Handbook of Middle American Indians, Volume 4: Archaeological Frontiers and External Connections 9781477306598

Archaeological Frontiers and External Connections is the fourth volume in the Handbook of Middle American Indians, publi

171 11 234MB

English Pages 376 [377] Year 2015

Report DMCA / Copyright

DOWNLOAD FILE

Polecaj historie

Handbook of Middle American Indians, Volume 4: Archaeological Frontiers and External Connections
 9781477306598

Citation preview

HANDBOOK OF MIDDLE AMERICAN INDIANS, VOLUME 4 Archaeological Frontiers and External Connections

THIS PAGE INTENTIONALLY LEFT BLANK

HANDBOOK OIF MIDDLE AMERICAN INDIANS EDITED

AT MIDDLE AMERICAN RESEARCH INSTITUTE, TULANE

ROBERT W A U C H O P E , General

Editor

MARGARET A. L. HARRISON, Associate INIS PICKETT, Administrative

Editor

Assistant

T H O M A S S. SCHORR a n d DAVID S. P H E L P S , Art K E N N E T H E . O W E N , Art

WITH

Editors

Assistant

L O R E N E GREGG C A M P B E L L , ASSEMBLED

Indexer

T H E AID O F A GRANT

FROM

T H E NATIONAL

FOUNDATION, AND UNDER T H E SPONSORSHIP O F T H E NATIONAL COUNCIL C O M M I T T E E O N LATIN AMERICAN ANTHROPOLOGY

Editorial

Advisory

UNIVERSITY

Board

IGNACIO BERN AL, HOWARD F . CLINE, GORDON F . E K H O L M , N O R M A N A. M C Q U O W N , M A N N I N G NASH, T. DALE STEWART, EVON Z. VOGT, ROBERT C. W E S T , GORDON R. W I L L E Y

SCIENCE RESEARCH

THIS PAGE INTENTIONALLY LEFT BLANK

HANDBOOK OF MIDDLE AMERICAN R O B E R T W A U C H O P E , General Editor INDIANS VOLUME

FOUR

Archaeological Frontiers and External Connections GORDON F. EKHOLM GORDON R. WILLEY Volume Editors

U N I V E R S I T Y

OF

T E X A S

PRESS

A U S T I N

Copyright © 1966 by the University of Texas Press First paperback printing 2015 All rights reserved Printed in the United States of America The preparation and publication of The Handbook of Middle American Indians has been assisted by grants from the National Science Foundation. Requests for permission to reproduce material from this work should be sent to:  Permissions   University of Texas Press   P.O. Box 7819   Austin, TX 78713-7819  http://utpress.utexas.edu/index.php/rp-form

Library of Congress Catalog Number 64-10316 isbn 978-1-4773-0658-1, paperback isbn 978-1-4773-0659-8, library e-book isbn 978-1-4773-0660-4, individual e-book

C O N T E N T S

1. Archaeology and Ethnohistory of the Northern Sierra Charles C. DiPeso

3

2. Archaeology of Sonora, Mexico Alfred E. Johnson

26

3. Archaeology and Ethnohistory of Lower California William C. Massey

38

4. Archaic Cultures Adjacent to the Northeastern Frontiers of Mesoamerica Walter W. Taylor

59

5. Mesoamerica and the Southwestern United States J. Charles Kelley

95

6. Mesoamerica and the Eastern United States in Prehistoric Times James B. Griffin

.

.

111

7. Archaeological Survey of El Salvador John M. Longyear, II1

132

8. Archaeological Survey of Western Honduras John B. Glass

157

9. Archaeology of Lower Central America S. K. Lothrop

180

10. Synthesis of Lower Central American Ethnohistory Doris Stone

209

11. Mesoamerica and the Eastern Caribbean Area Irving Rouse

234

12. Mesoamerica and Ecuador Clifford Evans and Betty J. Meggers

243

13. Relationships between Mesoamerica and the Andean Areas Donald W. Lathrap

. . . .

14. The Problem of Transpacific Influences in Mesoamerica Robert Heine-Geldern

265 277

15. The Role of Transpacific Contacts in the Development of New World Pre-Columbian Civilizations Philip Phillips

296

References

319

Index

349 vii

THIS PAGE INTENTIONALLY LEFT BLANK

HANDBOOK OF MIDDLE AMERICAN INDIANS, VOLUME 4 Archaeological Frontiers and External Connections

GENERAL

EDITOR'S

NOTE

The manuscripts for the following articles were submitted at various dates over a period of two and one-half years. Because of revisions and minor updatings made from time to time, it is difficult to assign a date to each article. In some cases, an indication of when an article was completed can be had by noting the latest dates in the list of references at the end of each contribution.

1. Archaeology and Ethnohistory of the Northern Sierra1

CHARLES

THAT portion of Mexico here defined as the northern Sierra2 consists of a littleknown rectangle (fig. 1) of mountainous terrain, approximately 87,750 sq. km. It lies mostly in northwestern Chihuahua but 1

Because this article was written in the field during the Joint Casas Grandes Expedition pertinent references may have been omitted. Assistance is acknowledged for field photography (T. Carroll), maps ( E . Contreras), typing and proofing (A. Withers and B. Colvin). 2 Until the Geophysical Year of 1958, the northern Sierra Madre Occidental remained virtually unmapped. During this year the Mexican Army joined with the U.S. Air Force in an aerial mapping project which covered the Republic of Mexico and was reported in 1958 under the title of Estados Unidos Mexicanos. It is a mercator projection scaled 1:500,000. Sections 12R-II, 12R-IV, 13R-I, and 13R-III cover the area under discussion. The Millionth Map Series for Sonora and Chihuahua, as well as the World Aeronautical Chart, sheets 405, 406, 470, and 471 and the USAF Preliminary Base Section 471 B, contain innumerable inaccuracies which render them useless in the Sierra country. The lumber company known as the Bosques de Chihuahua holds large tracts of land in the northern Sierra district and has, in its private files, aerial photographs which it uses in surveying its lumber tracts. These are not generally available but are of great help when working in the area.

C. DlPESO

includes part of northeastern Sonora and the southern margins of Arizona and New Mexico just north of the international border in the United States.3 GEOGRAPHY

The physical characteristics of the northern Sierra are described by Brand (1936, 1937), Goldman (1951, pp. 119-24, 421-25), and Lister (1958, pp. 3-7). It is wild country, full of topographic contrasts and rich in natural resources. The area was considerably utilized by prehistoric peoples, some of whom are associated with 3 This area measures some 355 km. north-south and 270 km. east-west. The distribution of Casas Grandes pottery types as trade pieces overflows the northern Sierras. Sherds of these pottery types are found as far west of the northern Sierra as the Trincheras site in Sonora, on the Rio Magdalena (Sauer and Brand, 1931, figs. 1, 2, pp. 91-94; Brand, 1935, p. 294); as far north as the Gobernador Canyon, in the Rio Arriba drainage of New Mexico (Brand, 1935, p. 294); and as far south as the Valley of Mexico (Kidder, Jennings, and Shook, 1946, p. 251; also see Linné, 1942, pp. 175, 178, fig. 318).

3

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

FIG. 1—THE NORTHERN SIERRA, MEXICO

the manufacture of Casas Grandes polychromes.4 The high, flat mountains, the narrow, fertile mountain valleys, as well as the broader valleys which lie along the eastern flanks of the Sierras (in the basin and range province) contain many remains of early populations (fig. 2). The northern spine of the Sierra Madre Occidental covers some 80 per cent of the northern Sierra district (see fig. 3). These highlands form a great curved rim which separates the three major basin-river valleys of the area from Sonora on the west and southern Chihuahua to the south. The northeastern flanks of the Sierra Madre have been formed into basin and range topography. This section appears as a great cuplike depression which receives the bulk of the Sierra east slope waters. This basin is dotted with numerous dry lake beds which are very likely to contain water during the rainy season. Volcanic mountains separate this con-

4 Descriptions of Casas Grandes Polychromes can be found in: Amsden, 1928; Brand, 1935; 1943, pp. 156-57; Carey, 1931, pp. 338-59; Chapman, 1923; Harcum, 1923; Hawley, 1936 (1950 e d . ) , pp. 60-62, 94-98; Hewitt, 1908; Hough, 1923, p. 34; Kidder, 1916; 1924; Lister, 1958, pp. 69-77; Lumholtz, 1902; Sayles, 1936a. The terminology commonly used today is that established by Sayles and agreed upon by Brand.

4

cavity into three distinct river valleys. These streams drain from south to north and debouch into a series of lesser lake basins. The dominant valley, commonly known as the Casas Grandes valley (fig. 4), lies nearest the spine of the Sierras and parallels the main mountain range. This, the most fertile of these valleys, is in the center of the northern Sierra, a physical advantage which drew prehistoric populations as a magnet. Undoubtedly, the cultures which formed here at one time dominated the entire northern Sierra district. The Santa Maria valley lies 50 km. east of the Casas Grandes drainage. It is smaller, contains less surface water and less bottom land, but it, too, supported prehistoric pueblos particularly in the area of the Cienega of Navacoyan. Little is known of the archaeology of this drainage. A third valley, Carmen, 45 km. east of the Santa Maria valley, is the least promising. It contains fewer prehistoric pueblos and sites than do the other two. These three major valleys deposit their waters into shallow lakes. The Casas Grandes waters drain northward into Lake Guzman basin, the Santa Maria ends in Laguna de Santa Maria, and the Carmen drainage supplies Laguna de Patos. These lake beds are associated with nearby salinas, a con-

FIG. 2—ARCHAEOLOGICAL SITES IN THE NORTHERN SIERRA

FIG. 3—CERRO DE MOCTEZUMA. View to south from basin and range area to the high northern Sierra. On the peak in the foreground lies one of the great circular stone towers associated with the Casas Grandes culture. (Courtesy, Amerind Foundation. Photo, Tommy Carroll.)

