Marcos. The site consists of four panels on the southern wall of a relatively deep canyon. Most of the carvings represent human faces with masks, with and without ornamentation. In addition, there are geometric forms, crosses, concentric circles, and zoomorphic images in the forms of birds and jaguars. Two adjacent boulders had depressions in their surfaces. Laguna de Moyuá (Matagalpa)

During the 1996 field season Kerri Finlayson (1996, 1998) conducted a systematic surface survey and collection of the Moyuá site in the department of Matagalpa (figure 5.5). The site designated as Moyuá consists of two islands: Isla Honda and Isla Seca (Moyuá), which are located in the Laguna de Moyuá, some 72 km north of the capital of Managua. This project was incorporated as one of the objectives of the Managua Metropolitan Area Project in order to increase the previous archaeological samples collected by Lange et al. (1992:44–­45), allowing for better registration of the prehistoric data from the islands as well as a more detailed study of the ceramics and the lithics. RESEARCH RESULTS

Above all, the Metropolitan Area Project was a three-­year research effort that made the following important contributions to developing a more complete understanding of Nicaraguan prehistory: 1. promoting neighborhood cooperation with National Museum and other cultural officials in efforts to protect and preserve the cultural heritage of the Managua metropolitan area; 2. promoting institutional cooperation among the Ministry of Culture and National Museum of Nicaragua and other ministries and land-­managing agencies; 114

LANGE

Figure 5.5. Partially dry Laguna de Moyuá. Photo by author. 3. promoting professional training for both Nicaraguan nationals and selected North American, European, and Asian students; 4. identifying the earliest ceramics (ca. 1800 bc) and domestic remains (ca. 600 bc) known to date in Nicaragua; 5. proving the existence of regional trade in obsidian from Guatemala and Honduras to Nicaragua by 1200 bc, more than 1,000 years earlier than previously known; 6. recovering the largest collection of scientifically excavated shoe-­shaped burial urns to date; 7. demonstrating the long term (more the 6,000 years) of permanent human occupation of the shoreline of Lake Managua and Lake Nicaragua; and 8 . identifying several key sites (Los Placeres, San Cristóbal, and Laguna de Moyuá) with evidence of social complexity. RESEARCH DESIGNS, SITE RECORDATION, LABORATORY ANALYSIS, AND PUBLIC ARCHAEOLOGY

In addition to the survey and testing activities reported above, the project also emphasized the importance of establishing a research design prior to T he M anagua M etropolitan P roject

115

each field season and summarizing the results at the end of that field season (Lange 1995a, forthcoming, a and b; Lange 1996; Lange and Espinoza Pérez 1995; Lange, Espinoza Pérez, and García Vásquez 1996). An ongoing theme in the research design was to increase the understanding of the behavior of the prehistoric residents of Managua in their spatial location in between southern Mesoamerica and the non-­Mesoamerican peoples to the south and east. To increase this understanding, additional project report chapters synthesized ceramic and lithic assemblages and the human skeletal remains from multiple sites, or (as the recorded sample increased) from the entire project (Bargnesi 1996; Boyette and Zambrana Fernández 1995; Espinoza Pérez 1995a, 1995b; Espinoza Pérez and García Vásquez 1995; García Vásquez 1995; García Vásquez, forthcoming; Lange 1995b). While the techniques for flotation (Dickau 1999, forthcoming), GPS (Kokaly 1995), and the INAA evaluation of ceramic samples (Bishop, Lange, and Lange 1988; Bishop et al. 1992; see also chapter 10, this volume) have become much more commonplace in the past twenty years, in 1995–­1997 these were innovative approaches to the study of Nicaragua’s past. The project also met with variable success in meeting its goal to enlist the support of the general populace of Managua in the struggle to preserve the past, and three articles summarized these efforts (Lange and Espinoza Pérez 1995; Lange, Espinoza Pérez, and García Vásquez 1996; Sánchez Barquero and Pichardo Pichardo, forthcoming). Finally, continually reminding ourselves of the archaeologist’s mantra that “more work needs to be done,” the project reports were careful to set goals for the future. Primarily among these were further spatial and temporal investigations, increased INAA applications to ceramic analysis and x-ray diffraction analysis of obsidian specimens (Eilert, Greiner, and Kehoe 1996), and dietary reconstruction from human bone analysis (Wheeler, García Vásquez, and Diers 1996), as well as the need to develop a focus on the historic period and representative early Spanish colonial sites in the metropolitan area (Werner 1995). At this juncture León Viejo is much better known in this regard. These unrealized targets are repeated briefly here in the hopes that they will motivate some ambitious student to follow up on the proposed research themes. In summary, the Managua Metropolitan Project was a Nicaraguan project focused on creating a greater awareness for the importance of preserving the nation’s past for future generations. It is therefore entirely appropriate to finalize this summary using the words of the humble but proud residents of one of the poorer sections of Ciudad Sandino who contributed their respect for their nation’s cultural heritage to the prologue to the 1996 project monograph: 116

LANGE

. . . Reconocemos que las urnas encontradas son Patrimonio Nacional. Por lo tanto, hacemos la donación de las mismas para colaborar así tanto para que Nicaragua se beneficie y obtenga más información de nuestros antepasados aborígenes, como para la arqueología que estudie la estructura y antigüedad del hallazgo. A la vez, queremos que nuestros hijos puedan conocer, visualizar y aprender de los aborígenes, sus costumbres, mitos, etc. como parte de la historia nicaragüense.2 ACKNOWLEDGMENTS

The Managua Metropolitan Archaeological project was made possible through the leadership of the Alcaldía of Managua, with the support of the National Museum of Nicaragua and the Office of Cultural Patrimony of the Ministry of Culture. The University of Colorado, Boulder; Iowa State University; the Colorado School of Mines; and the University of Mobile, Latin American campus proudly provided logistical resources, staff, and training and educational opportunities. Arnoldo Aleman Lacayo, Roberto Cedeño Borgen, and Myriam Fonseca López were the mayors of Managua during this period. The Honorable John Maisto and the Honorable Lino Gutiérrez were both United States ambassadors to Nicaragua during the period of the project; both visited some of the sites and the project laboratory and interacted with the students and Nicaraguan archaeologists. REFERENCES

Bargnesi, Keely. 1996. “Los restos óseos humanos excavados en 1996.” In Abundante cooperación vecinal: La segunda temporada del proyecto “Arqueología de la Zona Metropolitana de Managua,” edited by Frederick W. Lange, 125–­129. Managua: Alcaldía de Managua. Bargnesi, Keely, Krystal Hartman, and Maribeth Dirksen. 1996. “Excavación en el Sitio N-MA-­62 (UNI).” In Abundante cooperación vecinal: La segunda temporada del proyecto “Arqueología de la Zona Metropolitana de Managua,” edited by Frederick W. Lange, 37–­48. Managua: Alcaldía de Managua. 2. “We accept that the urns found [in the excavations] are cultural Patrimony of the nation. Therefore, we donate them in collaboration so that all Nicaraguans will benefit and obtain more information about our aboriginal ancestors, as will the archaeologists who study the structure and antiquity of these discoveries. “At the same time, we want our children to be able to know, visualize, and learn from the Indigenous peoples—­their customs, myths, etc.—­as part of Nicaraguan history.” T he M anagua M etropolitan P roject

117

Bishop, Ronald L., Frederick W. Lange, Suzanne Abel-­V idor, and Peter C. Lange. 1992. “Compositional Characterization of the Nicaraguan Ceramic Sample.” In The Archaeology of Pacific Nicaragua, edited by Frederick W. Lange, Payson D. Sheets, Aníbal Martínez, and Suzanne Abel-­V idor, 135–­162. Albuquerque: University of New Mexico Press. Bishop, Ronald L., Frederick W. Lange, and Peter C. Lange. 1988. “Ceramic Paste Compositional Patterns in Greater Nicoya Pottery.” In Costa Rican Art and Archaeology: Essays in Honor of Frederick R. Mayer, edited by Frederick W. Lange, 11–­44. Boulder: University of Colorado. Boyette, Melissa, and Jorge E. Zambrana Fernández. 1995. “Análisis de la muestra lítica: Proyecto Arqueología de la Zona Metropolitana de Managua—­1995.” In Descubriendo las huellas de nuestros antepasados: El proyecto “Arqueología de la Zona Metropolitana de Managua,” edited by Frederick W. Lange, 51–­60. Managua: Alcaldía de Managua. Brown, Marshall, Margaret Krieg, and Christopher Wilmott. 1996. “La segunda temporada en el sitio Villa Tiscapa (N-­MA-­36).” In Abundante cooperación vecinal: La segunda temporada del proyecto “Arqueología de la Zona Metropolitana de Managua,” edited by Frederick W. Lange, 9–­36. Managua: Alcaldía de Managua. Brown, Michelle Lee. 1995. “Prospección y rescate en Barrio La Cruz, San Marcos.” In Descubriendo las huellas de nuestros antepasados: El proyecto “Arqueología de la Zona Metropolitana de Managua,” edited by Frederick W. Lange, 71–­74. Managua: Alcaldía de Managua. Demarest, Arthur A., and Robert J. Sharer. 1982. “The Origins and Evolution of the Usulutan Ceramic Style.” American Antiquity 47(4):810–­822. Dickau, Ruth E. 1999. “Paleoethnobotany of the Lake Managua Region, Nicaragua.” MA thesis, University of Calgary. Dickau, Ruth E. Forthcoming. “Flotación e investigaciones paleoetnobotánicas del área del Lago de Nicaragua.” In Debajo de tus pies: La tercera temporada del proyecto “Arqueología de la Zona Metropolitana de Managua.” Manuscript on file, Archivo Informes, Departamento de Arqueología, Museo Nacional de Nicaragua (MNN), Dirección de Patrimonio Cultural (DPC), Instituto Nicaragüense de Cultura (INC), Managua [1997]. Eilert, Eloise, Aaron Greiner, and Justin Kehoe. 1996. “La autogestión cultural: Hacia una tercera etapa del Proyecto ‘Arqueología de la Zona Metropolitana de Managua.’ ” In Abundante cooperación vecinal: La segunda temporada del proyecto “Arqueología de la Zona Metropolitana de Managua,” edited by Frederick W. Lange, 153–­162. Managua: Alcaldía de Managua.

118

LANGE

Espinoza Pérez, Edgar. 1995a. “Acahualinca al pie de la letra: Una revaluación del sitio.” In Descubriendo las huellas de nuestros antepasados: El proyecto “Arqueología de la Zona Metropolitana de Managua,” edited by Frederick W. Lange, 91–­94. Managua: Alcaldía de Managua. Espinoza Pérez, Edgar. 1995b. “La cerámica temprana de Nicaragua y sus vínculos regionales.” In Descubriendo las huellas de nuestros antepasados: El proyecto “Arqueo­ logía de la Zona Metropolitana de Managua,” edited by Frederick W. Lange, 17–­24. Managua: Alcaldía de Managua. Espinoza Pérez, Edgar, and Ramiro García Vásquez. 1995. “Cerámica del sur de Honduras encontrada en los sitios del Proyecto Arqueología de la Zona Metropolitana.” In Descubriendo las huellas de nuestros antepasados: El proyecto “Arqueología de la Zona Metropolitana de Managua,” edited by Frederick W. Lange, 99–­102. Managua: Alcaldía de Managua. Espinoza Pérez, Edgar, Fumiyo Suganuma, and Luvy Pichardo Pichardo. Forthcoming. “Reconocimiento en el sitio La Arenera (N-­MA-­65).” In Debajo de tus pies: La tercera temporada del proyecto “Arqueología de la Zona Metropolitana de Managua.” Manuscript on file, Archivo Informes, Departamento de Arqueología, Museo Nacional de Nicaragua (MNN), Dirección de Patrimonio Cultural (DPC), Instituto Nicaragüense de Cultura (INC), Managua [1997]. Fink, Zak, Lucas Newman, Jacob Yacko, Sarah Piziali, and Matthew Cunningham. Forthcoming. “Excavaciones en el sitio San Cristóbal (N-­MA-­2).” In Debajo de tus pies: La tercera temporada del proyecto “Arqueología de la Zona Metropolitana de Managua.” Manuscript on file, Archivo Informes, Departamento de Arqueo­ logía, Museo Nacional de Nicaragua (MNN), Dirección de Patrimonio Cultural (DPC), Instituto Nicaragüense de Cultura (INC), Managua [1997]. Finlayson, Kerri. 1996. “Prospección y excavación preliminar en la zona de la Laguna de Moyuá.” In Abundante cooperación vecinal: La segunda temporada del proyecto “Arqueología de la Zona Metropolitana de Managua,” edited by Frederick W. Lange, 133–­149. Managua: Alcaldía de Managua. Finlayson, Kerri. 1998. “Stone Tool Industries of Managua, Nicaragua and Their Evolution through Time.” MA thesis, University of Illinois, Urbana-­Champaign. García Vásquez, Ramiro. 1995. “Secuencia cerámica de Ciudad Sandino.” In Descubriendo las huellas de nuestros antepasados: El proyecto “Arqueología de la Zona Metro­ politana de Managua,” edited by Frederick W. Lange, 37–­40. Managua: Alcaldía de Managua. García Vásquez, Ramiro. 1996. “Hallazgos recientes en Acahualinca (N-­MA-­61).” In Abundante cooperación vecinal: La segunda temporada del proyecto “Arqueología