6

ARCHAEOLOGY & ETHNOHISTORY: NORTHERN SIERRA

stant source of salt, as well as with barren sand dunes. The western margin of the area can be correlated with the Bavispe-Yaqui drainage in the state of Sonora. The Bavispe drainage begins near Tres Rios in the high Top-ofthe-Mountain country. Innumerable small tributaries gather at this point and drain northward through a narrow mountain valley past the town of Bavispe to Colonia Oaxaca,5 where the river makes a great bend to the west. At the town of Batepito the water volume is increased because of the San Bernardino junction which comes in from the north. The stream then turns south and flows along the craggy foothills of the Sierra Madre and receives waters from the west slope of the northern Sierra. Near Monte Cristo the Bavispe joins the Aros (Papagochic) River and the two form the bulk of the waters which flow into the major Yaqui drainage of Sonora. On the west this Bavispe-Yaqui drainage marks a natural border of the northern Sierra territory. The Sierra Madre, as seen from the Sonora side, has a threatening appearance. The rough jagged face consists of masses of red-brown crags which have been dissected by innumerable steep-sloped arroyos. The western slope of the Sierra Madre rises to an average elevation of 1900 m.; some of the peaks rise to over 2700 m. (Brand, 1937, p. 29). Despite their wild appearance, the western slopes can be penetrated in several places. One of the oldest trails, still used, passes through the Mala Noche country in the Nacori basin.6 The Spanish explorers coming up the Yaqui drainage from the south 5 It was in this area that Jesuit missionaries found the Opata, Jova, and Eudeve tribes living in the 1640's. These sedentary peoples reportedly were enemies of the Janos, Sumas, and Jacomes, who occupied the basin and range area in the northeastern part of the northern Sierra country. 6 The hot, miserable climate and insect-infested nights give this area its name. In this ecological niche the Yerba Flecha bush grows, a "poison tree" feared by the natives for its purgative powers which reportedly can kill man and beast alike.

may have followed this trail in traversing the Sierra Madre Occidental. The uplands of this area presented no barrier to the aboriginal peoples. Throughout the northeast section of the north Sierra are open sites, cave sites, rock-wall terraces, stone towers, and a network of ancient trails. From the findings of the Joint Casas Grandes Expedition and others, the inhabitants of the valleys, as well as those of the mountains, made abundant use of the many natural resources of the mountain terrain. Most arroyos were terraced to conserve topsoil and add tillable areas. Pine was plentiful and used extensively for house construction and fuel. The Indians quarried a red sandstone from the Arroyo de Tapiecitas and the Pajaritos Mountains; they obtained quartz crystals from the west slopes of the Asita Hueca Mountains, east of Basarac. White kaolin, for the making of pottery, was mined from beds in the Dos Cabezas range, southwest of Valle Seco, and in the headwaters of San Pedro de Arriba, known today as Alta Mirano. Surface metal deposits in this area furnished the makers of Huerigos and Carretas Polychromes with glaze paint. Raw copper, found in the ruins of Casas Grandes, outcrops 10 km. northwest of Huachinera; sodasalt is found in the Bacadehuachic area. The presence of turkey bones in the basin-valley sites suggests that this bird was taken from its mountain habitat and penned in the lower valley villages. Deer and other mammals were hunted. Seed crops gleaned from the highlands added to the subsistence of the old inhabitants of the northern Sierra. Large stone pueblos, as well as those of puddled adobe, are scattered throughout the eastern highlands in the Cuesta Blanca region, Pacheco, Cave Valley, Valle Seco, and the Tres Rios area. Often these pueblos contain over 100 rooms. The network of native trails which wind through the high Sierra, and the strategic positions of circular and rectangular lookout towers at valley intersections, show that the Sierras formed no

7

FIG. 4—CASAS GRANDES RUIN, CHIHUAHUA. View to south, down the valley. Note the relationship of the bottom land to the desert-like terraces. (Courtesy, Amerind Foundation. Photo, Tommy Carroll.)

8

ARCHAEOLOGY & ETHNOHISTORYI NORTHERN SIERRA

barrier between east-west communication. Millions of sea shells, recently uncovered by the Joint Casas Grandes Expedition, were traded from the Gulf of California into the inland basin area of the north Sierra (Brand, 1938; Tower, 1945). The highlands of the northern Sierra have attracted many cultural groups in historic times. First, the Spaniards penetrated the area in search of metals, then settled down to ranch. The Apache then used it as a refuge, where the last band reported was that of Indian Juan and his followers, who met their death in the late 1920's (Hatch, 1954, pp. 100-04). A number of colonies of Latter-day Saints (Mormons) escaped the effects of the Edmunds Act of 1882 by settling in the Sierras and brought their culture to the area (Hatch, 1954; Schwartzlose, 1952). In the 1920's a group of Kickapoo Indians bought a ranch and established themselves in the middle Bavispe area two days north of Tres Rios (Amsden, 1928, p. 12). Tarahumar Indians are frequent travelers in these mountains as well as in the basin and range area. Each spring and fall, groups of them move through the country begging for a living. The Sierra Madre apparently has always been a highway for the transmission of culture during both prehistoric and historic times (Lister, 1958, pp. 119). The northern limits of the northern Sierra must be defined arbitrarily because no mountain ranges or river valleys mark a natural border, the basin and range area continuing unbroken into southern Arizona and New Mexico. This point is essential to an understanding of relationships between the prehistoric populations in this general area. There was no barrier to north-south movements along this border. Recent findings of Lister (1958) in the mountain area and of the Joint Casas Grandes Expedition in the Casas Grandes valley indicate a close connection between the inhabitants of the northern Sierra and those of the Mogollon area in the north during the early part of

the historical continuum of the former district. The eastern border of the northern Sierra is marked by a drastic change in biotic zones. The area east of this margin is desolate dry alkali country, not conducive to sedentary productivity and, at best, probably utilized only by small nomadic groups. Little is known of its archaeology except from cursory surface surveys. The terrain changes gradually into the environment described above. The Carmen River valley which ends in Laguna de Patos marks the eastern border of the northern Sierra both culturally and topographically. Along the southern border lies high mountain country, not unlike that on the western border. For convenience the Aros (Papagochic) River drainage can be designated a southern boundary. This drainage cuts haphazardly from east to west and joins the Yaqui drainage in Sonora. Surface surveys indicate a break in prehistoric occupation south of the Aros drainage (Brand, 1939, pp. 75-105). Little material culture has been reported from the Aros River to the site of Zape in the state of Durango (ibid.). The Sierras north of the Aros drainage supply water for most, if not all, of the three main river valleys of the northern Sierra basin and range area. In this district Carey (1931) and Kidder (1939) reported on excavations in the Babicora sites, which resemble those throughout the northern Sierra area. RECORDED ANTHROPOLOGICAL STUDIES OF THE AREA

Many travelers have ventured through the northern Sierra.7 One of the first was Alvar Núñez, 8 better known as Cabeza de Vaca, 7 Brand (1935, pp. 289-90 and note 10) reviews the authorities who have penetrated the eastern margins of the Sierra country. 8 The Cabeza de Vaca Journal was translated by Fanny R. Bandelier and published in 1905 as The Journey of Alvar Núñez de Vaca. Hallenbeck produced another version in 1940, entitled The Journey and Route of Alvar Núñez Cabeza de Vaca. A

9

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

and his companions. These incredible Spaniards made their way from Florida across the continent to the junction of the Conchos and the Rio Grande rivers from 1534 to 1536. Their exact route from this point to Sonora is in question, but it is possible that these men, walking in the direction of the setting sun, crossed through the northern Sierra country. The leader of the next Spanish adventurers to penetrate the northern Sierras, as presently known, was General Francisco Ibarra, 9 who went in search of new lands and mineral wealth during 1564 and 1565. This group apparently made their way up the Yaqui drainage and struck the southwest corner of the northern Sierra country study made by Davenport and Wells appeared as The First Europeans in Texas. Sauer rerouted de Vaca (1932, pp. 14-21), but the Obregón Journal (Hammond and Rey, 1928) is full of inferences that the Ibarra party traveled along part of de Vaca's trail while in the northern Sierra area. On p. 147 Obregón states flatly that it took de Vaca 17 days to cross the northern Sierra from Chihuahua to Sonora. De Vaca is also mentioned on pp. 19495 and 290. Obregón claimed that de Vaca prayed for rain in the vicinity of Paquime (p. 202) and that he was given five bags of silver in the Sierra country (p. 207). Sauer (1932, p. 16) thought that the northern Sierra was a difficult crossing, but actually this country is crisscrossed by prehistoric trails. Bandelier (1890b, pp. 24-67 and frontispiece) routes Cabeza de Vaca through the Sierras after crossing northwestern Chihuahua near Casas Crandes. 9 The Ibarra journey described by Obregón was copied by Cuevas in 1924 and translated by Meecham in 1927 as well as by Hammond and Rey in 1928. In 1932 Sauer discussed Ibarra's route (pp. 38-50, 5 4 - 5 8 ) . All agree that the area of Paquime was the Casas Grandes valley, as did Hackett (1923-37, introduction). I have made several journeys to study the possible Ibarra routes through the Sierras and believe that the Spanish party crossed from west to east, entering the highlands east of the Mala Noche country northeast of Sahuaripa, following the old Bonita Trail to the Top-of-the-Mountain country either to Tres Rios or Chuhuichupa, and then going on to the eastern rim of the Sierras and into Casas Grandes valley. This belief runs counter to Sauer (1932, pp. 3 8 50), who routed Ibarra much farther north. Recent excavations at Casas Grandes have uncovered stuctures such as those described in Obregón's account of the abandoned city of Paquime.