T he M anagua M etropolitan P roject

119

de la Zona Metropolitana de Managua,” edited by Frederick W. Lange, 99–­104. Managua: Alcaldía de Managua. García Vásquez, Ramiro. Forthcoming. “Costumbres funerarias en la Managua precolombina.” In Debajo de tus pies: La tercera temporada del proyecto “Arqueología de la Zona Metropolitana de Managua.” Manuscript on file, Archivo Informes, Departamento de Arqueología, Museo Nacional de Nicaragua (MNN), Dirección de Patrimonio Cultural (DPC), Instituto Nicaragüense de Cultura (INC), Managua [1997]. González Rivas, Rafael. 1995. “La secuencia cerámica de sondeo #4, sitio Huellas de Acahualinca (N-­MA-­61).” In Descubriendo las huellas de nuestros antepasados: El proyecto “Arqueología de la Zona Metropolitana de Managua,” edited by Frederick W. Lange, 27–­36. Managua: Alcaldía de Managua. González Rivas, Rafael. Forthcoming. “Excavación de unidades 1–­4 del sitio UCA (N-­MA-­36).” In Debajo de tus pies: La tercera temporada del proyecto “Arqueología de la Zona Metropolitana de Managua.” Manuscript on file, Archivo Informes, Departamento de Arqueología, Museo Nacional de Nicaragua (MNN), Dirección de Patrimonio Cultural (DPC), Instituto Nicaragüense de Cultura (INC), Managua [1997]. García Vásquez, Ramiro, Serena Algozar, and Amanda Diers. 1996. “Hallazgo arqueo­ lógico en el Barrio Las Torres (N-­MA-­38), Managua: Un posible cementerio con entierros múltiples.” In Abundante cooperación vecinal: La segunda temporada del proyecto “Arqueología de la Zona Metropolitana de Managua,” edited by Frederick W. Lange, 105–­124. Managua: Alcaldía de Managua. González Rivas, Rafael, Jorge E. Zambrana Fernández, and Bayardo Gámez Montenegro. 1996. “Excavaciones en Ciudad Sandino (N-­MA-­12).” In Abundante cooperación vecinal: La segunda temporada del proyecto “Arqueología de la Zona Metro­ politana de Managua,” edited by Frederick W. Lange, 79–­84. Managua: Alcaldía de Managua. Haberland, Wolfgang. 1992. “The Culture History of Ometepe Island: Preliminary Sketch (Survey and Excavations, 1962–­1963).” In The Archaeology of Pacific Nicaragua, edited by Frederick W. Lange, Payson D. Sheets, Aníbal Martínez, and Suzanne Abel-­Vidor, 63–­117. Albuquerque: University of New Mexico Press. Hasegawa, Estuo. Forthcoming, a. “Investigaciones arqueológicas en el sitio Ticomo (N-­MA-­64).” In Debajo de tus pies: La tercera temporada del proyecto “Arqueología de la Zona Metropolitana de Managua.” Manuscript on file, Archivo Informes, Departamento de Arqueología, Museo Nacional de Nicaragua (MNN), Dirección de Patrimonio Cultural (DPC), Instituto Nicaragüense de Cultura (INC), Managua [1997]. 120

LANGE

Hasegawa, Estuo. Forthcoming, b. “Investigaciones arqueológicas en el sitio Villa Fontana Norte, Managua, Nicaragua.” In Debajo de tus pies: La tercera temporada del proyecto “Arqueología de la Zona Metropolitana de Managua.” Manuscript on file, Archivo Informes, Departamento de Arqueología, Museo Nacional de Nicaragua (MNN), Dirección de Patrimonio Cultural (DPC), Instituto Nicaragüense de Cultura (INC), Managua [1997]. Keller, Chad, Charles Lawson, and Ryan Sievers. 1996. “Excavaciones en Ciudad Sandino (N-­MA-­37).” In Abundante cooperación vecinal: La segunda temporada del proyecto “Arqueología de la Zona Metropolitana de Managua,” edited by Frederick W. Lange, 85–­98. Managua: Alcaldía de Managua. Kokaly, Raymond F. 1995. “Aplicación de la tecnología GPS (‘Sistema de Ubicación Global’) en la arqueología nicaragüense.” In Descubriendo las huellas de nuestros antepasados: El proyecto “Arqueología de la Zona Metropolitana de Managua,” edited by Frederick W. Lange, 95–­98. Managua: Alcaldía de Managua. Lange, Frederick W., ed. 1995a. Descubriendo las huellas de nuestros antepasados: El proyecto “Arqueología de la Zona Metropolitana de Managua.” Managua: Alcaldía de Managua. Lange, Frederick W. 1995b. “La presencia de la cerámica de la Gran Nicoya en los sitios arqueológicos investigados en el proyecto metropolitano.” In Descubriendo las huellas de nuestros antepasados: El proyecto “Arqueología de la Zona Metropolitana de Managua,” edited by Frederick W. Lange, 103–­118. Managua: Alcaldía de Managua. Lange, Frederick W., ed. 1996. Abundante cooperación vecinal: La segunda temporada del proyecto “Arqueología de la Zona Metropolitana de Managua.” Managua: Alcaldía de Managua. Lange, Frederick W. Forthcoming, a. “Debajo de tus pies: La tercera temporada del proyecto ‘Arqueología de la Zona Metropolitana de Managua’ [1997].” In Debajo de tus pies: La tercera temporada del proyecto “Arqueología de la Zona Metropolitana de Managua.” Manuscript on file, Archivo Informes, Departamento de Arqueo­ logía, Museo Nacional de Nicaragua (MNN), Dirección de Patrimonio Cultural (DPC), Instituto Nicaragüense de Cultura (INC), Managua [1997]. Lange, Frederick W. Forthcoming, b. “Los resultados científicos del proyecto metro­ politano en su contexto regional.” In Debajo de tus pies: La tercera temporada del proyecto “Arqueología de la Zona Metropolitana de Managua.” Manuscript on file, Archivo Informes, Departamento de Arqueología, Museo Nacional de Nicaragua (MNN), Dirección de Patrimonio Cultural (DPC), Instituto Nicaragüense de Cultura (INC), Managua [1997].

T he M anagua M etropolitan P roject

121

Lange, Frederick W., and Edgar Espinoza Pérez. 1995. “Los recursos arqueológicos urbanos de Nicaragua y su preservación.” In Descubriendo las huellas de nuestros antepasados: El proyecto “Arqueología de la Zona Metropolitana de Managua,” edited by Frederick W. Lange, 128–­134. Managua: Alcaldía de Managua. Lange, Frederick W., Edgar Espinoza Pérez, and Ramiro García Vásquez. 1996. “La autogestión cultural: Hacia una tercera etapa del Proyecto ‘Arqueología de la Zona Metropolitana de Managua.’ ” In Abundante cooperación vecinal: La segunda temporada del proyecto “Arqueología de la Zona Metropolitana de Managua,” edited by Frederick W. Lange, 173–­176. Managua: Alcaldía de Managua. Lange, Frederick W., Erin L. Sears, Ronald L. Bishop, and Silvia Salgado González. 2003. “Local Production, Non-­local Production, and Distribution: Usulutan and Usulutan-­like Negative Painted Ceramics in Nicaragua.” In Patterns and Process: A Festschrift in Honor of Dr. Edward V. Sayre, edited by Lambertus van Zelst, 157–­171. Suitland, MD: Smithsonian Center for Materials Research and Education. Lange, Frederick W., Payson D. Sheets, Aníbal Martínez, and Suzanne Abel-­V idor. 1992. The Archaeology of Pacific Nicaragua. Albuquerque: University of New Mexico Press. Pichardo Pichardo, Luvy. 1996. “Sitio Las Placeres: Un posible centro regional a orillas del Lago de Managua.” In Abundante cooperación vecinal: La segunda temporada del proyecto “Arqueología de la Zona Metropolitana de Managua,” edited by Frederick W. Lange, 69–­7 8. Managua: Alcaldía de Managua. Pichardo Pichardo, Luvy, and Jorge E. Zambrana Fernández. 1995. “Prospección y excavación arqueológica sitio N-MA-­62 UNI, Managua.” In Descubriendo las huellas de nuestros antepasados: El proyecto “Arqueología de la Zona Metropolitana de Managua,” edited by Frederick W. Lange, 75–­82. Managua: Alcaldía de Managua. Piperata, Bárbara A. 1995. “La prospección del sitio San José Robleto, San Marcos, Carazo, Nicaragua.” In Descubriendo las huellas de nuestros antepasados: El proyecto “Arqueología de la Zona Metropolitana de Managua,” edited by Frederick W. Lange, 61–­69. Managua: Alcaldía de Managua. Piperata, Bárbara A., and Raymond F. Kokaly. 1995. “Sitio con petroglifos en San Marcos: Sitio N-CA-­4.” In Descubriendo las huellas de nuestros antepasados: El proyecto “Arqueología de la Zona Metropolitana de Managua,” edited by Frederick W. Lange, 83–­90. Managua: Alcaldía de Managua. Pullen, Heidi. 1995. “Prospección arqueológica en Villa Tiscapa sitio N-MA-­36.” In Descubriendo las huellas de nuestros antepasados: El proyecto “Arqueología de la Zona Metropolitana de Managua,” edited by Frederick W. Lange, 41–­50. Managua: Alcaldía de Managua. 122

LANGE

Rinfret, Laurie, Ed Fielder, and Kate Armstrong. Forthcoming. “Investigaciones en el sitio UCA (N-­MA-­36) sondeos 5, 6, 7, 8, and 9.” In Debajo de tus pies: La tercera temporada del proyecto “Arqueología de la Zona Metropolitana de Managua.” Manuscript on file, Archivo Informes, Departamento de Arqueología, Museo Nacional de Nicaragua (MNN), Dirección de Patrimonio Cultural (DPC), Instituto Nicaragüense de Cultura (INC), Managua [1997]. Sánchez Barquero, Claudia, and Luvy Pichardo Pichardo. Forthcoming. “Alternativas de divulgación de la arqueología.” In Debajo de tus pies: La tercera temporada del proyecto “Arqueología de la Zona Metropolitana de Managua.” Manuscript on file, Archivo Informes, Departamento de Arqueología, Museo Nacional de Nicaragua (MNN), Dirección de Patrimonio Cultural (DPC), Instituto Nicaragüense de Cultura (INC), Managua [1997]. Salgado González, Silvia, Alejandra Bolaños, and Edgar Guerrero. Forthcoming. “Prospección y excavaciones en el sitio La Cruz (N-­CA-­3).” In Debajo de tus pies: La tercera temporada del proyecto “Arqueología de la Zona Metropolitana de Managua.” Manuscript on file, Archivo Informes, Departamento de Arqueología, Museo Nacional de Nicaragua (MNN), Dirección de Patrimonio Cultural (DPC), Instituto Nicaragüense de Cultura (INC), Managua [1997]. Stauber, Daniel. 1996. “Excavaciones arqueológicas e investigaciones preliminares en el sitio Las Placeres (N-­MA-­1).” In Abundante cooperación vecinal: La segunda temporada del proyecto “Arqueología de la Zona Metropolitana de Managua,” edited by Frederick W. Lange, 49–­68. Managua: Alcaldía de Managua. Werner, Patrick S. 1995. “Nuevas perspectivas obtenidas de los fuentes históricos y lingüísticos sobre la zona metropolitana precolombina.” In Descubriendo las huellas de nuestros antepasados: El proyecto “Arqueología de la Zona Metropolitana de Managua,” edited by Frederick W. Lange, 119–­128. Managua: Alcaldía de Managua. Wheeler, Brandy, Ramiro García Vásquez, and Amanda Diers. 1996. “Reconstruyendo los regímenes alimentarios de la población precolombina de Managua, utilizando isótopos de huesos.” In Abundante cooperación vecinal: La segunda temporada del proyecto “Arqueología de la Zona Metropolitana de Managua,” edited by Frederick W. Lange, 163–­172. Managua: Alcaldía de Managua. Wyss, Sue Bursey. 1983. “San Cristóbal Archaeological Site, Managua, Nicaragua: Site Report and Preliminary Ceramic Analysis.” MA thesis, Texas A&M University. Zambrana Fernández, Jorge E. 1995a. “Descripción del nuevo tipo cerámico: Las Brisas Impreso.” In Descubriendo las huellas de nuestros antepasados: El proyecto “Arqueología de la Zona Metropolitana de Managua,” edited by Frederick W. Lange, 153–­154. Managua: Alcaldía de Managua.

T he M anagua M etropolitan P roject

123

Zambrana Fernández, Jorge E. 1995b. “Excavación de entierro primario en sondeo no. 4, sitio N-MA-­61 Acahualinca.” In Descubriendo las huellas de nuestros antepasados: El proyecto “Arqueología de la Zona Metropolitana de Managua,” edited by Frederick W. Lange, 25–­26. Managua: Alcaldía de Managua. Zambrana Fernández, Jorge E., and Ramiro García Vásquez. 1995. “Rescate de entierros secundarios en urnas funerarias en los sitios N-MA-­24, Las Brisas and N-MA-­35, El Ferrocarril.” In Descubriendo las huellas de nuestros antepasados: El proyecto “Arqueología de la Zona Metropolitana de Managua,” edited by Frederick W. Lange, 9–­16. Managua: Alcaldía de Managua.