10

in the vicinity of the Sahuaripa Valley. Led by native guides, they probably entered the western flanks of the Sierra Madre near Nacori Chico and crossed the Top-of-theMountain country at Chuhuichupa or Tres Rios. In this mountainous terrain their Sonoran guides fled one night, leaving them without interpreters. The battle-worn group wended its way eastward across the high plateau to the Casas Grandes valley. The actual routes of both de Vaca and Ibarra must be verified before a historical bridge between prehistory and history can be established for the northern Sierra district. Ibarra (Hammond and Rey, 1928, pp. 20708) mentions the presence of a rude tribe living near the ruins of Paquime, thought to be the present ruins of Casas Grandes (ibid, p. 206, note 278; Sauer, 1932, pp. 48-49). These people did not associate themselves with the builders of Paquime. They told General Ibarra, by sign language, that the master builders of this impressive city had been defeated in a war with their enemies who lived on the west side of the northern Sierras. As a result of this war, the ancient urbanites were forced to retreat five jornadas north. Bandelier (1892, p. 523-24) was informed by Opata at Basarac that the people of Casas Grandes were hostile toward the Opata of Batesopa and Baquigopa, but this does not necessarily mean that the people of Casas Grandes at this time were the builders of Paquime (ibid., pp. 571-72). Jesuit missionaries were the first to enter the northern Sierra country after Ibarra. In the 1620's they were laboring along the western and northern margins.10 In the 10 J. F. Bannon (1955) of the Jesuit Order compiled an excellent history of Jesuit activities in Sonora between 1620 and 1687. He includes information about the missions of Sahuaripa, San Mateo, Bacadehuachic, Huachinera, Basarac, and Bavispe, all missions held under the jurisdiction of the Martyrs of Japan and considered to be Opata settlements. Decorme (1941), Bolton (1936), and Dunne (1940, 1944, and 1948) have all produced secondary works regarding missionary activity irr this area of the western flanks of the Sierras. Earlier

ARCHAEOLOGY & ETHNOHISTORY: NORTHERN SIERRA

1660's Franciscans of the Zacatecas Order came to work along the eastern flanks. At this time Fray Andrés Perez of this Order was sent by Don Francisco Beaumont to the Casas Grandes area (Bandelier, 1892, p. 572, note 2). 1 1 The Pueblo Revolt in 1680 along the Rio Grande in New Mexico stimulated Spanish colonists on the Rio Grande to settle in the river valleys of the northern Sierra area. This in turn sparked native revolts among the Janos, Sumas, and perhaps the Conchos living in the vicinity of Casas Grandes. After the 1680's came numerous Apache bands such as the Chiricahuas, Gilenos, Mímbrenos, and Farones. 12 The presence of these wild bands practically closed the northern Sierra for more than 200 years. The mountains gained a fearful reputation, and even the natives shunned them because of Apache warfare. La Flora visited the

works produced by missionaries concerned with the area include such studies as those by Pfefferkorn, Perez de Ribas, Nentuig, Florencia, and Alegre. 11 Bandelier (1892, pp. 543-44) suggests that Franciscans from the Order of Zacatecas administered to the natives living along the eastern flanks of the Sierras around 1667. The Joint Casas Grandes Expedition is currently indexing and microfilming the Parral Archives in search of new information regarding this period. Bancroft (1886) has little to add, Forbes (1957) and A. E. Hughes (1935) are concerned primarily with the El Paso district, and Almada (1945) is too general. Sauer (1934, 1935) has made considerable contribution. J. B. Johnson (1950) reported on the Opata of the Sonora area near Tonichic. Beals (1932), Hrdlicka (1904), Lombardo (1702), Núñez (1777), Ocaranza (1933), Pinart ( n . d . ) , Cabeza de Vaca (1871), and Zúñiga (1835) deal with the Opata and other tribes living along the western flanks of the northern Sierra. In a language affiliation chart and map of Mexico, Swadesh (1959) included the northern Sierra in the Macro-Nawan division by relating such languages as Concho, Suma-Jumano, Jova, and Opata, which were spoken at one time or another in the northern Sierra area. 12 In 1796 Cordero described the Apache bands (translated by Matson and Schroeder, 1957). Details regarding the Apache infiltration of this area and Pimeria Alta can also be obtained from Orozco y Berra (1864, chap. 2 5 ) , Benavides (1954), DiPeso (1956, pp. 3 3 - 3 5 ) , and Forbes (1960).

Casas Grandes valley in 1766 and briefly described the ruins. A few intrepid travelers such as Hardy (1829), Escudero (1834), García Conde (1842, 1849), Bartlett (1854), Bourke (1886), and GuilleminTarayre (1867, 1869) described the geography. 13 The first professional anthropologist to interest himself in the area was Adolph F. Bandelier,14 who traveled there in 1885. In his studies of the Casas Grandes area he posed such questions as: "What the tribe was, what language they spoke, what were the causes that produced their downfall and what has become of them." He felt (1892, p. 571) that the Concho and Suma tribes were not part of the prehistoric culture which controlled the northern Sierras. Some five years later, Lumholtz (1902, vol. 1) left Bisbee, Arizona, and cut southeast across northeastern Sonora to the Bavispe drainage. He traveled to Huasasbas and then east across the northern Sierra 13 Tassin (1902) stated that Don Enrico Muller, director of the Chihuahua mint, excavated in the Casas Grandes ruins, where one of his men tunneled into a large room and found a huge meteorite wrapped in "coarse linen." This specimen was taken to the National Museum in 1876, when Tassin analyzed it. Bandelier (1892) gives additional details. The Joint Casas Grandes Expedition has uncovered remnants of these tunnels and has located a smaller bit of meteoric iron in a multistoried house-cluster. 14 Bandelier traveled afoot and by pack train across the Sierras from Fort Huachuca, going south down the Sonora River valley to Babicora and then east to Moctezuma and to Huasabas. Penetrating the Sierra less than a year after General Crook had attended to the Apache problems there, he went northeast to Tesorobabi and north up the eastern branch of the Bavispe River to Bavispe. After going east through the pass at Fusiles Carretas and on across the Llanos de Carretas to Janos along the San Pedro drainage, he turned south upstream along the river of Casas Grandes. His findings were reported in 1892. His portfolio (1880-85) of drawings of Indian artifacts and ruins made en route was sent to the Pope and is housed in the Vatican Library. News articles appearing in the Nation (1890) gave an account of his travels. Those data pertaining to the Spanish history and ethnohistory of the area were published in his final report (1890b).

11

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

through the Top-of-the-Mountain country to Pacheco and down the Piedras Verdes River to the Casas Grandes valley. He took copious notes on the nature of the country and excavated various ruins en route. 15 While Lumholtz was in the field, Schwatka (1893) published on the land of cave and cliff dwellers; Blackiston (1905, 1906a, 1906b, 1909) reported on certain cave finds in the mountains. Interest in the Casas Grandes culture, which dominated the northern Sierra in prehistoric times, was greatly stimulated by these men; many illicit ceramic collections were bought by museums which offered markets for good pieces.16 Hewitt (1908), working under the auspices of the Archaeological Institute of America, continued Bandelier's pioneer work and defined the geographical extent of the northern Sierra and the homeland of the Casas Grandes culture. He became interested in the Aztec migration myth which had been perpetuated by missionaries and Spanish soldiers and he suggested that Casas Grandes ruins were not the third stopping place in the peregrinations of the Aztecs, but that the inland basin of northern Chihuahua might well be the "Vale of Aztlan" (Hewitt, 1923, p. 50). In 1924 Kidder defined the northern Sierra as part of the Southwestern culture area. He felt (1916, p. 268) that in some 15 Lumholtz (1902) brought to light many artifacts which belonged to the Casas Grandes culture. His collections are in the American Museum of Natural History. He excavated in the Cave Valley area of the Piedras Verdes drainage, in the vicinity of Colonia Juarez, and also near San Diego where the Casas Grandes valley begins. 16 Lumholtz, however, obtained permission to make a collection. A number of these collections are described by Kidder (1916, pp. 253-68; 1924), Chapman ( 1 9 2 3 ) , Harcum (1923), Hough (1923), and Sayles. See Gladwin's discussion in Sayles' report (1936b, p. 9 2 ) , which was used as a base for Sayles' study of Chihuahua ceramics (Sayles, 1936a). Also see Gladwin, 1957, pp. 3 2 9 37.

12

way the Casas Grandes ceramic tradition was connected with the Mimbres culture. In 1916 he carried on field work in the Babicora area along the Arroyo de las Varas and in the Garabato Canyon area (Kidder, 1939). In 1926 Noguera published a report on the Casas Grandes area which contained a certain amount of survey data and surmise (1926). Two years later, Monroe Amsden (1928) carried on a reconnaissance which took him south from Agua Prieta in Sonora to Colonia Morelos, on to the Bavispe, along this valley to its headwaters at Nacori, and down the Yaqui drainage, where he turned northwest and traveled to Huasabas. From this point he turned west to Babicora and then up the Sonora valley to Arispe. Here he cut east across the mountains to Nacozari, north to Fronteras, and back to Agua Prieta. En route he described some 49 prehistoric settlements along the eastern fringes of the area. Twenty-four of these were thought to be "Peripheral Casas Grandes" located along the western Bavispe drainage to the international border; 25 sites in the Sonora Valley were defined as "Rio de Sonora" culture (1928, pp. 44-49). The same year Marquina (1928) included the Casas Grandes area in his architectural comparisons of archaeological monuments in Mexico and listed the Casas Grandes area in his location of principal architectural structures. This work was followed by Robles' study (1929) of the archaeology of the Casas Grandes region. In 1930 Hewitt summarized his beliefs regarding the Casas Grandes culture. Brand, after having worked a year with Sauer along the western margins of the northern Sierra (Sauer and Brand, 1930, 1931), spent 1930 in the area (1932). His sustained interest led to more thorough investigation and also inspired students such as Lister toward this field.