124

LANGE

6 In 2000, archaeologists from the University of Calgary began investigating late pre-­ Hispanic cultures that lived along the shore of Lake Cocibolca (also known as Lake Nicaragua). Specifically, research sought to evaluate historical claims of migrations of Mesoamerican groups from Central Mexico who spoke languages associated with Otomanguean and Uto-­ Aztecan language families (Abel-­V idor 1981; Carmack and Salgado González 2006; Chapman 1974; Fowler 1989). Sixteenth-­ century accounts by Motolinía ([1540] 1951), Fernández de Oviedo y Valdés ([1526] 1950), and Torquemada ([1615] 1975–­83) describe the migration of these groups and their cultural practices, especially religious beliefs (León-­Portilla 1972; for a discussion of these accounts, see chapter 2, this volume). These ethnohistorical sources have been the basis for interColumbian Indigenous history, in preting late pre-­ what has been termed the “out of Mexico” hypothesis (Hoopes and McCafferty 1989; but see G. McCafferty 2015a, b). Yet relatively little archaeological research has been conducted to evaluate these claims (Carmack and Salgado González 2006; Healy 1980, 1988; Salgado González and Vázquez Leiva 2006). In fact, some archaeologists (most notably Frederick Lange), have argued that foreign influences were negligible in the region in pre-­Hispanic times (Lange 1994). Nicaragua is one of the least known archaeological regions of the western hemisphere. It was first explored by Ephraim Squier (1852, 1853), an American

Twenty Years of Nicaraguan Archaeology Results of the University of Calgary Projects Geoffrey G. McCafferty

DOI: 10.5876/9781646421510.c006

125

prehistorian of the mid-­nineteenth century who, among other things, discovered monumental stone sculptures of standing and seated individuals on the islands of Lake Cocibolca. Several other European and North American scholars explored the country in the later nineteenth century (e.g., Bovallius 1886; Bransford 1881), but more archaeological attention was directed toward the “high civilizations” of Mesoamerica and the Andes. Through political turmoil and natural disasters of the twentieth century, archaeology was not a significant endeavor, with a few notable exceptions (e.g., Haberland 1992; Healy 1980; Norweb 1964; Wyss 1983). This began to change in the 1990s, with settlement pattern studies (Niemel 2003; Román Lacayo 2013; Salgado González 1996), and rescue excavations (Espinoza Pérez, García Vásquez, and Suganuma 1999; Lange 1995, 1996a; Zambrana Fernández 2012). The objective of these projects was to identify and inventory archaeological sites and to collect baseline data to develop a fundamental culture history (Lange 1996b). Nevertheless, the overarching historical paradigm for the final centuries of the pre-­Hispanic era assumed a Mexican migration into Pacific Nicaragua that displaced autochthonous populations (probably comprised of Chibchan speakers). With these issues in mind, investigations were begun at the site of Santa Isabel, believed to have been the contact-­period capital of the Nahuat-­speaking Nicarao encountered by Spanish conquistador Gil González in 1522 (figure 6.1). The teyte, or ruler, of the Nicarao was known as “Nicaragua,” and so this investigation had important implications for the political identity of the modern nation (G. McCafferty 2015a). Following four seasons of excavations and subsequent analyses (G. McCafferty 2008; Steinbrenner 2010), results were ambiguous as to the importance of Mexican migration. A second project subsequently focused on the Granada region at the northern end of the lake, at the sites of Tepetate and El Rayo (figure 1.1, this volume; G. McCafferty and Dennett 2013; G. McCafferty, Salgado González, and Dennett 2009). This area was historically associated with the Chorotega, the other major Mesoamerican group in Nicaragua at the time of Spanish contact. Again, despite important discoveries relevant to the larger question, results were ambiguous. A third project (G. McCafferty, Pavón Sánchez, and Galeano Rueda 2013) was conducted at the famous site of Sonzapote on Zapatera Island, where monumental sculpture had been attributed to the migrant Chorotega (Bovallius 1886; Lothrop 1921; Navarro Genie 2007). And once again, the evidence contradicted expectations. This chapter presents interpretations from twenty years of archaeological investigation by the University of Calgary, projects that represent the most intensive archaeological research ever conducted in Nicaragua and that 126

M CCA F F ERT Y

Figure 6.1. Public mural in Rivas depicting 1522 encounter of Gil González with teyte Nicaragua on shore of Lake Cocibolca. Photo by author.

challenge the traditional culture history based on ethnohistorical accounts (G. McCafferty 2015a). Following an overview of the research that was conducted at the four key sites named above, it is divided into sections on chronology, foodways, architecture, mortuary patterns, ornamentation, and specialized craft production. The chapter concludes with an evaluation of the “out of Mexico” hypothesis in light of a diverse range of archaeological evidence. Research for the first two projects in Rivas and Granada was funded by major grants from the Social Sciences and Humanities Research Council of Canada and by the University of Calgary, while the Sonzapote project was funded by the National Geographic / Waitt Foundation. Recent research at El Rayo was funded by the Institute for Field Research as an archaeological field school. Although the initial projects were not conducted as formal field schools, they did offer opportunities for nearly 100 undergraduate students to participate, many experiencing life in Central America for the first time. Numerous graduate students have also participated, resulting in several theses and dissertations (Debert 2005; Dennett 2016; Leullier-­Snedeker 2013; López-­Forment Villa 2008; Manion 2016; Rice 2019; Steinbrenner 2002, 2010; Wilke 2012). The projects also integrated many students from Central American countries, especially Nicaragua, El Salvador, and Costa Rica. T wenty Y ears of N icaraguan A rchaeology

127

RESEARCH OVERVIEW

Santa Isabel

Santa Isabel was identified in the foundational book The Archaeology of Rivas, Nicaragua by Paul Healy (1980), which was based on archaeological work by Gordon Willey and his student Albert Norweb (1964) and later revisited during the archaeological survey of Karen Niemel (2003; see chapter 7, this volume). In these studies, it was interpreted as one of the most important sites of Nicaragua’s Sapoá-­Ometepe/“Postclassic” period (ad 800–­1522) and as possibly corresponding to the historical Quauhcapolca, the Nicarao capital at Spanish contact (Lothrop 1926:6; cf. Fowler 1989:68). From 2000 until 2005 the Proyecto Santa Isabel, Nicaragua investigated domestic features to infer daily practices of the Sapoá period (ad 800–­1300; G. McCafferty 2008) but without finding evidence of the Ometepe period (ad 1300–­1522) that is traditionally associated with the Nicarao. Seven loci were investigated at Santa Isabel, using both shovel tests and horizontal excavations. Mounds 1, 3, and 6 were the most extensively excavated (figure 6.2). Emphasis was placed on broad exposure of the uppermost levels relating to the final occupants. Radiocarbon dates suggest that the site was abandoned by ad 1250–­1300 (G. McCafferty 2008; G. McCafferty and Steinbrenner 2005a). Following the traditional interpretation of ethno­ historical sources, this timing roughly corresponds to the arrival of the Nicarao; the site occupation therefore probably related to the Otomanguean Chorotega cultural group with abandonment possibly relating to events associated with the arrival of the Nicarao. Note, however, that this “traditional” cultural sequence is currently being reevaluated by members of the Calgary team. The Santa Isabel project was one of the first projects in Lower Central America to investigate domestic practices, including residential architecture, foodways, specialized production, and religious ideology (G. McCafferty 2008). Another important result was the use of radiocarbon dating of samples associated with polychrome ceramics to revise the Postclassic chronology (G. McCafferty 2008; G. McCafferty and Steinbrenner 2005a; chapter 4, this volume). Additional analyses have included ceramics (Dennett 2016; G. McCafferty and Steinbrenner 2005b; Steinbrenner 2002, 2010), lithics (Debert and Sheriff 2007), faunal remains (Hoar 2006; López-­Forment Villa 2008), textile production (S. McCafferty and G. McCafferty 2008), and figurines and adornment (Leullier-­Snedeker 2013; G. McCafferty and S. McCafferty 2009, 2011; see also discussions of ceramics in this volume by Steinbrenner and Dennett in chapters 9 and 10). 128

M CCA F F ERT Y

Figure 6.2. Plan of Santa Isabel, indicating excavation loci and mounds. Adapted from Steinbrenner (2010:188).

Tepetate

Tepetate is located on the northern edge of modern Granada and has been known to archaeologists (and looters) for over 100 years (in recent years, the site has been nearly destroyed). Silvia Salgado González (1996) identified it T wenty Y ears of N icaraguan A rchaeology

129

Figure 6.3. Mound 1 at Tepetate. Photo by author.

as a regional center, probably relating to the contact-­period Indigenous community of Xalteva that was known as a capital of the Chorotega. In 2008, excavations by the Proyecto Arqueológico Granada, Nicaragua, explored three loci on the northern edge of the site (G. McCafferty, Salgado González, and Dennett 2009; Zambrana Fernández 2008). Architectural remains were found at Mound 1 (figure 6.3), though the upper levels were badly disturbed by looting activities. Architectural rubble on the surface of the mound suggests that it may have been faced with lajas (flat stone slabs), as was described for mounds by early twentieth-­century visitors to the site (Salgado González 1996). Deeper excavations encountered remnants of paving stones, probably associated with floors, and also a slab-­lined enclosure that was likely a tomb. A second low mound was found on the northern edge of the site, and excavations exposed a stone foundation wall. In Locus 3 we excavated a cemetery of poorly preserved human remains, some associated with “shoe-­shaped” burial urns. 130

M CCA F F ERT Y

Figure 6.4. Plan of El Rayo, indicating excavation loci, 2009–­2016. Adapted from Dennett (2016:119).

El Rayo

The site of El Rayo is located on the Asese Peninsula extending into Lake Cocibolca. The peninsula was formed through the partial collapse of the Mombacho Volcano and subsequent lahar deposits (Shea, van Wyk de Vries, and Pilato 2008; Stansell 2013; see also chapter 10, this volume). El Rayo was discovered by Salgado González (1996) as part of her inventory of sites in the Granada region. In 2007, road construction exposed a Postclassic cemetery, with remains of shoe-­shaped urns. To recover information from the site, investigations began in 2009 and continued in 2010. In 2015 and 2016 investigations began again as part of an archaeological field school. El Rayo is one of the most important archaeological sites excavated in Nicaragua, in part because of the continuous occupation from late Bagaces times through most of the Sapoá period (ad 550–­1200) but also because of its excellent preservation (figure 6.4). It features a range of burial practices, including ritual interments associated with a possible shrine. A well-­defined civic-­ceremonial structure may relate to the overall religious importance of the site. Domestic debris found at Locus 2 allows comparisons of material culture relating to the Bagaces-­Sapoá transition, when ethnic groups from Mexico allegedly moved into Pacific Nicaragua. Careful analysis of these contexts permits a rigorous reevaluation of the migration myth (G. McCafferty and Dennett 2013). T wenty Y ears of N icaraguan A rchaeology

131

Sonzapote

Whereas the previous projects were primarily directed to domestic contexts relating to the Chorotega culture, Sonzapote has a prominent position in the published literature as a ceremonial center, purportedly founded by the same cultural group (Guido Martínez 2004; Lothrop 1921; Navarro Genie 2007). In order to obtain a more complete perspective on Chorotega culture, a small pilot project mapped and sampled archaeological resources at the site (G. McCafferty, Pavón Sánchez, and Galeano Rueda 2013). The main focus of previous investigations was the monumental sculptures found at the site (Bovallius 1886), but most of these were later moved to the nearby city of Granada, where they are now on exhibit at the ex-­Convento San Francisco museum. Very little archaeological research was ever conducted at the site, which among other things featured a dense grouping of low masonry mounds (but see Castillo Barquero 1989; Navarro Genie 2007). A short field season in 2013 accomplished several important goals (G. McCafferty, Pavón Sánchez, and Galeano Rueda 2013). A survey map of the central site core identified seventeen low stone mounds, including evidence of site planning in plaza groups never before recognized in Pacific Nicaragua (figure 6.5). An inventory of more than fifty existing monuments (including petroglyphs, utilized boulders, and statue fragments) augmented the corpus known previously from Carl Bovallius’s and Rigoberto Navarro Genie’s studies. And most surprisingly, excavations at Mound 14 demonstrated that the original occupation dated to the late Tempisque period (ad 1–­300), nearly 500 years prior to Chorotega migration into the region. Additional research is needed to further elaborate on these findings, but the preliminary interpretation is of incipient urbanism far earlier than expected. CHRONOLOGICAL FOUNDATIONS

Archaeological cultures are usually described based on their spatial and temporal dimensions. As such, regional chronology is an important foundation for interpretation. The chronology of Pacific Nicaragua has traditionally been linked to that of nearby Costa Rica, in part because of the relatively greater amount of scientific archaeology conducted there (Bonilla et al. 1990; but see Niemel, Román Lacayo, and Salgado González 1998). In fact, northwestern Costa Rica and Pacific Nicaragua are often lumped into the related cultural region known as Greater Nicoya (Norweb 1964). The result has been a sequence of five time periods of relatively long duration:

132

M CCA F F ERT Y

Figure 6.5. Plan of Sonzapote, indicating mounds and terraces. Map by Shawn Morton, Meaghan Peuramaki-­Brown, Gracia Silva, and Ligia Obando.

Orosí (2000–­500 bc) Tempisque (500 bc–­a d 300) Bagaces (ad 300–­800) Sapoá (ad 800–­1350) Ometepe (ad 1350–­1522). Seventeen radiocarbon samples were processed from Santa Isabel, and all dated between ad 686 and 1280 (calibrated, 2σ), consistent with the Sapoá period (G. McCafferty 2008; G. McCafferty and Steinbrenner 2005a; see table 4.5, this volume). This result was surprising because previous research at the site had concluded that it was occupied through the Ometepe period up until European contact (Healy 1980; Niemel 2003). These previous interpretations were based on the presence of pottery believed to be diagnostic of the Ometepe period, such as Vallejo, Bramadero, Madeira, and El Menco polychromes and Castillo Engraved (Bonilla Vargas et al. 1990). The conclusion T wenty Y ears of N icaraguan A rchaeology

133

based on these new dates is that the traditional ceramic sequence requires revision, with many “Ometepe” diagnostics actually introduced several centuries earlier. It also suggests that the site was abandoned no later than ad 1300 (as previously noted), a date that probably—­for reasons discussed in chapter 4 of this volume—­represents a better ending date for the Sapoá period than the traditional ad 1350. When Ometepe-­period remains were not found at Santa Isabel, the site of Tepetate was selected because it had supposedly also been occupied at the moment of European contact (Carmack and Salgado González 2006; Salgado González 1996). Excavations recovered ceramics associated with the traditional sequence, including Vallejo: Mombacho and other Vallejo polychromes, but not in contexts that could be dated to the Ometepe period. Instead, two 14 C dates once again suggested that the site was occupied in the late Sapoá period, and thus support the hypothesis that the traditional sequence does not correspond well with the new information. El Rayo was also excavated in hopes of encountering Ometepe-­period remains, and again the site appears to have been abandoned during the late Sapoá period, by ad 1200. Surprisingly, deep levels at Locus 2 included rich deposits of late Bagaces-­period pottery. Radiocarbon dates (see chapter 4, this volume) indicate that the site was first occupied about ad 600, and between ad 650 and ad 800 innovative pottery types such as Momta Polychrome were introduced. Sapoá pottery was adopted relatively quickly, perhaps between ad 750 and ad 850, with major changes in the types and vessel forms used. The brief exploratory excavations at Sonzapote, on Zapatera Island, encountered ceramics of both the late Tempisque period (ad 1–­300) and the Sapoá period. While it was hoped that carbon samples from a contaminated burial context with diagnostically early pottery would produce a date relating to the construction of Mound 14, and therefore a foundation date for the ceremonial center, the two 14C dates that were recovered (Beta-­441227, cal. ad 1020–­1160, and Beta-­441228, cal. ad 1165–­1265) apparently related to a later burial disturbance associated with Sacasa Striated urns and polychrome pottery (see chapter 14, this volume). Future investigations are planned to clarify the architectural history of the site. The University of Calgary projects have also provided radiocarbon dates for a late Tempisque-­period cemetery, Las Delicias, excavated in Managua by the Nicaraguan Institute of Culture. So far two 14C dates (Beta-­257989 and Beta-­ 257990, both ad 120–­380 calibrated, 2σ), provide a chronological framework for the Las Delicias cemetery, where nearly 100 individuals have been found in association with diagnostic pottery and objects of adornment. 134

M CCA F F ERT Y

Recent projects in Santa Isabel, Tepetate, El Rayo, and Sonzapote have contributed twenty-­eight radiocarbon dates relating to the late Bagaces through late Sapoá periods, ad 600–­1300 (see table 4.5 for a summary of dates from the first three projects). The results challenge the traditional ceramic sequence, especially relating to Ometepe-­period diagnostics, and clarify the Bagaces-­ Sapoá transition (ad 650–­850). Other projects are currently considering the Tempisque-­Bagaces transition. One possible result might be the creation of a chronological sequence for Pacific Nicaragua that is distinct from that of the Costa Rican section of Greater Nicoya, though recent work reported by Felipe Solís Del Vecchio (2017; see chapter 4, this volume) now suggests that chronology of the southern sector of Greater Nicoya has been similarly misinterpreted and that its chronological sequence may actually parallel the Nicaraguan sequence after all. FOODWAYS

Foodways are a fundamental means of inferring cultural identity, especially ethnicity (Twiss 2012). In order to identify possible migrant groups in Pacific Nicaragua that are consistent with the ethnohistorical sources, identifying archaeological foodways has been an important goal of the Calgary projects. Pacific Nicaragua has been described as an ecological paradise, with abundant wild plants and animals (Pohl and Healy 1980). This is reflected in the rich diversity of remains found in archaeological contexts, including fish; turtle; deer; armadillo; and assorted birds, reptiles, and mammals (Hoar 2006; López-­Forment Villa 2008). Fish provided the most abundant faunal remains found in well-­preserved contexts at Santa Isabel and El Rayo. Deer, however, made up the greatest meat contribution to the diet at Santa Isabel (López-­ Forment Villa 2008) but were less abundant at El Rayo. Archaeological evidence for hunting and fishing included some of the most abundant artifact classes (figure 6.6). This was particularly true of ceramic net weights used for fishing (Wilke 2011; see chapter 11, this volume). Bone fishhooks were also found at both Santa Isabel and El Rayo. Projectile points, probably used on spears for hunting large animals (or for warfare), were rare. More common were clay balls that were probably used as blowgun projectiles for hunting birds or iguanas. Excellent preservation of plant remains has included many examples of carbonized wood and seeds. Microscopic analysis of the carbonized wood is providing information on the pre-­Hispanic forest. The most numerous of the seeds are of jocote, a fruit that can be used to produce a fermented wine. Other T wenty Y ears of N icaraguan A rchaeology

135

Figure 6.6. Fishing and hunting tools. Photo by author.

seeds include beans, cacao, and coyol palm nuts. Notably, no evidence for maize has been found among the carbonized seeds, and analyses of phytoliths and starches from organic soil samples have also failed to recover evidence of maize. This is particularly curious since the alleged Postclassic migration should have included maize-­dependent Mexicans. Other artifacts relate to the preparation of food. Lithic scrapers were used to butcher game animals. Other stone tools were probably used to cut trees and dig for roots. Grinding stones were used for crushing seeds and fruits, though recent studies of phytoliths again indicate that maize was not processed (Dennett and Simpson 2010). Small chipped stone tools, raspaditas, feature a narrow, chipped tip where they may have been embedded in wooden grater boards for processing root crops such as manioc (figure 6.7; Debert 2005; Debert and Sheriff 2007). Notably, while thousands of raspaditas were found at Santa Isabel, they were scarce from Tepetate and El Rayo, suggesting that grater boards were not an important factor in food preparation in the Granada region. Large utilitarian vessels provide additional information on cooking practices. Cazuelas (open pots) were probably used for preparing stews with a mix of wild plants and animals. Shoe-­shaped pots may have been used for boiling liquids, with the condensed steam trapped in the upper surface of the vessel. Alternatively, these may have been used for extracting cacao oils to produce a drink akin to pinolillo (G. McCafferty and S. McCafferty 2012). A notable omission from the ceramic assemblage of Pacific Nicaragua are comals, the large griddles typically used for heating tortillas and found in abundance in 136

M CCA F F ERT Y

Figure 6.7. Chipped chert raspaditas. Photo by Brett Watson.

Postclassic sites of Central Mexico. The absence of comals was one of the first clues that the “out of Mexico” hypothesis may be flawed. Recent archaeological investigations provide extensive information on pre-­ Hispanic foodways, especially for the Sapoá period. People consumed diverse plants and animals obtained from the verdant natural environment. In contrast to expectations, however, almost no evidence was found of domesticated T wenty Y ears of N icaraguan A rchaeology

137

plants or animals: no evidence for maize, turkey, or dog was recovered among the hundreds of carbonized seeds or hundreds of thousands of faunal remains. Nutritional studies using stable isotope analysis are now in process to further clarify ancient dietary patterns (Waters-­Rist, Carroll, and G. McCafferty, forthcoming). ARCHITECTURE

Little information exists about pre-­Hispanic architectural forms of Nica­ ragua. In ethnohistorical sources from the sixteenth century (e.g., Fernández de Oviedo y Valdés [1526] 1950) there are descriptions of impermanent structures made of wattle-­and-­daub and with thatch roofs. Some of these had long, rectangular shapes similar to Indigenous buhíos from Panama (Steinbrenner 2010). Historical photos also provide images of Indigenous houses that likely resemble pre-­Hispanic forms. However, until the beginning of the University of Calgary projects, archaeological excavations had never recovered much information on architectural patterns from Pacific Nicaragua. The excavations at Santa Isabel investigated domestic remains in several artificial mounds; these were built up over time from collapsed architectural material and living debris, and there is little evidence of intentional mound construction. Although it was not possible to identify complete building “footprints,” numerous floors and wall fragments were encountered. The majority of the floors consisted of compacted sand, but in some cases there were also thin layers of a stucco-­like substance with a high phosphate content. On Mound 3, a sequence of eight superimposed floors spanning a time period of about 350  years suggests that each living surface was occupied for 30 to 50  years before the next construction episode. Walls were made of wattle-­and-­daub, a construction technique whereby thin poles are woven together and then covered in dried mud. Archaeological remains of wattle-­and-­daub architecture consists of chunks of burnt mud with impressions of the poles. The most complete mound at Tepetate had been badly looted in recent years. Excavations discovered remains of wall rubble in the upper levels of the structure, as well as stone slabs that may have comprised a covering for the mound. A meter below the surface was evidence of floors made of stone, associated with wall bases; a similar stone floor was also found at Locus 3. In another locus of the site a well-­preserved stone foundation wall was excavated (figure 6.8). A rare example of monumental architecture was found at Locus 2 at El Rayo: a stone wall of about 1 m in thickness that probably served as a retaining wall for an artificial platform. This was constructed during the late Bagaces period, 138

M CCA F F ERT Y

Figure 6.8. Stone foundation wall at Tepetate, Locus 2. Photo by author.

based on ceramics embedded in the wall. Additional architectural remains of residential structures were located on the Locus 2 platform, also associated with the Bagaces-­period occupation. A more impressive foundation wall was discovered during the 2015 season at Locus 4 (Rice 2019). A double row of vertical lajas formed the north wall of a large structure that measured about 25 m × 10 m (figure 6.9). Vestiges of a similar foundation wall were visible on the south side, but the shorter east and west sides lacked the vertical stone foundation. It is possible that the double row of stones, separated by about 30 cm, may have supported a perishable wall, perhaps a palisade of tree trunks. Also associated with the north wall were several large stones that may have been vertical monoliths. Three superimposed walking surfaces were associated with the platform, with the lowest consisting of a well-­smoothed surface of talpuja, a mix of volcanic ash and sand. Few artifacts were found with this structure, which apparently dates to the Sapoá period based on diagnostic ceramics, leading to the interpretation that this was a civic-­ceremonial structure that may have been used in rituals associated with the mortuary complex. A small structure, represented by stone rubble and measuring about 1 m × 2 m, was located at Locus 3 at El Rayo. This structure, which may have been a T wenty Y ears of N icaraguan A rchaeology

139

Figure 6.9. Double-­row foundation from El Rayo, Locus 4, including upright stone. Photo by author.

small shrine or altar, was adjacent to a cluster of shoe-­shaped Sacasa urns and other possible offerings, including scattered human skeletal remains. Although minimal, the architectural evidence uncovered in these projects is among the first ever discovered in Nicaragua. It indicates that the Indigenous population lived in permanent houses and in long-­term settlements. The presence of a platform wall and mortuary structure at El Rayo implies a degree of political organization whereby the inhabitants may have worked together for the public good, probably relating to ritual activities. In contrast to these rudimentary architectural forms, excavations at the Sonzapote site mapped and exposed much more substantial structures. These 140

M CCA F F ERT Y

Figure 6.10. Masonry construction at Sonzapote. Photo by author.

were constructed of dry laid masonry blocks of igneous rock (abundant at the site) and measured 1–­3 m in height (figure 6.10). Some evidence of corners and one possible staircase were found in association with rectangular mounds, and alignments on top of the low platforms indicate superstructures. Some of the mounds were arranged around possible plazas. Bovallius’s (1886) nineteenth-­ century sketch map of the site indicated the arrangement of monumental sculptures around the periphery of mounds, and the Calgary monument inventory recorded both statues and petroglyphs in association with mounds. Excavations at Mound 14 encountered two superimposed walking surfaces abutting the building foundation, and associated ceramics tentatively date the structure to the late Tempisque period (ad 1–­300). Domestic refuse was rare in association with the mounds, suggesting that they did not serve a residential function. Instead, the Sonzapote mounds were probably civic-­ceremonial in nature, and the dense clustering and apparent site planning suggest a degree of incipient urbanism at this very early date. MORTUARY PATTERNS

Recent archaeological investigations in Pacific Nicaragua have encountered a range of mortuary patterns dating to the late Tempisque, late Bagaces, and Sapoá periods (ad 1–­1300). These periods encompass the supposed arrival of Mesoamerican migrants, as asserted in historical accounts, and so relate to cultural changes (see chapter 14, this volume, for more detailed discussion). T wenty Y ears of N icaraguan A rchaeology

141

Figure 6.11. Sacasa Striated shoe-­pots typical of Sapoá-­period mortuary practices. Photo by author.

The best-­known burial form for the Postclassic period involves large shoe-­ pots (figure 6.11). At Santa Isabel, these pots were consistently associated with infant burials, generally buried without grave goods. In contrast, one adult and two adolescents were also found buried directly in the soil. The adult (male) was buried with pieces of greenstone, lapidary tools, and an unusual pot decorated with multiple owl faces. Near the adult and with a parallel orientation, a subadult of seven-­to-­nine years of age was buried with its head resting on a turtle shell. At Tepetate, clusters of crushed burial urns, including shoe-­pots and large-­ mouth ollas, were found with poorly preserved human skeletal remains. All identifiable individuals were adults. Some were buried in urns, but others were placed in extended position above the vessels. Associated grave goods included bowls, miniature vessels with appliqué faces, and ornamentation. A possible burial crypt was found within Mound 1, with alignments of flat stones delimiting a rectangular space. A similar rectangular crypt was excavated at La Chureca (Zambrana Fernández 2012), and crypts of vertical stones have been identified at El Rayo. A single human tooth was the only skeletal 142

M CCA F F ERT Y

Figure 6.12. Fragmentary human skull (left) with the bifacial knife found embedded in its mouth (right), El Rayo Locus 1. Photos by author.

remain found in the Tepetate crypt, in what was a badly disturbed context with abundant evidence of looting. The El Rayo site featured burials from both the Bagaces-­and Sapoá-­period components, and therefore presents excellent evidence for changing burial patterns as well as osteoarchaeological analysis (see chapter 15, this volume). The main cemetery was exposed in a roadcut, revealing burial urns and human remains. Several clusters of shoe-­pot urns were found containing miniature vessels and cobbles of volcanic rock but only occasional human skeletal remains. Instead, isolated human crania were found adjacent to the “burial” urns as possible evidence for trophy heads. One exception was a disintegrating human skull found in an urn and with a chert biface placed in the mouth (figure 6.12). A cluster of three urns was associated with a grouping of well-­formed biface knives, two large earspools, and a small bowl filled with at least eighty beads. Beneath the level of Sapoá urns were primary burials in flexed position associated with ceramics from the Bagaces period. One was associated with a ground stone mano, and another was buried with a spindle whorl. Unfortunately, the skeletal remains were too poorly preserved for sexing of the individuals. Near this cluster of primary burials were three complete vessels: one with modeled and incised decoration in the form of a nude female, another with a constricted neck, and one with modeled and incised decoration to form a rodent face. This “rodent vessel” contained fragmented human skeletal remains of several individuals. At Locus 2, three individuals were found relating to the Bagaces occupation. A fetus was found in a deep deposit, associated with domestic refuse. An adult T wenty Y ears of N icaraguan A rchaeology

143

Figure 6.13. Burial cluster associated with shrine at El Rayo Locus 3. Photo by author.