ARCHAEOLOGY & ETHNOHISTORY: NORTHERN SIERRA

DEVELOPMENT OF HISTORICAL OPINION

In 1928 Carey began his study of museum collections, such as the Lumholtz collection in the American Museum of Natural History, preparatory to his excavations in the Casas Grandes drainage in the vicinity of Corralitos and in the Babicora district (1931, 1954, 1955). He approached the problem as a time-depth study. In the Babicora area he uncovered structures which had two floor levels (Carey, 1931, pp. 33536); the upper contained Chihuahua polychromes, the lower contained only a blackon-red decorated ware. He felt (ibid., p. 371) that the architecture of the Casas Grandes mounds showed "contrast" rather than similarity to both the Southwest and Mexico and believed (p. 372) that the Casas Grandes culture of the northern Sierra showed many local developments. It held a closer affiliation with the Southwest than with Mexico, but a minimal contact with the southern Mexican area was reflected in the cultural evidence. Zingg (1940) excavated several prehistoric cave sites in southern Chihuahua. Brand's further studies (1936, 1937) contain the best available description of the geography of the northern Sierra. He emphasized the southwestern affiliations of this culture and denied any Mexican influence (1935, p. 287). He attacked the time-depth problem but theorized (1) that there were no Basketmaker or early Pueblo 1 or perhaps Pueblo 2 remains to be found in the northern Sierra, (2) that the absence of central Mexican and early Pueblo culture suggested that the first sedentary occupation of the northern Sierra occurred in late Pueblo 2 or early Pueblo 3 times by Southwestern people out of the Southwest, and (3) that the population density for the northern Sierra was greatest in Pueblo 3-4 times when the culture was disrupted by an invasion of eastern tribes which forced the people, living along the eastern flanks of the

basin and range area, into the high mountains of the Sierras. He believed that the eastern invaders became the Opata, Jova, and other tribes occupying the area when the Spaniards came (ibid., pp. 304-05). In the mid-1930's Kidder and the Cosgroves excavated at Pendleton Ruin near Cloverdale in southwestern New Mexico in hopes of finding a northern extension of the Casas Grandes culture (Kidder and Cosgrove, 1949). While Carey continued to work alone, students under Brand from the University of New Mexico in 1936 visited cave sites in the northern Sierra area and made several stratigraphic excavations in the high basin country on the eastern flank of the mountains some 55 miles northwest of Casas Grandes at the ruins of Agua Zarca and La Morita (Lister, 1939, 1946). Osborne and Hayes (1938) reported the expedition's activities across the international border in southeastern New Mexico. At the same time, Brand surveyed the Zape area in Durango in an attempt to correlate the Casas Grandes culture with Mesoamerica. Altogether, Brand and his students recorded some 500 sites in the northern Sierra country (Brand, 1939). Sayles' (1936b) reconnaissance of the area included some 200 sites not located by Brand. The purpose of the Gila Pueblo survey was to trace the southeastern extension of the Hohokam culture and to gain better insight into the general archaeological problems of the Chihuahua area. In summarizing his reconnaissance (ibid., pp. 85-88, Table 1) Sayles proposed a phase sequence. (1) The earliest defined occupation was a probable Mogollon-affiliated red-on-brown period called the Medanos phase. This culture was associated with camp sites in the sand dunes in the northeast corner of the north Sierra area. (2) Babicora Polychrome found in cliff dwellings and in adobe and stone pueblos marked the Babicora phase, which was probably inspired by the western 13

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

spread of the Caddoan Root. (3) This inspiration fostered the Ramos phase, which became highly developed at the ruins of Casas Grandes. (4) The Animas phase followed next and consisted of large pueblos of both stone and adobe which are associated with such northern ceramic intrusives as the Gila polychrome series, El Paso Polychrome, and Chupadero Black-onWhite. (5) Evidence in the subsequent Carretas phase intimated a time block when there was a break-up of the large pueblos into smaller house-units, located principally in the high upstream water ways of the eastern mountain area. (6) The Conchos phase was marked by Spanish contact. (7) The Lipan phase marked the Apache intrusion into the northern Sierra. In this same report, Gladwin (pp. 89105) recapitulated Sayles' findings and declared (1) that the Medanos phase probably dated around A.D. 1000-1100 and that the associated culture had a marked Mogollon relationship, (2) that the Babicora phase with its architectural and ceramic shifts was caused by influences from the Mimbres branch of the Southwest, and (3) the subsequent Ramos phase was inspired by the Salado people who staged an exodus from southern Arizona and migrated into Chihuahua. He reaffirmed this phase structure in 1957 and clung to the Salado migration theory as the inspiration for the Ramos phase great-houses found in the northern Sierra (Gladwin, 1957, pp. 128, 134-36, 294, 329-35). In 1941 the Mexican government published its archaeological atlas of the Mexican Republic (Marquina, 1939) and listed (pp. 55-67) the main sites located as of that date in the state of Chihuahua by Sayles, Brand, Lumholtz, and others. In the same year Wheat (1948-49, pp. 8-10) carried on surface reconnaissance in the vicinity of Ramos. In his report on the Chihuahua culture area (1943) Brand attempted to divide the "culture corridor" of the northern Sierra 14

country into geographical proveniences on the basis of river drainage systems. He was interested in three main problems: the extent of the dominant culture in the northern Sierras in time and space, its affiliations with surrounding areas, and the utilization of the area by the prehistoric inhabitants. In 1948 Lehmer published material gathered in 1940 on the Jornada branch of the Mogollon, in which he touched upon the northeast corner of the north Sierra in the vicinity of Villa Ahumada, and O'Neale (1948) reported on a group of textiles taken from Chihuahua. After a preliminary note (1953b) on his excavations in the Cave Valley area in 1951 and 1952, Lister (1958, pp. 110-19) disagreed with Brand's idea that the northern Sierra cultures originated in the river valleys along the eastern flanks of the mountains and migrated into the highlands to escape pressures from nomadic infiltration from the east. He also disagreed with the Gladwin school which proposed that the cultures originated in the highlands and moved to the eastern valleys and developed along Pueblo lines. He suggested instead that the early Medanos or red-on-brown phase be labeled Mogollon and that the later Casas Grandes phase material showed strong Southwest Puebloan influences after the bearers of the Mogollon culture drifted out of the western highlands into the river valleys of the eastern basin and range area. In his cultural-historical summary (ibid., pp. 112-15) he proposed three primary occupational phases: (1) The earliest level of culture preceded Mogollon 3, dated around A.D. 900; this lowest level was carried by a preceramic culture which had corn as well as certain stone core and flake tools. During this early horizon, the entire Sierra Madre Occidental served as a cultural corridor which connected Mexico and the American Southwest. (2) After this old, preceramic (?) level there appeared a Mogollon-like group which may date in the 10th century A.D. In this horizon the culture complex gave

FIG. 5—RUINS IN T H E NORTHERN SIERRA, a, Unexcavated large stone pueblo, one of many surface ruins in the narrow valleys of the high mountains. b,c, Room excavation of a cobblestone ruin in the high mountains. Note central post hole and subfloor fire hearth in b, the firebox, open ash pit, firedogs, and stone slab griddle of fire hearth in c. (Courtesy, Amerind Foundation.)

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

little evidence of house building in the cave sites, but contained a wide assortment of Mogollon-like traits. (3) The multistoried cliff houses were built by the same Mogollon-like people who were stimulated by the Pueblo culture of the Southwest. Later in the same horizon, these cliff dwellers migrated to the eastern valleys of the Casas Grandes, Santa Maria, and Carmen drainages and formed the dominant culture of the area. Lister likened this Mogollon absorption into the Pueblo pattern to development in the Mimbres area, where it is believed that a Mogollon group was absorbed by a Pueblo group to form the Classic Mimbres horizon. Mangelsdorf and Lister (1956) traced the distribution of maize through the northern Sierra and suggested that such types as Harinoso de Ocho spread through the area around A.D. 700. Kelley (1956, pp. 136-38) suggested, in view of recent archaeological studies in Durango and southern Chihuahua, that the dominant culture of the northern Sierras should be re-examined in the light of possible Mesoamerican origins, although he admitted that available knowledge of the area indicated strong Southwestern ties with the Pueblo area. Gerald (1957) excavated early Spanish material near Janos, and the next year Ascher and the Clunes dug Waterfall Cave in southern Chihuahua (Ascher and Clune, 1960; Clune, 1960). Cutler (1960) reported on the plant complex recovered during this work, and noted a close relationship between this area and the American Southwest. TENTATIVE ARCHAEOLOGICAL RECONSTRUCTION OF THE HISTORICAL CONTINUUM AND INVOLVED PROBLEMS OF THE NORTHERN SIERRA

The Joint Casas Grandes Expedition, operating under the auspices of the Amerind Foundation and the Instituto Nacional de Antropología e Historia, has been actively engaged in field work in the northern Sierra 16

district since September 1958. It is hoped that future excavations will expose the historical continuum of the area. The focal point of this study is the ruin of Casas Grandes, or the city of Paquime, as called by Obregón in the 16th century (Hammond and Rey, 1928, pp. 205-08) (figs. 4, 8, 9). To date, some 20 hectares, approximately 2/5 of the city, have been completely excavated.17 In addition, three nearby sites have been examined for data to provide an accurate historical reconstruction of the river valley cultures. Other sites in this vicinity have been surveyed and coordinated with the work of Brand, Carey, Sayles, and Lister. The Parral archives are being indexed and microfilmed for information vital to an archaeohistorical approach. Little ethnohistory is available for filling the gaps which will occur in the final analyses of these studies. Solely on the basis of current field work we offer the following very generalized and most tentative outline of historical continuum. Preceramic Horizon The early hunters and foragers who may have roamed the northern Sierra have left little evidence with which to reconstruct the past. There should be cave sites in the mountains and camp sites on the old beaches of the now dry lakes in the basin and range area that may provide further finds. Lister (1958) has uncovered what may prove to be a preceramic level in Cave Valley in the Piedras Verdes drainage. Students from the University of New Mexico under the leadership of Reiter reported "a possible Clovis Fluted point" and other tools (Marrs, June 11, 1960, personal communication). The Joint Casas Grandes Ex17 The excavations have exposed five ceremonial mounds, two large water reservoirs and an acequia system, a ball court, a large town plaza, as well as 200 secular rooms and 16 intra-house plazas, which represent approximately 10 per cent of the total number of dwelling rooms.

ARCHAEOLOGY & ETHNOHISTORY: NORTHERN SIERRA

FIG. 6—ARROYO D E MALPAIS SITE, a, Level of the Pleistocene horse-kill, b, Horse jaw, preceramic horizon. (Courtesy, Amerind Foundation. Photo, Jimmy Cohn.)

pedition has located what appears to be a Pleistocene horse-kill site (fig. 6), but no tools or fire-hearths in association. The trail of the ancient hunters remains tantalizingly elusive. CASADOR PERIOD. Early in the history of the northern Sierras man roamed the area as a hunter, perhaps tracking Pleistocene fauna and consuming Pleistocene flora. Mayhap these ancient hunters were related to the hunters of southern Arizona who produced Clovis Fluted points and ate the meat of mammoth and other Pleistocene beasts (Haury et al, 1959). FORRAJE PERIOD. Evidence for the presence of foragers in the Sierras rests solely on Lister's cave finds (1958, pp. 15-22, 96108, 112) in the Piedras Verdes valley. He reports a preceramic stratum which contained pre-chapalote corncobs, acorn shells, stone flake and core tools, but no ceramics.