was placed in an extended, dorsal position, while another adult, probably an elderly female, was in a flexed position. Grave goods were not associated with these burials. A Sacasa Striated shoe-­pot urn was also found, relating to the Sapoá period, but no skeletal remains were found. Locus 3 featured additional shoe-­pot urns, aligned in a north-­south orientation in front of the possible shrine (figure 6.13). Human remains were again rare in these vessels, but bones were found scattered around the urns, together with smaller vessels. One of the Sacasa shoe-­pots did contain remains of a nearly complete, semidisarticulated subadult placed in the vessel on top of the mandible of an adult. In another area, poorly preserved skeletal remains were found associated with several complete vessels, a ceramic ocarina, weaving tools of bone, and a copper bell; this interment was secondary but placed directly in the ground. In sum, recent excavations have recovered a variety of mortuary practices. Late Tempisque–­period burials from Las Delicias (Managua) were typically primary burials associated with ceramics and occasionally ornamentation as grave goods. During the late Bagaces period, interments were also primary, with skeletons in flexed or extended positions. Shoe-­shaped urns were typical 144

M CCA F F ERT Y

of the Sapoá period, though how human remains were placed in or around these vessels was variable. These results differ from the pattern found at Santa Isabel, where infants were consistently buried in the ovoid vessels and adults and adolescents were primary burials. Mortuary offerings were also inconsistent, with miniature vessels found in the Granada region. At El Rayo, cobbles of volcanic rock were often placed in shoe-­pots, and the interior bases of the pots were occasionally lined with broken potsherds. Some of the urns seem to have been placed one inside another, forming concentric deposits in cross section. ORNAMENTATION AND IDENTIT Y

One of the primary goals of research in Pacific Nicaragua has been the interpretation of cultural identities, especially ethnicity. Ornamentation is one aspect of material culture that is most directly related to identity (G. McCafferty and S. McCafferty 2009, 2011). In Nicaragua ornamentation consists of such items as beads, pendants, and earspools, made of greenstone, bone, shell, and ceramic (figure 6.14). Other ornamentation of perishable materials included textiles, featherwork, and body stamping. These can be inferred from the polychrome figurines that are common in archaeological contexts as well as actual objects of adornment. Beads were one of the most numerous objects of ornamentation found. Most were made of ceramic, but bone and greenstone were also used. At El Rayo, a small bowl was found with at least eighty ceramic and bone beads, probably as a burial offering. Greenstone beads were found at all three sites; blanks and misdrilled pieces indicate that beads were being manufactured at Santa Isabel. A large ceramic sphere incised with the face of the Mexican storm god, Tlaloc, as characterized by large goggle eyes and fangs, was found at Santa Isabel. Six greenstone beads were the only grave goods found associated with a Sapoá-­ period burial at Mound 14 at Sonzapote. Pendants were made of clay, bone, shell, and greenstone. The most elaborate pendants were carved from bone: one from Santa Isabel depicted serpents and avian species, while another probably represented the mandible of a crocodilian. A skeletal pendant was found at El Rayo. Perforated teeth—­including those of peccary, jaguar, and shark—­were found at all sites; perforated human teeth were only found at Santa Isabel and may be related to the practice of making bracelets of teeth found in northwestern Costa Rica (see chapter 16, this volume, for examples). Shell was most common at Santa Isabel, where there was evidence for manufacture using marine shell. Greenstone pendants were also found at Santa Isabel, again with evidence of manufacture. At Santa T wenty Y ears of N icaraguan A rchaeology

145

Figure 6.14. Objects of adornment. Photos by author.

Isabel a tiny native copper human figure pendant was recovered, and at El Rayo a copper bell was found—­rare examples of metal artifacts. These objects may represent a long tradition of exotic adornment, as a greenstone avian pendant was found at the late Tempisque cemetery of Las Delicias, as well as a metal bell. The most abundant form of pendant was made of reworked potsherds, at least at Santa Isabel, where several hundred perforated worked sherds were found. However, these were quite rare at Tepetate and El Rayo and so may indicate a localized form of ornamentation. Most earspools were made of fine clay, burnished to a brownish-­black color. The most common form consisted of hollow cylinders with thin, concave walls. Some earspools were more elaborate, solid with incised decorations in a cross 146

M CCA F F ERT Y

Figure 6.15. Figurines. Drawings by Sharisse McCafferty. Photos by author.

pattern. Others were made from fish vertebrae. The variation is size, from 1 cm to 5 cm in diameter, may relate to age grades or status. Unfortunately, no earspools have been found with aged or sexed skeletons. Figurines, as miniature portraits, provide valuable insight into the way that ancient Nicaraguans viewed themselves, while also including details of ornamentation that do not preserve archaeologically, such as hairstyles, tattoos, body paint, and clothing (figure 6.15). Careful analysis of figurines, along with other aspects of ornamentation, offers important information on the aesthetics of concepts relating to the “body beautiful” (G. McCafferty and S. McCafferty 2009, 2011; Reischer and Koo 2004). From the perspective of diachronic change between the Bagaces and Sapoá periods, when foreign groups allegedly migrated into the region, figurines show dramatic differences (Leullier-­Snedeker 2013). Bagaces figurines from El Rayo are generally monochrome female figures with prominent bellies and buttocks. Sapoá figurines are also typically female but are painted with representations of clothing, T wenty Y ears of N icaraguan A rchaeology

147

body paint, and occasional ornamentation. Interestingly, at Santa Isabel both monochrome and polychrome figurines were found together, as possible evidence of cultural mixing during the Sapoá period. SPECIALIZED CRAF T PRODUCTION

One of the fundamental characteristics of complex societies is the specialized production of materials. This activity implies the allocation of time and energy for labor apart from basic subsistence. It also suggests a degree of social hierarchy and long-­distance trade of precious commodities (Helms 1992). Investigations in Pacific Nicaragua, and particularly at Santa Isabel, have discovered a variety of specialized craft activities. The most abundant evidence of specialized production is in the form of thousands of fragments of pottery, many of which were decorated with painted and engraved motifs. Identifying ceramic potting traditions using detailed decorative and morphological characteristics was the theme of Larry Steinbrenner’s PhD dissertation (2010; see also chapter 9, this volume). He inferred several diachronic trajectories as different polychrome types evolved through time. Ceramic manufacture using compositional analyses was the topic of Carrie Dennett’s more recently completed doctoral research (Dennett 2016; see also chapter 10, this volume), which built on the foundational study by Bishop and Lange (Bishop et al. 1992; Bishop, Lange, and Lange 1988; Bishop and Lange 2013). Thin sections of different ceramic types were studied microscopically to identify different clay matrices, representing distinct production centers. This analysis has also been supplemented by instrumental neutron activation analysis (INAA) using the Smithsonian Institution’s nuclear facilities, and also with X-ray diffraction analysis (G. McCafferty, Logee, and Steinbrenner 2007). Another prominent example of specialized production is found in thousands of pieces of chipped stone, a by-­product of stone tool production. The majority of lithic materials used along the shore of Lake Cocibolca were of a whitish chert, but a dark red chert, chalcedony, and obsidian were also used. Some tools may have arrived premade, but the abundance of flakes of all materials indicates local manufacture as well. Obsidian was more common at Tepetate, where it represented about 8 percent of the total lithic assemblage. It was most often found as prismatic blade fragments, a common Mesoamerican form. One expended core was found, as evidence that at least some of the blades may have been produced locally. Obsidian was much less common at El Rayo or Santa Isabel, at about 1 percent of the total lithic assemblage, as an indication that these sites were less involved in consumption of this rare commodity. 148

M CCA F F ERT Y

Figure 6.16. Monumental art from Sonzapote. Photos by author.

While the majority of the chipped stone tools recovered were expedient flakes or crudely made bifaces, a high level of craft ability was represented in bifacial knives recovered at El Rayo. These were found in association with isolated crania and therefore may be indicative of decapitation ritual and trophy heads. Ground stone objects were also present, including manos and metates. These were made of basalt and andesite. A few examples featured carved decoration, though never to the elaboration of the flying-­panel metates best known from Costa Rica. A residue analysis of ground stone tools from Santa Isabel recovered evidence of fruit starches (Dennett and Simpson 2010), but there was no indication that they were ever used for maize. A more impressive form of ground stone production was found at Sonzapote in the form of the monumental sculptures and petroglyphs (figure 6.16). These were likely produced on-­site, using the abundantly available natural stone. Although they are currently impossible to date, the relationship of the statues and petroglyphs to the masonry mounds suggests that they were produced in the late Tempisque period. T wenty Y ears of N icaraguan A rchaeology

149

Figure 6.17. Spinning and weaving tools. Photos by author.

Evidence for textile production was abundant at Santa Isabel in the form of spindle whorls, as well as a variety of weaving tools of bone (S. McCafferty and G. McCafferty 2008). In contrast, whorls were relatively rare at Tepetate and El Rayo, indicating that they probably imported thread from production centers such as Santa Isabel. Due to the exceptional preservation conditions at Santa Isabel and El Rayo, numerous weaving tools of bone were recovered (figure 6.17). These are classified as needles, picks, and battens; some of the small 150

M CCA F F ERT Y

Figure 6.18. Evidence of greenstone production. Photos by author.

picks were carved with decorative patterns. A few textile tools from Santa Isabel (two whorls and several batten fragments) were made out of polished greenstone, suggesting an ideological significance for these objects. Painted figurines depict netted blouses and possible skirts as evidence for the kinds of clothing produced, but thread was probably also used for making fishing nets and hammocks. Whereas textile production is generally associated with female production in Mesoamerica, no secure relationship has been established between spinning and weaving and gender identity in pre-­Columbian Pacific Nicaragua. Further evidence for specialized production from Santa Isabel involves polished greenstone, which takes on a luster that has resulted in it being called “social jade”—­not true jade, but rather a local imitation (Lange 1993) (figure 6.18). Santa Isabel featured evidence for the production of social jade beads, pendants, and amulets using locally available stone. Raw materials showed evidence of string-­saw cut marks as the rough forms were shaped, as well as wasters that were discarded during manufacture. The exceptional preservation of bone at Santa Isabel and El Rayo included examples of bone tools such as fishhooks and needles. At Santa Isabel these remains also included evidence of their production. For example, long bones of deer had cut marks from string saws, probably for the production of fishhooks (figures 6.6 and 6.19). Other bones were modeled to produce weaving tools and ornamentation. At Santa Isabel there were various examples of shell jewelry and also waste from marine shell cores as evidence of manufacture. The abundance of these T wenty Y ears of N icaraguan A rchaeology

151

Figure 6.19. Deer long bones sawed for fishhook production. Photos by author.

cores implies production for exchange. It is interesting, therefore, that there were no examples of shell jewelry at either Tepetate or El Rayo. CONCLUSION

Excavations in Santa Isabel, Tepetate, El Rayo, and Sonzapote represent the most extensive archaeological program ever conducted in Nicaragua and have produced many new discoveries and interpretations. We are in the process of “reinventing” the prehistory of the country based on scientific data, in contrast to the mythical histories that have previously been the basis of interpretation. The University of Calgary projects demonstrate the value of scientific research and reveal the diversity and complexity of Indigenous Nicaraguan culture. Although the inhabitants lived in simple houses, they enjoyed the abundance available from the land and the lake. They practiced mortuary rituals in a variety of ways, perhaps relating to different regional and social identities. The beautifully decorated pottery demonstrates highly developed artistic abilities. Ornamentation and specialized production indicate the complexity of their social organization as well as interaction with distant regions. The overarching goal of our recent research in Pacific Nicaragua has been to evaluate, archaeologically, the ethnohistorically based “out of Mexico” 152

M CCA F F ERT Y

Figure 6.20. Feathered serpent and Ehecatl iconography from Pacific Nicaragua. Photos by author.

hypothesis that Nicarao and Chorotega migrants from Central Mexico settled the region during the Postclassic period. Two of the principal centers of Indigenous culture, Santa Isabel and Tepetate, were specifically selected for investigation because settlement surveys of the region by Niemel (2003) and Salgado González (1996) had identified them as being prominent at the time of Spanish contact. The Sonzapote site, long believed to have been a Chorotega ceremonial center (Lothrop 1921), was also tested, albeit with new evidence that the major architectural and monument program predated the Chorotega occupation. Revised radiocarbon chronologies now indicate that at least in the areas sampled, these sites were abandoned several centuries prior to European contact, and so the final centuries of pre-­Hispanic life have not yet been explored. Nevertheless, the Sapoá period, when alleged Otomanguean-­speaking Chorotega migrants settled in the area, is abundantly represented. Archaeo­ logical evidence now presents an ample record of daily life and mortuary ritual. At El Rayo in particular, the Bagaces-­Sapoá transition shows diachronic change occurring between ad 750 and ad 850, consistent with the arrival of foreign innovations in material culture. Some symbolic elements, such as the use of the “feathered serpent” and Ehecatl wind god vessel supports, indicate affiliation with Mexican religious cults (figure 6.20; Manion and McCafferty 2013; G. McCafferty 2019). However, while there are significant changes in pottery, figurine styles, and mortuary practices at the time of transition, there are other elements traditionally associated with T wenty Y ears of N icaraguan A rchaeology

153

Mesoamerican culture that are lacking, especially the use of maize, dog, and turkey in the diet and the use of incense burners in ritual practice. Lack of these fundamental traits makes it impossible to casually affirm the presence of Mexican migrants (G. McCafferty 2015a), and instead we must consider a less direct form of cultural change, involving ethnic groups from intermediate regions that may have had more contact with, or incentive to adopt, Mesoamerican religious ideology. Thus we are currently leaning away from an “out of Mexico” scenario to consider other Central American regions in El Salvador and/or Honduras as origins of migration (G. McCafferty and Dennett 2013). Research over the past twenty years has produced a material record useful for characterizing ancient Chorotega ethnic identity (G. McCafferty 2011). Despite this solid foundation, however, more questions remain, requiring additional research. Paramount is the problem of the Nicarao of the late Postclassic Ometepe period. No Nicarao sites have been significantly investigated, and many of what were thought to be Nicarao diagnostic artifacts are now recognized as having been introduced hundreds of years earlier, probably by the Chorotega themselves. Similarly, the Bagaces-­Sapoá transition remains key to understanding cultural changes brought about by possible migrant groups, perhaps deriving from Central Mexico (following the ethnohistorical accounts) or from intermediate zones in El Salvador or Honduras (as is now suggested by ceramic associations from El Rayo). Further research is currently being planned, and hopefully the next decade will produce additional advances to address these questions. REFERENCES

Abel-­V idor, Suzanne. 1981. “Ethnohistorical Approaches to the Archaeology of Greater Nicoya.” In Between Continents / Between Seas: Precolumbian Art of Costa Rica, edited by Elizabeth P. Benson, 85–­92. New York: Harry N. Abrams. Bishop, Ronald L., and Frederick W. Lange. 2013. “Frederick R. Mayer’s Legacy of Research Support: The Prehispanic Ceramic Schools of Greater Nicoya.” In Pre-­ Columbian Art and Archaeology: Essays in Honor of Frederick R. Mayer, edited by Margaret Young-­Sánchez, 27–­46. Denver: Mayer Center for Pre-­Columbian and Spanish Colonial Art, Denver Art Museum. Bishop, Ronald L., Frederick W. Lange, Suzanne Abel-­V idor, and Peter C. Lange. 1992. “Compositional Characterization of the Nicaraguan Ceramic Sample.” In The Archaeology of Pacific Nicaragua, edited by Frederick W. Lange, Payson D.