The dating of either the Casador or the Forraje period is hindered by lack of definite data. Bands who derived their sustenance from hunting and foraging may have occupied the area throughout its entire continuum, living side by side with the more advanced cultures. Village-Farming

Community

Horizon18

VIEJO PERIOD. The Joint Casas Grandes Expedition has recently excavated an open site in the Casas Grandes yalley which, in many respects, correlates closely with Lister's cave studies (Lister, 1958). Some segment of the northern Sierra population seems to have attained the level of a village18

This form is taken from Braidwood's study of Near Eastern prehistory (1958) because it apparently best fits the general scheme of growth in the northern Sierra district. Also see Willey (1960).

17

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

farming community, a period that can be broken down into three phases based primarily on architectural stratigraphy, shifts in house forms, and changes in ceramics. (For the purposes of this article, the following descriptions are not only tentative but brief and oversimplified.) The Viejo period represents a time when the population in both the mountain and the basin and range areas gathered in small villages. Their residue culture apparently resembles (according to Lister, 1958, and Gladwin, 1957) that of the Mogollon, which has been defined as part of the cultural complex of Arizona and New Mexico (Wheat, 1955). In my opinion the Viejo period remains thus far found were rooted in the earlier Cochise culture (Sayles and Antevs, 1941) and related to the Ootam or desert Mogollon culture (DiPeso, 1956) of southern Arizona. There is a strong possibility that this low-level culture was extremely widespread, having both geographic continuity and perhaps also a common linguistic base, covering an area from the Mogollón Mountain area in the United States south through the northern Sierra country of Chihuahua and perhaps into Durango. It might be wiser, therefore, to use a more general term for this culture than "Mogollon," as has been proposed by Lister (1958, pp. 11219). This is a minor but nevertheless important point, which may be resolved in further study. The presence of such a group over a wide area may well explain the early movements and interchange of ideas and perhaps of people between central Mexico and the United States. In this period, the natives built small villages on high points of land. The settlement pattern is one of small houses-in-pits built at random around a larger ceremonial structure, which would be defined as a Great Kiva in Mogollon terminology. Apparently these early inhabitants also used caves in the mountain area. It appears that this culture dominated the northern Sierra district 18

for a considerable period, an occupation defined as Sayles' Medanos phase (Sayles, 1936b, p. 84, Table 1) and Lister's Mogollon 3-4 phases (Lister, 1958, pp. 112-19). The period covers three phases. Convento Phase. In this phase, the earliest definitive phase yet uncovered by the Expedition, the village-farming community consisted of a few scattered houses-in-pits located on high points of land near available water. The houses were small (averaging 11 sq. m.) and were round, oval or beanshaped. The side, stepped entries were not always oriented to the east. A simple firepit lay in the center of the rooms; around the interior periphery of the house depression were closely aligned post holes. This indicates a post-reinforced wall which probably formed a dome-shaped, earth-covered dwelling. Large pits scattered between the houses apparently were used for storage as well as for cooking. Burials, consisting of semiflexed individuals with no associated funerary furniture, lay at random around the village premises, often in the fill of old pits. The pottery of this period was a red-brown ware often decorated with broad-lined-onfinger-applied, rectangular elements. These open-line designs took the shape of pendant triangles on the shoulders of jars. The necks of this brown ware were designed with a multitude of intricate, textured patterns. A red-on-brown decorated bowl form, finished by polishing over the decoration and thus smearing it, is also associated with this phase. There is a close resemblance between these ceramic types and those defined as Mogollon. One notable lack, mentioned by Lister (1958, p. 110), is absence of a redslipped pottery. No trade pottery has yet been definitely associated with this phase. The lithic complement included crude boulder metates in the forms of slab, basin, and horseshoe troughs. The rectangular block manos, as well as taper manos, again appear to be related to the Ootam culture of the north.

FIG. 7—CONVENTO SITE. This site produced evidence of the Spanish colonial period. Church of San Antonio de Padua (see excavations along right half of photograph) was destroyed by native revolt in 1684. Early pit houses and surface dwellings of Viejo period in left half. (Courtesy, Amerind Foundation. Photo, Tommy Carroll.)

On the floors of these rooms were corn kernels, which suggest that agriculture was practiced to a certain extent. Pilon Phase. In this phase the village settlement pattern and location remain the same but the house forms change. Neat circular pit houses (averaging 21 sq. m.) now overlie the earlier, smaller houses-inpits. Neat plastered firehearths are encountered near the doorways, which remain mere step entries. The ceremonial structure apparently was enlarged during this phase. Burials remain the same save for the ap-

pearance of grave furniture. The natives buried their dead both in abandoned pits and in abandoned houses. The ceramic shift consists of executing the design in finer line, adding a white slip, and omitting to rub over the design with a polishing stone. Trade pottery such as Mangus Black-on-White appears to have come in from the north. The lithic tools show little or no change. Perros Bravos Phase. The latest phase in this period is marked by an interesting shift in house forms and in trade pottery. The 19

FIG. 8—CASAS GRANDES RUIN, CHIHUAHUA. Excavations at lower right are of Medio period structures. Multistoried structures are of Tardio period. Ball court and other ceremonial mounds surround part of the town court, Central Plaza. (Courtesy, Amerind Foundation. Photo, Tommy Carroll.)

houses are rectangular surface houses, constructed in clusters of two or three rooms tied together by wing walls around small plazas. The room walls are raised on puddled-adobe footings. The trade pottery is Mimbres Black-on-White and Mimbres Polychrome, Three Rivers Red-on-Terracotta, and Lincoln Black-on-Red. There is little apparent change in the local ceramics. Burial customs and the lithic complement remain the same. MEDIO PERIOD. The Medio period belongs to the village-farming community horizon, 20

but shows rapid and very noticeable progress in architecture and ceramics. The houses of the Viejo period, Perros Bravos phase, continue in vogue, but there is a drastic shift in ceramic styles as the Casas Grandes polychrome cluster, a red ware as well as a black ware, are added to the older brown ware series. Burials are placed under the floors of houses and are accompanied by a greater amount of funerary furniture. One of the outstanding changes of this period is the development from simple house clusters of two or three rooms built on puddled-

ARCHAEOLOGY & ETHNOHISTORY: NORTHERN SIERRA

adobe footings to massive single-story house clusters. This change in housing structure and the introduction of a water system employing acequias (ditches supplying city water), canals (ditches irrigating fields), and subterranean drains suggest an increase in population as well as an advance in engineering. The house clusters built toward the end of this period indicate a defensive form of settlement pattern; they have high outer walls staggered for more effective protection, and corner bastions. The introduction of stairs with landings and stairwells, the idea of the plumed-serpent motif, and the retention of the Great Kiva as an integral part of village construction, all mark this period as one of great change and growth. Reyes Phase. The Perros Bravos phase house form continues, but the ceramic type shifts and burials are placed under floors. It is at this time that the red wares and black wares are added to the growing textured brown ware tradition. The shift from bichrome decoration to polychrome, which marks the Casas Grandes culture, is a crucial clue to the changing way of life of the indigenous peoples, but presents a serious problem as to the reason for this shift. Buena Fe Phase. The salient characteristic of this phase is the drastic shift in architectural features. The building of house-clusters of puddled adobe is emphasized. New traits such as T-shaped doorways, raised firehearths, staircases, and alcove platform beds are introduced. A subterranean room is retained in the ceremonial architecture but is now rectangular instead of round. In addition to the introduction of the plumedserpent motif, special bins in the plazas for the breeding and raising of the macaw indicate that the natives practiced animal husbandry. The lithic complement changes noticeably. There is an apparent increase in population in the valley area. These developments point to a new and vigorous force at work in the valley culture. A major question, still unanswered, is Where did this inspiration come from? It

appears to be associated with such northern trade pottery as the Little Colorado polychrome series, including Saint Johns as well as El Paso Polychromes and Chupadero Black-on-White. The absence of the Gila polychrome series is significant and perhaps indicates that this inspiration was not engendered by a Salado movement as proposed by Gladwin (1957; also see his discussion in Sayles, 1936b, pp. 101-05). There are many suggestions besides ceramics that the local culture was developing at its own rate, inspired by a strong impetus from the south. Urban Civilization Horizon19 TARDIO PERIOD. It may appear presumptive to suggest that the culture of the Casas Grandes valley can be described as an urban civilization. No evidence of writing has been found, but the buildings of Casas Grandes became an actual city in the Tardio period (fig. 9). The settlement pattern indicates a master building plan which covered more than 50 hectares and possibly accommodated over 5000 inhabitants. The formal placement of ceremonial mounds, the presence of guild or artisan classes, the utilization of a city water system, and a well-maintained trail system with pyro intercommunication all support this tentative assumption. This period sees a sharp change in ceremonial architecture. Many architectural features described by Brand (1939, p. 96) as belonging to the Quemada, or Zacatecas, culture appear at the city of Paquime. Such items as a town court, round stone houses, stairways, pyramidal structures, ball courts, human sacrifice, and truncated mounds are present in the Tardio period. Perhaps the archaeology of Casas Grandes is atypical for the entire northern Sierra, as the city may have been the largest trading center in the area, but its power over the district and its influence on the cultures of the Southwest may have been tremendous.

19 Braidwood, 1958, p. 1419-20.

21

FIG. 9—CASAS GRANDES, CHIHUAHUA. Part of the Tardio period multistoried construction. (Courtesy, Amerind Foundation. Photo, Tommy Carroll.)

The archaeological evidence strongly suggests that in the Tardio period the older Medio period city settlement pattern, which was a series of puddled-adobe house clusters scattered over a small area, was radically modified. Those clusters in the areas where the Tardio period ceremonial mounds were built were purposely abandoned. The city acequia system was revamped to accommodate the building of a two-bodied truncated mound which was part of a ball-court complex. Those Medio period house clusters in the secular dwelling areas were remodeled or razed to accommodate the massive multi22

storied constructions of this period of urbanization. It would appear from running building walls, placement of plaza underground drains, and methods of amassing and utilizing building supplies that a master plan was put into eflFect by some controlling force. It seems that this force used the knowledge of construction already present in the area but introduced many advanced ideas of architecture. These innovations in the Tardio period included a change in settlement pattern, the construction of multistoried apartments, the addition of ceremonial mounds, ball courts, as well as a

ARCHAEOLOGY & ETHNOHISTORY: NORTHERN SIERRA

town court, and the practice of secondary burials placed in large jars and accompanied by human-bone phalangeal necklaces and human longbone musical rasps. All this evidence permits the speculation that a powerful religious group from central Mexico gained control of the Casas Grandes valley and imposed their leadership on the people. The massive construction, the production of food, the maintenance of the water system, and the production and distribution of trade goods called for an incalculable number of man-hours. Nor can one estimate the number of individuals needed to maintain protection for the populace. The great building florescence of the Tardio period is marked by excessive group energy in the first, or Paquime, phase of the period. The second phase, the Diablo, is marked by a violent secularization of the temples. Many of the grand galleries were partitioned off to make small living rooms; old doorways were sealed and new ones opened. The end of the period came with the burning of the great metropolis. The presence of human bodies scattered in the fill of the collapsed rooms, the suggestion of profaning the temple areas by breakage of idols and other religious items, and a murmur of Opata myth (Bandelier, 1892, pp. 520-21, 571) suggest that a war brought on the final collapse of a greatly advanced culture. This supposition was also mentioned in Obregón's journal (Hammond and Rey, 1928, pp. 207-08). He suggested that the city met its end through tribal warfare or, perhaps more accurately, through internal revolt which worked to the advantage of neighboring tribes who sacked the city. The question of remnant population is a moot point which may never be settled as there apparently is a break between the original population which built Casas Grandes and the rude tribesmen met by the Spaniards when they first traveled through the valley of Paquime. Paquime Phase. This phase of the Tardio period is marked by a sharp change in the

settlement pattern of the city of Paquime. The house clusters of the Medio period are abandoned, razed, or remodeled to fit the new master plan of multistoried areas which surrounded a town court and the building of ceremonial mound structures and a ball court. The ceramic pattern indicates an interesting shift in northern trade wares as Gila, Tonto, and Tucson polychromes became popular and the local Ramos Polychrome artisans became strongly influenced by these trade pottery designs. The use of copper in the form of needles, bells of cire-perdue, and pseudocloisonné become popular. The artistic designs begin to feature plumed serpents as well as the macaw, and a ceramic hand drum decorated in red and green paint appears. These changes form a cluster of traits surrounding a religious idea which affected the indigenous populace. Diablo Phase. This phase contains architectural changes that are clues to an internal breakdown of the city of Paquime. Temple rooms and long, impressive galleries are broken up by poorly made partition walls; firehearths are added; temples may well have been profaned by the working classes. The mute evidence of a burned city and the remains of human bodies in the collapse of the great buildings, partially charred and scattered near doorways, suggest the disastrous end of a culture which had approached the threshold of civilization only to weaken and fall. Internal revolt, perhaps, dealt the first blow, and then a coup de grâce struck by surrounding tribes which moved in to complete the final destruction. Historic Contact Horizon After the fall of the city of Paquime, we know little of what occurred in the northern Sierra area. The innumerable outlying villages may have remained active for a time, but survey suggests that many were also burned. The possibility of dual occupancy in the area by two different cultures remains to be worked out; there is but a hint of 23

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

group movements after the destruction of the city. SPORADIC CONTACT PERIOD. This is a period when the first Spaniards entered the area and left written records. Perhaps Cabeza de Vaca's journals will contain information on this matter; certainly General Francisco Ibarra's reports are important in filling the gaps. There are no self-identified Indian groups in the area today which retain any folklore or mythology applicable to this problem. The scanty Opata myths gathered by Bandelier (1892) may prove helpful. SPANISH CONTACT PERIOD ( A.D. 1660-84).

40 years later along the eastern portion. This period is still little known.20 The current program of microfilming the Parral archives should afford pertinent data. A native revolt in 1684 forced the Spaniards out of the northern Sierra country and permitted entrance of Apache bands which held control. APACHE PERIOD (POST-1684). This period is mentioned merely as part of the historical continuum of the northern Sierra district. For over 200 years these wild Apache bands dominated an area which once supported a culture that had approached the level of a civilization and then collapsed.

Actual contact—settlement of Spaniards in the northern Sierra district—begins in the 1620's along the western margin and some

20 See Gerald, 1957, which covers the Presidio of Janos, and A. E. Hughes, 1935.

REFERENCES Almada, 1945 Amsden, 1928 Ascher and Clune, 1960 Bancroft, 1886 Bandelier, 1880-85, 1890a, 1890b, 1892 Bannon, 1955 Bartlett, 1854 Beals, 1932 Benavides, 1954 Blackiston, 1905, 1906a, 1906b, 1909 Bolton, 1936 Bourke, 1886 Bradfield, 1923, 1931 Braidwood, 1958 Brand, 1932, 1935, 1936, 1937, 1938, 1939, 1943 and Harvey, 1939 Breternitz, 1959 Cabeza de Vaca, 1871 Carey, 1931, 1951, 1954, 1955 Chapman, 1923 Clavigero, 1787 Clune, 1960 Cosgrove, H. S. and C. B., 1932 Cutler, 1960 Decorme, 1941 DiPeso, 1951, 1956 and Cutler, 1958 Dunne, 1940, 1944, 1948 Escudero, 1834 Forbes, 1957, 1960

24

Forrestal, 1954 García Conde, 1842, 1849 Gerald, 1957 Gladwin, 1957 Goldman, 1951 Guillemin-Tarayre, 1867, 1869 Hackett, 1923-37 Hammond and Rey, 1928 Harcum, 1923 Hardy, 1829 Hatch, 1954 Haury et al., 1959 Hawley, 1936 Hewett, 1908, 1923, 1930 Hough, 1923 Hrdlicka, 1904 Hughes, A. E., 1935 Johnson, J. B., 1950 Kelley, 1956 Kidder, 1916, 1924, 1939 and Cosgrove, 1949 , Jennings, and Shook, 1946 Kinnaird, 1958 Lehmer, 1948 Linné, 1942 Lister, 1939, 1946, 1953b, 1958 Lombardo, 1702 Lumholtz, 1902, 1912 Mangelsdorf and Lister, 1956 Maps, 1904, 1948, 1958

ARCHAEOLOGY & ETHNOHISTORY: NORTHERN SIERRA

Marquina, 1928, 1939 Martin et a l , 1952, 1954 Matson and Schroeder, 1957 Monterde, 1842-45 Nesbitt, 1931 Noguera, 1926, 1930 Núñez, 1777 Ocaranza, 1933 O'Neale, 1948 Orozco y Berra, 1864 Osborne and Hayes, 1938 Pinart, n.d. Rivera, 1736 Robles, 1929

Sauer, 1932, 1934, 1935 and Brand, 1930, 1931 Sayles, 1936a, 1936b and Antevs, 1941 Schwartzlose, 1952 Schwatka, 1893 Swadesh, 1959 Tassin, 1902 Tower, 1945 Wheat, 1948-49, 1955 Willey, 1956a, 1960 Zingg, 1940 Zuñiga, 1835

25

2. Archaeology of Sonora, Mexico

ALFRED

HE STATE of Sonora (fig. 1), in northT west Mexico, is bounded on the north by the international boundary with the United States, separating Sonora from Arizona and New Mexico; on the west by the Gulf of California, except for a short distance at the extreme northwest corner of the state where it is separated from Baja California Norte by the lower portion and delta of the Colorado River; on the east by the state of Chihuahua; on the south by Sinaloa.1 Sonora's 70,477 square miles fall within three distinct biotic provinces (Dice, 1943, p. 3). Northwest Sonora lies in the Sonoran biotic province, which extends into southwest Arizona, southeast California, and northeast Baja California Norte. The Sonoran province is largely desert, its extensive plains separated by isolated mountain 1 Acknowledgment is made to Emil W. Haury and Raymond H. Thompson, of the Department of Anthropology at the University of Arizona, for critical review of the manuscript and helpful suggestions. The ceramic figurines illustrated in fig. 2 are from the collection of Edward and Rebecca Moser, to whom thanks are due for the loan of these figurines to the Arizona State Museum for study and photography.

26

E.

JOHNSON

masses. Summers are long, and temperatures are high. Winters are moderate with occasional frosts. Winter rains are widespread and gentle; summer rains are restricted and of the thunderstorm type. The Sonoran province is well covered by vegetation, with cacti and thorny shrubs conspicuous (Dice, 1943, pp. 51-54). Northeast Sonora lies within the Apachian biotic province, which also includes parts of Arizona, New Mexico, and Chihuahua. This province is characterized by high grassy plains, ranging in elevation from 4000 to 7000 feet, and by interspersed mountainous zones, which occasionally rise to over 9000 feet. The rainfall pattern, like that of the Sonoran province, has mild winter rains and summer thunderstorms. Because of the higher elevations, the mean temperatures are somewhat more moderate, although summer days can be quite hot, and periods of extreme cold occur in winter (Dice, 1943, pp. 56-58). Southern Sonora is in the Sinaloan biotic province, which also extends into northwestern and central Sinaloa, and for some distance up the Yaqui River valley. Frosts do not occur, a feature that sets this province

FIG. 1—SONORA, WITH PLACE NAMES MENTIONED IN TEXT (Courtesy, Arizona State Museum.)