154

M CCA F F ERT Y

Sheets, Aníbal Martínez, and Suzanne Abel-­V idor, 135–­162. Albuquerque: University of New Mexico Press. Bishop, Ronald L., Frederick W. Lange, and Peter C. Lange. 1988. “Ceramic Paste Compositional Patterns in Greater Nicoya Pottery.” In Costa Rican Art and Archaeology: Essays in Honor of Frederick R. Mayer, edited by Frederick W. Lange, 11–­44. Boulder: University of Colorado.

Bonilla Vargas, Leidy, Marlin Calvo Mora, Juan Vicente Guerrero Miranda, Silvia Salgado González, and Frederick W. Lange. 1990. “La cerámica de la Gran Nicoya.” Vínculos: Revista de Antropología del Museo Nacional de Costa Rica 13(1–­2 [1987]):1–­327.

Bovallius, Carl. 1886. Nicaraguan Antiquities. Stockholm: Swedish Society of Anthropology and Geography. Bransford, John F. 1881. Archaeological Researches in Nicaragua, Smithsonian Institution Contributions to Knowledge 383. Washington, DC: The Smithsonian Institution.

Carmack, Robert M., and Silvia Salgado González. 2006. “A World-­Systems Perspective on the Archaeology and Ethnohistory of the Mesoamerican / Lower Central American Border.” Ancient Mesoamerica 17(2):219–­229.

Castillo Barquero, Julio Magdiel. 1989. “The Context and Meaning of the Zapatera Sculptures: Punta del Zapote. Mound I.” MA thesis, University of Texas, Austin.

Chapman, Anne M. 1974. Los Nicarao y los Chorotega según las fuentes históricas. 2nd ed. San José: Editorial de la Universidad de Costa Rica.

Debert, Jolene. 2005. “Raspadita: A New Lithic Tool Type from Santa Isabel, Nicaragua.” MA thesis, University of Manitoba, Winnipeg. Debert, Jolene, and Barbara L. Sherriff. 2007. “Raspadita: A New Lithic Tool from the Isthmus of Rivas, Nicaragua.” Journal of Archaeological Science 34(11):1889–­1901. Dennett, Carrie L. 2016. “The Ceramic Economy of Pre-­Columbian Pacific Nicaragua (ad 1–­1250).” PhD diss., University of Calgary.

Dennett, Carrie L., and Steven W. Simpson. 2010. “Tools of the Trade: Macro­ botanical Trace Analyses of Ground Stone Tools from El Rayo, Nicaragua.” Poster presented at the 75th Annual Meeting of the Society for American Archaeology, St. Louis, Missouri, April 14–­­18.

Espinoza Pérez, Edgar, Ramiro García Vásquez, and Fumiyo Suganuma. 1999. Rescate arqueológico en el sitio San Pedro, Malacatoya, Granada, Nicaragua. Managua: Departamento de Investigaciones Antropológicas, Museo Nacional de Nicaragua, Instituto Nicaragüense de Cultura.

T wenty Y ears of N icaraguan A rchaeology

155

Fernández de Oviedo y Valdés, Gonzalo. (1526) 1950. Sumario de la natural historia de las Indias. Edited by José Miranda. Biblioteca Americana. Serie de Cronistas de Indias 13. Mexico City: Fondo de Cultura Económica. Fowler, William R., Jr. 1989. The Cultural Evolution of Ancient Nahua Civilizations: The Pipil-­Nicarao of Central America. Civilization of the American Indian Series 194. Norman: University of Oklahoma Press. Guido Martínez, Clemente. 2004. Los dioses vencidos de Zapatera: Mitos y realidades. Managua: Academia Nicaragüense de la Lengua. Haberland, Wolfgang. 1992. “The Culture History of Ometepe Island: Preliminary Sketch (Survey and Excavations, 1962–­1963).” In The Archaeology of Pacific Nicaragua, edited by Frederick W. Lange, Payson D. Sheets, Aníbal Martínez, and Suzanne Abel-­Vidor, 63–­117. Albuquerque: University of New Mexico Press. Healy, Paul F. 1980. Archaeology of the Rivas Region, Nicaragua. Waterloo, ON: Wilfrid Laurier University Press. Healy, Paul F. 1988. “Greater Nicoya and Mesoamerica: Analysis of Selected Ceramics.” In Costa Rican Art and Archaeology: Essays in Honor of Frederick R. Mayer, edited by Frederick W. Lange, 293–­301. Boulder: University of Colorado Press. Helms, Mary W. 1992. “Thoughts on Public Symbols and Distant Domains Relevant to the Chiefdoms of Lower Central America.” In Wealth and Hierarchy in the Intermediate Area: A Symposium at Dumbarton Oaks, 10th and 11th October 1987, edited by Frederick W. Lange, 317–­330. Washington, DC: Dumbarton Oaks. Hoar, Bryanne. 2006. “The Isthmus of Plenty: Faunal Analysis for the site of Santa Isabel ‘A,’ Rivas, Nicaragua.” Honour’s thesis, University of Calgary. Hoopes, John W., and Geoffrey G. McCafferty. 1989. “Out of Mexico: An Archaeological Evaluation of the Migration Legends of Greater Nicoya.” Paper presented at the 54th Annual Meeting of the Society for American Archaeology, Atlanta, Georgia, April 5–­­9. Lange, Frederick W. 1993. “Formal Classification of Prehistoric Costa Rican Jade: A First Approximation.” In Precolumbian Jade: New Geological and Cultural Interpretations, edited by Frederick W. Lange, 269–­288. Salt Lake City: University of Utah Press. Lange, Frederick W. 1994. “Evaluación histórica del concepto Gran Nicoya.” Vínculos: Revista de Antropología del Museo Nacional de Costa Rica 18–­19(1–­2 [1992–­1993]):1–­8. Lange, Frederick W., ed. 1995. Descubriendo las huellas de nuestros antepasados: El proyecto “Arqueología de la Zona Metropolitana de Managua.” Managua: Alcaldía de Managua. 156

M CCA F F ERT Y

Lange, Frederick W., ed. 1996a. Abundante cooperación vecinal: La segunda temporada del proyecto “Arqueología de la Zona Metropolitana de Managua.” Managua: Alcaldía de Managua. Lange, Frederick W. 1996b. “Gaps in Our Databases and Blanks in Our Syntheses: The Potential for Central American Archaeology in the Twenty-­First Century.” In Paths to Central American Prehistory, edited by Frederick W. Lange, 305–­326. Niwot: University Press of Colorado.

León-­Portilla, Miguel. 1972. Religión de los nicaraos: Análisis y comparación de tradiciones culturales nahuas. Serie de Cultura Náhuatl. Mexico City: Instituto de Investigaciones Históricas, Universidad Nacional Autónoma de México. Leullier-­Snedeker, Natasha. 2013. “Changing Identities in Changing Times: Gendered Roles and Representations through the Ceramic Figurines of Greater Nicoya.” MA thesis, University of Calgary. López-­Forment Villa, Angélica. 2008. “Aprovechamiento cultural de los recursos faunísticos en el sitio de Santa Isabel, Nicaragua.” Licentiate’s thesis, Instituto Nacional de Antropología e Historia, Mexico City.

Lothrop, Samuel K. 1921. “The Stone Statues of Nicaragua.” American Anthropologist 23(3):311–­319.

Lothrop, Samuel K. 1926. Pottery of Costa Rica and Nicaragua. 2 vols. Memoir 8. New York: Museum of the American Indian, Heye Foundation. Manion, Jessica. 2016. “Remembering the Ancestors: Mortuary Practices and Social Memory in Pacific Nicaragua.” MA thesis, University of Calgary. Manion, Jessica, and Geoffrey G. McCafferty. 2013. “Serpientes emplumadas del Pacífico de Nicaragua / Feathered Serpents of Pacific Nicaragua.” Mi Museo y Vos 26:19–­24.

McCafferty, Geoffrey G. 2008. “Domestic Practice in Postclassic Santa Isabel, Nicaragua.” Latin American Antiquity 19(1):64–­82. McCafferty, Geoffrey G. 2011. “Etnicidad chorotega en la frontera sur de Meso­ américa.” La Universidad: Órgano científico-­sociocultural de la Universidad de El Salvador 14(April–­June):91–­112.

McCafferty, Geoffrey G. 2015a. “Evaluación arqueológica del mito histórico ‘Fuera de México’ / Archaeological Evaluation of the ‘Out of Mexico’ Historical Myth.” Mi Museo y Vos 32:21–­30.

McCafferty, Geoffrey G. 2015b. “The Mexican Legacy in Nicaragua, or Problems When Data Behave Badly.” In Constructing Legacies of Mesoamerica: Archaeological Practice and the Politics of Heritage in and Beyond Mexico, edited by David S. Anderson, Dylan J. Clark, and J. Heath Anderson, 110–­118. Archaeological Papers of

T wenty Y ears of N icaraguan A rchaeology

157

the American Anthropological Association, Special Issue, vol. 25, no. 1. Hoboken: American Anthropological Association, John Wiley and Sons.

McCafferty, Geoffrey G. 2019. “Mixteca-­Puebla Style Ceramics from Early Post­ classic Pacific Nicaragua.” Mexicon: Journal of Mesoamerican Studies 41 (3):77–­83. McCafferty, Geoffrey G., and Carrie L. Dennett. 2013. “Ethnogenesis and Hybridity in Proto-­Historic Nicaragua.” Archaeological Review from Cambridge 28(1):191–­215. McCafferty, Geoffrey G., Jillian Logee, and Larry Steinbrenner. 2007. “X-Ray Diffraction Analysis of Greater Nicoya Ceramics.” La Tinaja: Newsletter for Archaeological Studies 18(2):12–­17. McCafferty, Geoffrey G., and Sharisse D. McCafferty. 2009. “Crafting the Body Beautiful: Performing Social Identity at Santa Isabel, Nicaragua.” In Mesoamerican Figurines: Small-­Scale Indices of Large-­Scale Social Phenomena, edited by Christina T. Halperin, Katherine A. Faust, Rhonda Taube, and Aurore Giguet, 183–­204. Gainesville: University Press of Florida. McCafferty, Geoffrey G., and Sharisse D. McCafferty. 2011. “Bling and Things: Ornamentation and Identity in Pacific Nicaragua.” In Identity Crisis: Archaeological Perspectives on Social Identity. In Proceedings of the 42nd Annual Chacmool Archaeology Conference, edited by Lindsey Amundsen-­Meyer, Nicole Engel, and Sean Pickering, 243–­252. Calgary: Chacmool Archaeology Association, University of Calgary. McCafferty, Geoffrey G., and Sharisse D. McCafferty. 2012. “Ollas en forma de zapato tipo Sacasa Estriado: Función y significado.” Mi Museo y Vos 20:5–­17. McCafferty, Geoffrey G., Oscar Pavón Sánchez, and Ligia M. Galeano Rueda. 2013. “Investigaciones preliminares en Sonzapote, Isla Zapatera / Preliminary Investigations at Sonzapote, Zapatera Island.” Mi Museo y Vos 26:5–­13. McCafferty, Geoffrey G., Silvia Salgado González, and Carrie L. Dennett. 2009. “¿Cuándo llegaron los mexicanos? La transición entre los periodos Bagaces y Sapoá en Granada, Nicaragua.” In Proceedings of the III Congreso Centroamericano de Arqueología en El Salvador, October 2009. San Salvador, El Salvador: Museo Nacional de Antropología “Dr. David J. Guzmán.” McCafferty, Geoffrey G., and Larry Steinbrenner. 2005a. “Chronological Implications for Greater Nicoya from the Santa Isabel Project, Nicaragua.” Ancient Mesoamerica 16(1):131–­146. McCafferty, Geoffrey G., and Larry Steinbrenner. 2005b. “The Meaning of the Mixteca-­Puebla Stylistic Tradition: The View from Nicaragua.” In Art for Archaeology’s Sake: Material Culture and Style across the Disciplines. Proceedings of the 33rd Annual Chacmool Archaeology Conference, edited by Andrea Waters-­Rist, Christine Cluney, Calla McNamee, and Larry Steinbrenner, 282–­292. Calgary: Chacmool Archaeology Association, University of Calgary.

158

M CCA F F ERT Y

McCafferty, Sharisse D., and Geoffrey G. McCafferty. 2008. “Spinning and Weaving Tools from Santa Isabel, Nicaragua.” Ancient Mesoamerica 19(1):143–­156.

Motolinía (Toribio de Benavente). (1540) 1951. Motolinía’s History of the Indians of New Spain Translated and edited by Francis Borgia Steck. Publications of the Academy of American Franciscan History. Documentary Series 1. Washington, DC: Academy of American Franciscan History. Navarro Genie, Rigoberto. 2007. “Les sculptures préhispaniques en pierre du versant Pacifique du Nicaragua et du nord-­ouest du Costa Rica et leur contexte archéologique (650–­1830 apr. J.-­C.).” PhD diss., Université de Paris I, Panthéon-­Sorbonne.

Niemel, Karen S. 2003. “Social Change and Migration in the Rivas Region, Pacific Nicaragua (1000 bc–­a d 1522).” PhD diss., State University of New York, Buffalo.

Niemel, Karen S., Manuel A. Román Lacayo, and Silvia Salgado González. 1998. “Las secuencias cerámicas de los periodos Sapoá (800–­1350 dC) y Ometepe (1350–­1522 dC) en el Pacífico Sur de Nicaragua.” In XI Simposio de Investigaciones Arqueológicas en Guatemala, 1997, edited by Juan Pedro Laporte and Héctor L. Escobedo, 790–­798. Guatemala City: Museo Nacional de Arqueología y Etnología, Instituto de Antología e Historia, Ministerio de Cultura y Deportes y Asociación Tikal. Norweb, Albert Holden. 1964. “Ceramic Stratigraphy in Southwestern Nicaragua.” Proceedings of the 35th International Congress of Americanists 1, 551–­561. Mexico City, August 19–­­25, 1962.

Pohl, Mary, and Paul F. Healy. 1980. “ ‘Mohammed’s Paradise’: The Exploitation of Faunal Resources in the Rivas Region of Nicaragua.” In Archaeology of the Rivas Region, Nicaragua, edited by Paul F. Healy, 287–­292. Waterloo, ON: Wilfrid Laurier University Press.

Reischer, Erica, and Kathryn S. Koo. 2004. “The Body Beautiful: Symbolism and Agency in the Social World.” Annual Review of Anthropology 33(1):297–­317.

Rice, Shaelyn J. 2019. “Constructing the Past: Monumentality of the Bagaces and Sapoá Periods at the Site of El Rayo, Granada, Nicaragua.” MA thesis, University of Calgary. Román Lacayo, Manuel A. 2013. “Social and Environmental Risk and the Development of Social Complexity in Precolumbian Masaya, Nicaragua.” PhD diss., University of Pittsburgh. Salgado González, Silvia. 1996. “Social Change in a Region of Granada, Pacific Nicaragua (1000 B.C.–­1522 A.D.).” PhD diss., State University of New York, Albany.

T wenty Y ears of N icaraguan A rchaeology

159

Salgado González, Silvia, and Ricardo Vázquez Leiva. 2006. “Was There a Greater Nicoya Subarea during the Postclassic?” Vínculos: Revista de Antropología del Museo Nacional de Costa Rica 29(1–­2):1–­16. Shea, Thomas, Benjamin van Wyk de Vries, and Martín Pilato. 2008. “Emplacement Mechanisms of Contrasting Debris Avalanches at Volcán Mombacho (Nicaragua), Provided by Structural and Facies Analysis.” Bulletin of Volcanology 70(8):899–­921.

Solís Del Vecchio, Luis Felipe. 2017. “Bahía Culebra en el noroeste de Costa Rica: ¿Tuvo ocupación prehispánica durante el Período Ometepe (1350–­1550 d.C.)?” Paper presented at the XI Congreso de la Red Centroamericana de Antropología, University of Costa Rica, San José, February 27–­­March 3. Squier, Ephraim G. 1852. Nicaragua: Its People, Scenery, Monuments, and the Proposed Interoceanic Canal. 2 vols. New York: D. Appleton and Co.

Squier, Ephraim G. 1853. “Observations on the Archaeology and Ethnology of Nicaragua.” Transactions of the American Ethnological Society 3(1):85–­158.

Stansell, Nathan D. 2013. “Radiocarbon Ages for the Timing of Debris Avalanches at Mombacho Volcano, Nicaragua.” Bulletin of Volcanology 75(1):686. Steinbrenner, Larry. 2002. “Ethnicity and Ceramics in Rivas, Nicaragua, ad 800–­1550.” MA thesis, University of Calgary.

Steinbrenner, Larry. 2010. “Potting Traditions and Cultural Continuity in Pacific Nicaragua, ad 800–­1350.” PhD diss., University of Calgary.

Torquemada, Juan de. (1615) 1975–­1983. Monarquía Indiana. De los veinte y un libros rituales y monarquía indiana, con el origen y guerras de los indios occidentales, de sus poblazones, descubrimiento, conquista, conversión y otras cosas maravillosas de la mesma tierra. 3rd ed. 7 vols. Historiadores y Cronistas de Indias 5. Mexico City: Instituto de Investigaciones Históricas, Universidad Nacional Autónoma de México.

Twiss, Katheryn C. 2012. “The Archaeology of Food and Social Diversity.” Journal of Archaeological Research 20(4):357–­395.

Waters-­Rist, Andrea, Gina Carroll, and Geoffrey G. McCafferty. Forthcoming. “Stable Isotope Assessment of the Dietary Contribution of Maize in Pre-­Columbian Pacific Nicaraguans.” Journal of Anthropological Archaeology.

Wilke, Sacha. 2011. “Cambiando la tipología de las pesas de red de El Rayo, Nicaragua.” Mi Museo y Vos 17:6–­9. Wilke, Sacha. 2012. “From Remains to Rituals: Exploring the Changing Mortuary Program at El Rayo, Nicaragua.” MA thesis, University of British Columbia, Vancouver.

160

M CCA F F ERT Y

Wyss, Sue Bursey. 1983. “San Cristóbal Archaeological Site, Managua, Nicaragua: Site Report and Preliminary Ceramic Analysis.” MA thesis, Texas A&M University, College Station. Zambrana Fernández, Jorge E. 2008. “Excavaciones arqueológicas en Tepetate, Granada.” Mi Museo y Vos 6:4–­7. Zambrana Fernández, Jorge E. 2012. Estudios arqueológicos en el sitio “Los Martínez,” Sector Pantanal, Managua. Managua: Alcaldía de Managua.

T wenty Y ears of N icaraguan A rchaeology

161

7 The Development of Social Complexity in the Rivas Region, Pacific Nicaragua Karen Niemel Garrard

DOI: 10.5876/9781646421510.c007

162

From 1999 to 2000, the author conducted a program of settlement survey (figure 7.1) within the department of Rivas, Nicaragua, as part of her doctoral requirements (Niemel 2003). To date, this remains the only systematic large-­ scale archaeological survey undertaken in the southwestern-­most portion of Nicaragua, an area that figures quite prominently in early ethnohistorical descriptions of the region. The survey area encompassed 270 square kilometers east of the Pan-­American Highway, extending from the Río Ochomogo in the north to the department’s capital, the city of Rivas, in the south. Lake Nicaragua, an important natural resource for the region’s inhabitants from prehistoric times through the present day, defined the eastern boundary of the survey limits. Early Spanish chroniclers during the sixteenth century report that the Nicarao, a Nahuat-­speaking Mexican group, inhabited Rivas at the period of Spanish entrada and that Quauhcapolca—­the capital of the modern country’s namesake, Cacique, or Chief, Nicaragua—­was probably located somewhere in the vicinity of the modern city of Rivas (Fowler 1989:68). The remains of Quauhcapolca have yet to be identified, but one possibility is that it lay under San Jorge, a modern town of about 8,000 residents along the shore of Lake Nicaragua just east of the city of Rivas. Nicarao migratory accounts also state that when the Nicarao arrived in the Rivas area, another Mesoamerican group, the Otomanguean-­speaking Chorotega, already

Figure 7.1. Rivas survey area. Map by author.

inhabited it (see chapter 2, this volume). After the Nicarao settled in Rivas, the Río Ochomogo seems to have served as the border between their territory and remaining Chorotega lands to the north (Lothrop 1926). One long-­standing hypothesis for the chronology for these events places the Chorotega arrival at approximately ad 700/800 and the Nicarao arrival around ad 1250/1350, with each group at least partially displacing existing populations (Day 1984; Healy 1980; Lothrop 1926; Stone 1966). The approximate arrival dates of the two different Mesoamerican groups have been based mostly on material T he D evelopment of S ocial Complexity in the R ivas R egion

163

culture changes, particularly within the ceramic complexes, which are notable for their Mesoamerican style and iconography. However, as the amount of research data has slowly accumulated for the Nicaraguan–­Costa Rican area, reassessment of the ceramic chronology has begun, which may in turn affect the hypothesized migration chronology (cf. McCafferty and Steinbrenner 2005; see also chapters 4, 6, 9, and 10, this volume). The Rivas settlement data were subjected to three levels of analysis and interpretation. The first level entailed the reconstruction of changes in the prehistoric settlement pattern through time. Based on this analysis, the second level formulated inferences concerning sociopolitical organization. Diachronic reconstructions of settlement patterns, such as this one, reveal how economic, political, and ideological structures and processes are expressed through the spatial activities of a society. Each level of analysis—­the region, community, and household—­is shaped by factors that are unique to it. For example, site distribution across the landscape reflects the impact of trade, defense, and administration. Studies of sites and communities mirror the larger social organization, while households reflect family organization. Other factors responsible for settlement size, location, and pattern can include changes in the natural environment; the role of demographics, economy, technology; or the influence of social and political organization on spatial distribution. Implementing a program of settlement archaeology within Rivas therefore allowed for social organization to be explored from both a macrotheoretical and microperspective. The third level of analysis and interpretation placed Rivas within a larger sociopolitical and economic context to explore the extent to which its relationships with regions to the north and south were a factor in local sociopolitical development. PHYSICAL SETTING OF THE RESEARCH AREA

Rivas is located within the geological and physiographic province known as the Depression of Nicaragua, a large region of flatlands and rolling hills formed by a fork of the Central America volcanic axis (Healy 1980:9). The depression extends from the Cosigüina Peninsula in the north to the Río San Juan, which also marks Nicaragua’s border with Costa Rica (Incer Barquero 2000:55). The eastern boundary of the depression runs parallel to the eastern shores of lakes Nicaragua and Managua. The Pacific cordillera defines the western edge of the depression approximately 16 to 32 km inland from the Pacific Coast. The principal natural feature is Lake Nicaragua (or Cocibolca, as it was originally known), which is about 160 km long and 72 km wide, with a surface 164

GARRARD

area of 8,264 km2 (Incer Barquero 2000:59). The lake is the largest freshwater body between the Great Lakes of North America and Lake Titicaca in Southern Peru and Bolivia (Incer Barquero 2000). The presence of several adapted marine species, including tarpon and (possibly) a freshwater species of shark, suggests that the lake was at one time connected to the Pacific Ocean (Healy 1980:11). Hundreds of small islets are found in Lake Nicaragua, but there are two major islands, Ometepe and Zapatera, both of which feature important petroglyphs and other archaeological remains (Healy 1980; Lange et al. 1992). Ometepe Island is approximately 13 km east-­northeast of the city of Rivas and formed by the twin volcanoes Concepción (active) and Maderas (extinct). A narrow isthmus that is often partially submerged during the rainy season connects the two volcanoes. Ometepe Island figures prominently in the migratory myth of the Nicarao—­their “promised land” lay within site of the twin peaks (Lothrop 1926:6; see also chapter 2, this volume). Zapatera Island is also an extinct volcano and located farther north along the shore of the lake near the modern city of Granada. The narrowest strip of land between Lake Nicaragua and the Pacific Ocean is 15 to 20 km wide and located just south of the city of Rivas. During the nineteenth century, Lake Nicaragua was once championed as a component of the transoceanic canal instead of the canal eventually built through Panama (Squier 1852). The proposed canal through the Nicaraguan region would have stretched from the Río San Juan to Lake Nicaragua, and then would have continued via a short artificial canal built across the isthmus to the ocean. This would have capitalized on the Río San Juan, which served as a natural travel route between the Caribbean and the Pacific and northwestern Costa Rica from prehistoric to colonial times (Lange et al. 1992:5). More recently, this idea has been again revived, and a 172-­mile transoceanic canal is currently “under construction” across the isthmus of Rivas, though as of this writing little progress has been made and the projected completion date remains in flux. Soils within the Rivas research area consist of Cretaceous sediments, which lie adjacent to Eocene sediments to the west (Lange et al. 1992:136). Laminated clays having travertine sand, marl, and gravel insertions dominate the lake’s fertile watershed (Terán and Incer Barquero 1964:91). The high clay content facilitates moisture retention and improves agricultural potential. In spite of this common feature, noticeable variation exists between the northern and southern portions of the research area. Soils in the northern half, for example, show a fair degree of erosion, something that may be more a result of colonial-­modern agricultural practices than anything else. The northern half T he D evelopment of S ocial Complexity in the R ivas R egion

165

is characterized by a lack of visible stratigraphy due to heavy seasonal rainfalls and periods of intense heat, which leach color from the soils. This leaching makes excavation by natural soil layers difficult, if not impossible, a fact also originally noted in Rivas by archaeologist Albert Norweb (Healy 1980:38–­39). Two minor geologic features are found within the Rivas research area. A narrow wedge of andesite, known as Las Mesas, is embedded in the Cretaceous sediments approximately 325 m west of the city of Rivas (Terán and Incer Barquero 1964:91). The hills of Abejonal are also located at the mouth of the Río Ochomogo, where it drains into Lake Nicaragua (Healy 1980:11). The hills range in elevation from 61 to 147 m above sea level and consist of diorite masses interspersed with Eocene sediments. The remainder of the research area is flat to slightly rolling. Elevation ranges from 36 to 61 m above sea level. Although somewhat diminished in the present day, a rich variety of flora and fauna calls Rivas home. Important mammal species of the past included brocket and white-tailed deer, jaguar, ocelot, wolf, collared peccary, coyote, gray fox, spider and howler monkeys, rabbit, opossum, porcupine, raccoon, armadillo, tapir, and coati (Healy 1980:15–­16). Reptiles include iguana, a variety of small lizards, and serpents such as the rattlesnake and coral snake. Lacustrine wildlife includes fish, shellfish, turtles, and crocodiles. Sharks, sea bass, and sawfish reach the lake through the Río San Juan (Healy 1980:16). Common trees include ceiba, laurel, cedar, jocote, tobacco, papaya, mahogany, sapote, mango, coconut palm, and rosewood. Two prehistoric tree crops, cacao and sapodilla or nispero, are also found in the area. Rivas was known as an important pre-­Columbian cacao-­producing zone in the Americas (Steinbrenner 2006). After the conquest, cacao production initially centered in Granada until it was again pushed south in the early seventeenth century (Newson 1987:139). Cacao never became an important export crop during colonial times, and eventually the demand for it within Nicaragua was met by imports from Costa Rica. In spite of this fact, Rivas is locally known as the “Tierra de los mangos” in reference to the tall mango trees lining nearly every road around the modern town. These mango trees, and others no longer extant, once provided shade for the cacao plants. The modern economy of the department of Rivas is predominantly based on cattle ranching. However, within the research area, the presence of fertile soils has led to the increased importance of crop cultivation. Maize, beans, squash, and various vegetables, such as green pepper and tomato, are the primary subsistence crops on small landholdings. Larger farms cultivate sugarcane, papaya, banana, and plantain. The largest land manager during the 1999–­2000 fieldwork was the Ingénio Benjamin Zeladón, which processed 166