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

apart from those to the north. Rainfall is abundant in the summer, but generally lacking the rest of the year. The characteristic vegetation is thorn forest, dominated by Acacia cymbispina (Dice, 1943, pp. 55-56). At least some of the cultural variation noted below is associated with the varying ecological zones of Sonora. Other natural features of the state cross-cut the ecological zones and serve to relate otherwise disparate sections. Several of these features are the major river systems which cross-cut the biotic provinces: from north to south, the Magdalena, Sonora, Yaqui, and Mayo. These river valleys undoubtedly served as corridors along which peoples and ideas moved throughout northwest Mexico and, perhaps, into the Southwest United States. Another geographic feature which has tended to relate disparate areas, and perhaps peoples, is the long coast of the Gulf of California. As noted below in the summary of this coastal region, there are certain similarities along this entire section. As of 1960, nine distinct Indian groups resided in, or partially in, the state: the Cocopa, Papago, Jova, Opata, Seri, Warihio, Lower Pima, Yaqui, and Mayo (Spicer, 1962, Map 1). Ethnographic data have been collected from several of these groups and historical research in governmental and church archives has also been undertaken (Spicer, 1962). Ethnohistoric studies have yet to be made. HISTORY OF ARCHAEOLOGICAL RESEARCH

Passing notice of the archaeological remains in Sonora has been taken since early Spanish exploration. Manje (1954, p. 21) recorded his visit to the Soba Pima in February 1694. We said goodbye and continued our journey toward the east along the river, through plains. We passed in sight of, and around, a mountain where there are 100 terraces of stone walls in the form of a snail, spiraling to the top. They say it forms an armory, where in former wars those who gained the heights first were usually victors. Those who reached the first ring went 28

around to the second, and as far as was necessary to exhaust the supply of arrows of those below. Then they came down from the mountain and fell upon their enemies and killed them. After having gone 10 leagues, we camped at the river, which we named Santa Maria de Toava. This river now bears the name Magdalena. The terraced mountain is probably the great site of Las Trincheras immediately south of the modern community of Trincheras in northwest Sonora. Terraced hillside sites will be discusssed below. Such descriptions, scattered through the documentary material on Sonora, have stimulated archaeological research, but seldom include enough specific information to be usable in interpretation. Formal archaeological work in Sonora began with a survey of the northeast part by Adolph Bandelier during the 1880's (1892, pp. 480530), followed by the Lumholtz expeditions of the 1890's and early 1900's (1902, pp. 1-59; 1912, pp. 134-288), first in northeast Sonora, then in the northwest. During the 1920's surveys were conducted by Monroe Amsden (1928) in the northeast, and by Gila Pueblo in the northwest (Gladwin and Gladwin, 1929). The data gathered by Sauer and Brand (1931) came mostly from a survey of northern Sonora in 1930, some in 1928. Ekholm, in the late 1930's, conducted a more extensive expedition, which extended into Sinaloa (1939). Danson's survey (1941) of the upper Santa Cruz River valley included part of Sonora. Gifford (1946) recorded sites along the coast of the Gulf of California, especially in the vicinity of Punta Peñasco in 1944 and 1945. Lehmer (1949) made a reconnaissance of Sonora in the late 1940's. Additional data on the archaeology of Sonora, particularly the coastal region, were gathered by Fay (1953, 1959) between 1953 and 1958. Ezell's survey (1954) of Papagueria in the early 1950's extended into Sonora and located a number of sites in the northwest. In 1953 Lister (1958, pp. 41-57)

ARCHAEOLOGY: SONORA

led a field party in the test excavation of a number of caves along Arroyo el Concho in northeast Sonora. Hinton (1955) surveyed a section of the Altar valley in the northwest in 1954. Noguera (1958) directed an archaeological reconnaissance which covered a large portion of the state in the late 1950's. A. E. Johnson (1960, 1963) made a survey and minor excavation at La Playa site in 1959. The references listed above include most of the major archaeological projects reported for Sonora. Additional notes, scattered through the literature, will be mentioned in the following discussions. Other references, not cited in the text but listed at the end of this article, attempt to cover the subject as completely as possible. Nearly all our information comes from surface reconnaissance, most of this in the northern half of the state. The few excavations conducted have been on a small scale or have not been reported in detail. The resulting data are insufficient for an accurate delineation of cultural provinces, but to facilitate the following summary, the state has been somewhat arbitrarily subdivided into regions and cultural periods. PRECERAMIC REMAINS

Only a little information indicates that Sonora was occupied by the big-game hunters responsible for the Llano complex (Sellards, 1952, pp. 17-46). Supporting this are two Clovis projectile points from a blowout some 30 miles north of Guaymas near Punta Blanca (DiPeso, 1955). F. Η. Η. Roberts (1945, p. 147) mentions two Folsomlike points from Sonora. Llano complex sites are better known from just over the border in southern Arizona, where they are dated to around 900Q B.C. (Haury, Sayles, and Wasley, 1959). Sites of the Cochise culture, a regional variant of the widespread Desert culture (Jennings, 1957, p. 282), are also present in Sonora. A Peralta complex of the Cochise culture has been identified from sites near

Pozo Peralta on the Sonora River west of Hermosillo (Fay, 1955; 1956a; 1957a; 1957b; 1958; 1959). The Peralta artifact complex includes crude choppers, flake tools, occasional projectile points, basin metates, and hand stones. To date, only surface remains are known, and the stratigraphic associations have not been determined. The closest typological similarities are with San Pedro stage Cochise materials from southern Arizona, making it likely that the Peralta complex dates to between 1500 B.C. and A.D. 1. Although remains attributable to the earlier Sulphur Spring and Chiricahua stages of the Cochise culture (Sayles and Antevs, 1941, pp. 11-20) have not yet been identified in Sonora, the possibility of their presence is suggested by data collected from the Rio Mayo area. In this region, Ekholm (1940, pp. 326-27) recorded sites, on tributaries of the Rio Mayo, that lacked pottery and had a stone complex consisting of a few chipped implements, slab metates, and hand stones. There was an indication that these artifacts were from deposits which also included Pleistocene fauna. Another find from this region was a fossil human cranium, also from deposits with extinct mammals (Aveleyra, 1962, p. 17). In Arizona Cochise sites, slab metates are early in the sequence (Sayles and Antevs, 1941, fig. 11), and Sulphur Spring stage artifacts have been found in association with extinct mammals (ibid., p. 8; Haury, 1960). Other indications of a preceramic period in the Sonoran sequence of development are several lithic sites listed in the Arizona State Museum Archaeological Survey files. These sites are known from surface exposures in the northern part of the state. Artifact assemblages include hand stones, choppers, scrapers, basin metates, and occasional projectile points. Ives (1963a) has recently called attention to the association of shell mounds and old shorelines along the Gulf of California. This association suggests some interesting possi29

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

bilities for stratigraphic studies, particularly since some of the shell middens along the Sonoran coast have pottery throughout the entire mound (Hayden, 1956), others include pottery only in their upper layers (Ives, 1963a), and others lack ceramic remains (Hayden, 1956). A problem involved in the investigation of preceramic remains in Sonora was noted by Ezell (1954, p. 15) during his survey of northwest Papagueria, and it was also apparent from the study of the artifacts from the La Playa site (A. E. Johnson, 1963): the late retention of early tool types. At La Playa the stone artifacts had many similarities to the lithic assemblage of the San Pedro stage of the Cochise culture, as known from southern Arizona (Sayles and Antevs, 1941, pp. 21-26), but the La Playa artifacts occurred with pottery, indicating a much later date. The placement of lithic assemblages from Sonora therefore should be cautiously interpreted as, in special cases, stone tools which appear to be typologically early can actually be quite late. COASTAL SONORA

The long coastline of Sonora extends some 600 miles from the mouth of the Colorado River to the northern border of Sinaloa. Although probably inhabited by varying peoples at different times, this coastal region has certain archaeological features which tend to relate the entire zone. For this reason, and because of insufficient research to delineate archaeological provinces, the Sonoran coast is here considered as a unit. Prominent among the archaeological features are shell middens. Although the entire coast has not been surveyed for this purpose, enough work has been done to indicate that middens are distributed from the Punta Peñasco region (Gifford, 1946) to the Sinaloa border and south (Fay, 1953). The middens in the vicinity of Punta Peñasco have little depth, but instead have widely distributed shell remains. Occasional stone implements such as manos, metates, 30

and crude flake tools occur. Pottery types include Sacaton Red-on-buff, Gila Plain, Sells Plain, Lower Colorado Buff ware, and Trincheras Purple-on-red. Sacaton Red-onbuff and Gila Plain are River Hohokam types, probably from the Gila Basin of southern Arizona (Gladwin et al., 1937, pp. 171-78, 205-11). Sells Plain is the utility type of the Desert Hohokam of southwest Arizona (Haury, 1945a, pp. 6-12). Lower Colorado Buff ware, including a number of types, was made along the lower portions of the Colorado and Gila rivers (Schroeder, 1952, pp. 7-35). Trincheras Purple-on-red is the most common painted type associated with the Trincheras culture of northwest Sonora (A. E. Johnson, 1960, pp. 62-66). Two of the types, Sacaton Red-on-buff and Trincheras Purple-on-red, have been placed chronologically with some accuracy, so their association in the shell middens is suggestive of a temporal placement for the middens. Sacaton Red-on-buff was made between A.D. 900 and 1100 (Gladwin et al., 1937, fig. 106); Trincheras Purple-on-red occurs between A.D. 800 and 1100 (A. E. Johnson, 1963, p. 184). The discovery of a variety of types in the middens, representative of several cultural provinces in southern Arizona and northern Sonora, allows alternative explanations. The sites are apparently temporary, to judge from the lack of any depth of trash or evidence of permanent habitation. Therefore, they could be occasional camps of nomadic groups who dwelt along the coast and maintained trade relations with inland groups. Shell was widely used for ornaments in all the regions from which the pottery is believed to have come, so it is possible that the coastal groups were trading shells to the neighboring inland peoples. The other possibility is that the sites represent temporary camps of inland groups who visited the coast to gather shells. Shell mounds occur farther south along the coast, in the area about Bahia Kino, opposite Isla Tiburon. Few stone tools have

ARCHAEOLOGY: SONORA

appeared in surface collections from these mounds, although pottery is fairly common (Fay, 1953). The pottery is apparently of the same types as found between Bahia Kino and Guaymas, particularly near Estero de Tastiota (Owen, 1956), where two types have been recovered from the shell mounds. One is plain, with coarse sand temper and prominent scraping striations. The other is thin, hard, and buff to buff-gray; it is often polished and sometimes has a white slip. Besides shell middens, the area around Estero de Tastiota has stone arrangements which may represent shelter outlines (Hayden, 1956). A collection of surface artifacts described for an Estero de Tastiota site includes a large number of triangular projectile points. Unnotched, side-notched, and cornernotched varieties of the obsidian and basalt points occur. Some are serrated. Although a few are similar to Pinto and San Pedro styles, their association with pottery at the site makes it likely that they are of more recent manufacture. The pottery is generally similar to that mentioned above, but has a little more variety, including a type with brown paste and dark, organic tempering. Other artifacts are shell beads and pendants, basin metates, and hand stones (Holzkamper, 1956). A ceramic figurine complex occurs in the general region from Bahia Kino to Estero de Tastiota and north to the Desemboque vicinity. Other specimens have been found on Isla Tiburon and in Baja California (Dockstader, 1961). To date, all figurines have been surface finds from shell midden areas (Owen, 1956, p. 1). Both single examples and caches occur (Dockstader, 1961). Most of the specimens are handmodeled females, with expanded thorax, narrow pointed head, and either pointed base, bifurcated base, or rounded pedestal base. Decorations, perhaps representing tattoos, have been incised, probably by means of paired cactus spines. Red paint occasionally fills the areas between the lines

FIG. 2—CERAMIC FIGURINES FROM THE CENTRAL COAST OF SONORA. (Courtesy, Arizona State Museum.)