GARRARD

sugarcane into raw sugar for export. The Zeladón plant is located 5 km north of the city of Rivas. During the 1970s, a canal was dug stretching from the plant to Lake Nicaragua. The quantity of archaeological remains disturbed and destroyed during its construction is notorious among local inhabitants. FIELD SURVEY METHODOLOGY

The regional survey identified a total of forty-­eight archaeological sites, representing occupations dating from 500 bc to ad 1522 (figure 7.2). Briefly, the field methodology consisted of the systematic, 100 percent survey of a series of 150 m wide, east-­west transects. Opportunistic survey was also conducted outside transects based on information provided by local informants. Limited test excavations were undertaken at four sites to obtain additional ceramic materials in the hopes of refining the regional sequence of the area and to provide data that could be used to support inferences about settlement patterns, social organization, and migration. The forty-­eight archaeological sites were first ranked from a synchronic perspective in order to develop a working typology of site types. Ranking (from smallest to largest: I, II, and III) was based on three criteria: size, volume, and the presence of indicators of elevated sociopolitical importance. The latter—­ which included earthen mounds and imported obsidian and ceramics—­was particularly important for suggesting the presence of socio­ political elites. Temporal information, based on the presence or absence of diagnostic ceramic types, was then incorporated to reconstruct settlement patterns and, using the working site typology, settlement hierarchies through time. At the time of the settlement survey, the then-­current chronology for the region was used: Orosí period (1000–­500 bc); Tempisque period (500 bc–­a d 300); Bagaces period (ad 300–­800); Sapoá period (ad 800–­1350); and Ometepe period (ad 1350–­1522). As has been noted, much about this chronology has held as the amount of research has increased within Nicaragua; however, a reassessment of the temporal markers of the Sapoá and Ometepe periods is likely (cf. McCafferty and Steinbrenner 2005; see also chapter 4, this volume). SETTLEMENT PATTERN OF THE OROSÍ AND TEMPISQUE PERIODS

No evidence was recovered during the 1999–­2000 survey to indicate settlement of the Rivas region before 1000 bc, the beginning of the Orosí period. And, unfortunately, after 1000 bc the Rivas settlement data is problematical T he D evelopment of S ocial Complexity in the R ivas R egion

167

Figure 7.2. Archaeological sites identified during 1999–­2000 Rivas survey. Map by author.

due to the lack of well-­defined ceramic complexes in the Orosí regional sequence. There is only one known diagnostic ceramic for this period: Bocana Incised Bichrome, which has been dated anywhere between 800 bc and ad 1 (Haberland 1992; Healy 1980; Hoopes 1987, 1992). For the following Tempisque period, diagnostic artifacts greatly increase. Diagnostic ceramics include the types Rosales Zoned Engraved, Schettel Incised, Charco Black-­on-­Red: Puerto variety, Puchor Red-­Slipped and Punctated, 168

GARRARD

T able 7.1. Tempisque-­period sites Site

Area (ha)

RI-­7, Las Mesas

>1

RI-­14, Santa Lucía

8

RI-­15, San Martín

14

RI-­18, El Vergel

>1

RI-­16, San Félix

RI-­19, Paco Rojas

3

RI-­20, Sabana Grande RI-­23, Santa Elena RI-­24, San Jorge

RI-­25, Finca de Cana RI-­26, José Mercedes RI-­27, La Conchita

RI-­28, Miguel Mora

9

1

4

35

2

4

8

5

RI-­32, Sucuya

15

RI-­34, San Fernando

12

RI-­33, El Pital

RI-­36, La Esperanza RI-­37, Yamil Ríos RI-­38, El Paraíso

RI-­41, Alfredo Siazer

18

15

1

9

2

RI-­42, Chata

14

RI-­44, Santa Isabel

90

RI-­43, Chatilla

Imported Goods

4

ceramics –­

ceramics –­

–­

ceramics –­

ceramics

ceramics –­

–­

ceramics –­

ceramics –­

ceramics

ceramics

ceramics –­

–­

–­

–­

ceramics

Features

Site Type

–­

I

–­

–­

–­

–­

–­

–­

–­

–­

–­

–­

–­

–­

–­

–­

–­

–­

–­

–­

–­

–­

–­

–­

I

I

I

I

I

I

I

I

I

I

I

I

I

I

I

I

I

I

I

I

I

I

Source: 1999–­2000 Rivas settlement survey by author.

Huila Zoned Punctate, and Obando Black-­on-­Red. Twenty-­three of the sites identified within Rivas had components that were assigned to the Tempisque period (table 7.1). The sites were characterized by extremely low densities of surface remains and ranged in size from 25 m2 to 90 ha. The larger sites lacked continuous surface remains and are more aptly described as site clusters, each seven to thirty hectares in size. Three sites were single component. In spite of differences in site area found during the Tempisque period, there is no clear indication of a settlement hierarchy (figure 7.3). Imported artifacts were limited to the ceramic type Usulután, originating from El Salvador T he D evelopment of S ocial Complexity in the R ivas R egion

169

Figure 7.3. Tempisque-­ period settlements. Map by author.

and west-­central Honduras and ceramics best described as “Usulután-­related,” characterized by a distinctive negative resist painting and paste texture. The source for the latter may be somewhere in north-­central Nicaragua (Salgado González et al. 1998:6; cf. Dennett 2016 and chapter 10, this volume). The presence of imported ceramics at half of the Tempisque-­period sites (n = 11/23) does not point to centralized control over their distribution. Overall, the settlement pattern of the Tempisque period was dispersed, suggesting a lower 170

GARRARD

population than in subsequent periods. However, incipient aggregation can be hypothesized for several of the larger sites, where later Bagaces-­period surface remains covered larger areas. SETTLEMENT OF THE BAGACES PERIOD

Twenty-­nine of the sites located within the survey area have Bagaces-­ period components—­representing a striking 26 percent increase in the total number of settlements (table 7.2). Sites that had been occupied during the preceding Tempisque period were also two to five times larger in the Bagaces period, while ten sites were occupied for the first time. Collectively, this evidence is taken to indicate an increase in population levels. Ceramics diagnostic of the period include Rosalita Polychrome, Chávez White-­on-­Red, Galo Polychrome, Rivas Red: León variety, Tola Trichrome, Momta Polychrome, and the first of the Papagayo varieties—­Papagayo: Culebra variety. Three Bagaces sites were characterized by at least one earthen mound approximately one-­and-­one-­half to 3 m in height and ranging from 10 to 20 m in diameter. These sites are Paco Rojas (RI-­19), San Jorge (RI-­24), and Santa Isabel (RI-­44). Excavations conducted on the mound located in the central area of the Paco Rojas site revealed cultural deposits over 2 m in depth. The recovery of burnt daub from the mound also suggests that a structure once stood there. Unfortunately, repeated agricultural plowing has destroyed any evidence such as packed-­earth walking surfaces or post molds that may have been existed. A similar, low mound was present at the Santa Isabel site. The mound was located in the part of the site corresponding to Santa Isabel “B,” as originally reported by Paul Healy (1980:58). The mound was approximately 20 m by 2 m in size. Prior to Healy’s reporting of the site, the mounds of San Jorge were visited in 1959 and 1961 by Gordon Willey and Albert Norweb (Healy 1980); they have since been destroyed. Healy (1980:41) described these mounds, which differed from those of other sites in the areas, as being “located in rolling grazing lands which slope markedly toward the lakeshore.” He considers them to have been platforms for house residences, probably serving to protect the structures against flooding. In addition to the mounds found at these sites, they were also the only three sites that yielded imported ceramics and obsidian (for a discussion of the likely foreign origins of most of the obsidian tools found in Nicaraguan archaeological contexts, see chapter 12, this volume). Foreign ceramics during this period were limited to the Honduran polychromes Delirío Red-­on-­W hite and Tenampua. A total of sixty-­nine obsidian artifacts were also recovered, thirty-­nine of which were from excavated T he D evelopment of S ocial Complexity in the R ivas R egion

171

T able 7.2. Bagaces-­period sites Site

RI-­1, Las Piedras

RI-­6, Ingenio Xavier Guerra RI-­8, San Joaquín

RI-­10, Salvador García RI-­11, San Jerónimo RI-­14, Santa Lucía

Area (ha)

Imported Goods

Features

Site Type

29

–­

–­

I

22

–­

>1

–­

>1

–­

>1

–­

2

–­

–­ –­ –­ –­ –­

I I I I I

RI-­15, San Martín

30

RI-­19, Paco Rojas

45

ceramics, obsidian

1 mound

II

RI-­21, Sergio Martínez

>1

–­

–­

I

RI-­25, Finca de Cana

14

RI-­16, San Félix

RI-­20, Sabana Grande

RI-­24, San Jorge

RI-­27, La Conchita

RI-­28, Miguel Mora

7

40

RI-­37, Yamil Ríos RI-­38, El Paraíso RI-­40, San José

RI-­41, Alfredo Siazer RI-­42, Chata

RI-­43, Chatilla

RI-­44, Santa Isabel

RI-­45, San Francisco

–­

I I

I

II

24

–­

–­

I

–­

5

36

RI-­36, La Esperanza

–­

2+ mounds

RI-­33, El Pital

RI-­35, Santo Domingo

–­

–­

ceramics, obsidian

>1

RI-­34, San Fernando

–­

70

RI-­31, La Ceiba RI-­32, Sucuya

–­

–­ –­

31

–­ –­

35

ceramics

30

–­

2

–­

5

–­

9

–­

3

–­

2

–­

14 10

204

15

–­ –­

ceramics, obsidian

Source: 1999–­2000 Rivas settlement survey by author.

–­

–­ –­ –­ –­ –­ –­ –­ –­ –­ –­ –­ –­ –­ –­

2+ mounds –­

I I I I I I I I I I I I I I

III I

Figure 7.4. Bagaces-­period settlements. Map by author.

contexts at the Paco Rojas site. Of the excavated obsidian artifacts, thirty-­ seven were visually sourced to Güinope, Honduras, and two to Ixtepeque, Guatemala (Niemel 2003). Based on the differences between sites, a three-­tiered hierarchy settlement was defined for the Bagaces period. Santa Isabel was at the top of the hierarchy of three levels (figure 7.4). The site was clearly larger (approximately 204 hectares in size) and more densely occupied than any other site of the period. T he D evelopment of S ocial Complexity in the R ivas R egion

173

It covered an area approximately double that of both the Paco Rojas and San Jorge sites. In spite of this, the existence of artificial mounds and imported materials at these two sites differentiates them from the remaining sites of the period. They were also located approximately 5 km from Santa Isabel and each other, forming a triangle. It appears that Paco Rojas and San Jorge represent secondary sites in a three-­tiered settlement hierarchy. The settlement system during the Bagaces period indicates the emergence of a society with at least an incipient process of social and political differentiation. The Santa Isabel site emerged as a regional center, and most of the population was probably aggregated there. There was an increase in macro­ regional interaction from the preceding period with areas located to the north of Rivas, as indicated by the imported obsidian and Honduran ceramics. The distribution of foreign artifacts suggests that networks of long-­distance trade were controlled by inhabitants of the largest sites, if not solely by those of Santa Isabel. The foreign artifacts represent not only a mechanism of wealth accumulation but serve as status builders, associating emergent elite with distant places. SETTLEMENT OF THE SAPOÁ PERIOD

The Sapoá period is notable for a marked change in the ceramic complexes. Although some continuity remains from the preceding period, the ceramic complexes are overwhelmed by the introduction of Mesoamerican motifs and a white or cream slip on polychromes in particular. Diagnostic ceramics include Papagayo Polychrome (all varieties excepting Culebra, as already noted) and Granada, Mombacho, and Pataky polychromes, as well as Sacasa Striated, a very widespread monochrome ceramic with a distinct shoe-­shaped form, red-­slipped rim, and brushed exterior (Abel-­Vidor et al. 1990:229). A total of thirty-­four sites were occupied during the Sapoá period (table 7.3). Eight of the sites, in particular, revealed evidence of initial occupation. Five sites occupied during the Bagaces period (i.e., Salvador García [RI-­10], San Jerónimo [RI-­11], Sabana Grande [RI-­20], Sergio Martínez [RI-­21], and El Paraíso [RI-­38]) were apparently abandoned. Twelve additional Bagaces sites revealed a 60–­90  percent reduction in total settled area. All sites that were either abandoned or that decreased in size were located inland. At the same time, there was an increase in extension and artifact density of three lakeshore sites: Las Piedras (RI-­1), San Jorge (RI-­24), and Santa Isabel (RI-­ 44). This pattern suggests a change in use of environmental resources, specifically those of the lake. 174

GARRARD

T able 7.3. Sapoá-­period sites Site

RI-­1, Las Piedras

RI-­2, Santa Roasa

RI-­5, Humberto Bazarano

RI-­6, Ingenio Xavier Guerra

RI-­8, San Joaquín

RI-­9, San Román

RI-­12, El Corral

Area (ha) 27

Imported Goods