(fig. 2). A typology of the figurine varieties has been suggested by Owen (1956). Little is known of the history of these figurines. They occur within the historic Seri area (ibid.) but cannot be definitely attributed to the modern Seri as identical specimens do not occur in recent ethnological collections. Modern Seri figurines of wood and of clay do bear some resemblances. The paste is similar to the thin, hard, often scraped or brushed pottery found in the shell middens of this region, although it sometimes includes a high proportion of organic material as does the paste of recent Seri pottery. A number of investigators at present believe that both the hard, thin, scraped pottery and the figurines of the same paste mark a prehistoric Seri occupation (Fay, 1956b; Owen, 1956). Additional shell mounds have been recorded for the regions about Guaymas, Huatabampo, and Agiabampo, Sonora. They also extend south into Sinaloa (Fay, 1953). Beals (1942) has noted that the mounds extending south from the Mayo River to Playa San Ignacio in Sinaloa have a maximum height of 3 or 4 feet. Sherds collected from these mounds are similar to modern Mayo Indian pottery. NORTHWEST SONORA

A distinctive complex of archaeological traits, known as the Trincheras culture, is found throughout the northern half of 31

FIG. 3—EASTERN E N D O F T H E SITE O F LAS TRINCHERAS. (Courtesy, Arizona State Museum.)

Sonora. Recognized since the time of Bandelier (1892), the complex is marked by terraced hillside sites and Trincheras Purpleon-red pottery (A. E. Johnson, 1963). In Sonora, Trincheras sites spread from the foothills of the Sierra Madre (Bandelier, 1892, pp. 524-25) nearly to the coast, and from the international border south to about the middle of the state, being most abundant in the Magdalena basin, including the Altar valley in northwestern Sonora (Sauer and Brand, 1931, pp. 87-100; Hinton, 1955). The characteristic sites are situated on isolated hills or prominent knolls in mountainous sections (Sauer and Brand, 1931, pp. 87-100), but sometimes also in valleys (A. E. Johnson, 1963). The hillside sites are marked by dry-laid masonry walls rising in terraces up the hillsides. The walls retain 32

a rubble fill on which are constructed small houses, again with walls of dry-laid masonry. The back wall of the house is usually the retaining wall of the terrace above. Besides the terrace houses, many sites have a compound, either circular or rectangular, often at the crest of the hill. The compound may or may not enclose additional houses. The most elaborate of the terraced hillside sites is Las Trincheras itself (fig. 3), in northwest Sonora, just south of the modern village of Trincheras (Sauer and Brand, 1931, pp. 91-92). Depending on the height of different parts of the hill, from 15 to 30 levels extend from the base to the summit. Rectangular stone enclosures appear both at the summit and farther down the slope. Numerous house outlines are spread about the terraces; an extended oc-

ARCHAEOLOGY: SONORA

cupation is evidenced by quantities of sherds, stone tools, and shell ornaments. Of all the possible functions for trincheras sites, the one offered most consistently is that of defense, which seems obvious from their locations and from the fortlike structures which often crown their summits (Arizona State Museum Survey files). The purpose of defense is supported by descriptions and legends collected by early explorers in Sonora (Pfefferkorn, 1949, p. 154; Bandelier, 1890b, p. 71). Some sites seem to have served only during times of attack as there is no evidence of continued occupation (A. E. Johnson, 1963, p. 179); others, such as Las Trincheras, were perhaps occupied the year round, to judge from the quantity of trash accumulated on the terraces. Against whom the Trincheras people were defending themselves is unknown. Trincheras sites should not be confused with the systems of check dams or agricultural terraces in the Sierra Madre east of the Trincheras region (Lumholtz, 1902, pp. 21-22). The Sierra Madre features, apparently constructed to halt soil and water runoff, lack the evidence of houses on the Trincheras sites. R. B. Woodbury (1961, p. 11) has suggested the term terraces for erosion control devices, trincheras for habitation and defensive sites. Trincheras culture sites are also present in the intermontane valleys of northern Sonora. One of the largest is La Playa (A. E. Johnson, 1960, 1963), 27 miles west of Santa Ana, Sonora, on the old road from Santa Ana to Altar. La Playa occupies a large flat at the base of the Cerros Boquillas, which are to the north. The site was evidently a cienega or marsh at the time of occupation, with a permanent supply of water from the springs at Ocuca. Recently, La Playa has undergone extensive sheet and gully erosion, so most of the archaeological features have been destroyed. Archaeological features at La Playa include pit ovens, artificial mounds, and

burials. Two small knolls with trincheras are present in the Cerros Boquillas. The pit ovens are basin-shaped, straight-sided, or undercut, about 1.0 m. in diameter, and occasionally lined with clay. They were heated either by building a fire in the pit, in which case the sides of the pit are firehardened, or by dropping in rocks which had been heated in a separate fire. Pit ovens of this type were widely used, in the Southwestern United States and in northern Mexico, to cook mescal or other plants of the agave type (Beals, 1932, pp. 164-65). The 16 artificial mounds are either oval or circular in outline, and vary in diameter from 5 to 40 m. The average maximum height is about 1.5 m. All the mounds are covered with a mantle of fire-cracked rock, resulting, excavations show, from the erosion of numerous pit ovens. In one mound the excavator found a roughly flat, adobe floor, from which ovens extended into the fill. It is assumed, therefore, that the mounds were artificial bases for pit-oven cooking, perhaps necessitated by the moist conditions of the cienega. Both inhumation and cremation burials occur. Inhumations are either flexed or extended, offerings are few and simple, and the bones are occasionally covered with red ochre. Cremations are of the primary type, where the body is burned in a pit and all bones and offerings are left in place. Here, too, offerings are seldom found. Indigenous pottery in the Trincheras area includes Trincheras Purple-on-red, Trincheras Polychrome, Trincheras Red ware, and an unnamed brown ware. All have a brown ware paste which has been modified in various ways. The most common painted type is Trincheras Purple-on-red (fig. 4), where the brown ware paste is often covered with a red slip on which simple geometric designs have been applied in a purple paint manufactured from specular iron. Jars, seed jars, and bowls occur. A deep scoring, probably accomplished with a bundle of grass 33

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

FIG. 4—TRINCHERAS PURPLE-ON-RED JAR AND SEED JAR. From Arizona State Museum collections. (Courtesy, Arizona State Museum.)

stems, appears on the interiors of jars and seed jars but shows no regular orientation. Trincheras Polychrome differs only in that the red slip color is used as a counterpoint addition to the purple design and is applied directly to the brown ware paste. Trincheras Red ware is known from intrusive sherds at the Paloparado site near Nogales (DiPeso, 1956, p. 362). It resembles the slipped Trincheras Purple-on-red vessels but lacks painted designs. The unnamed brown ware, undoubtedly utilitarian, constitutes the bulk of the pottery in the Trincheras region. Although descriptions of Trincheras culture artifacts are available (A. E. Johnson, 1960), one aspect should be mentioned here: the importance of shell in the manufacture of ornaments. Shell occurs, at least in some of the Trincheras sites (ibid., pp. 16689), in nearly as great quantity as in Hohokam sites of southern Arizona (Gladwin et al, 1937, pp. 135-53). The Hohokam have long been noted for their extensive use of this material. The predominance of shell in Trincheras sites, with La Playa perhaps the best example, has led Woodward (1936) to postulate that the Trincheras people had an important place in the prehistoric shell trade of the Southwest, and that La Playa, at least, was a manufacturing center for many artifacts that were traded 34

north into other parts of the Southwest. Brand (1938, p. 8) also reached this conclusion. The temporal placement of Trincheras culture has not been finally established, but it was in existence between A.D. 800 and 1100 (A. E. Johnson, 1963, pp. 182-83), on the basis of Trincheras Purple-on-red sherds discovered in sites of southern Arizona which are assignable to the Santa Cruz and Sacaton phases of the Hohokam sequence. These phases are cross-dated by intrusive pottery from northern Arizona. Dates for the northern Arizona types are based on tree-ring associations (Gladwin et al, 1937, pp. 212-20). Trincheras culture possibly lasted somewhat later, as sherds of Chihuahua polychrome have been collected from some Trincheras hillside sites (Sauer and Brand, 1931, fig. 2). The closest ties of Trincheras culture are with the Southwest rather than with Mesoamerica. None of the elaborate traits from regions to the south are present in northern Sonora, but there are numerous similarities with the cultural remains of southern Arizona, mostly (in trincheras sites and in artifacts of pottery, stone, bone, and shell) with the Desert Hohokam culture of southwest Arizona (A. E. Johnson, 1960, pp. 195219). Because of these close parallels, the Trincheras culture is believed to be a more southerly manifestation of the Desert Hohokam culture. NORTHEAST SONORA

Although the archaeological remains in northeast Sonora share many features with surrounding regions, a blend of a number of these features produced a somewhat distinctive complex and so merits a separate description. In a survey of the headwaters of the Santa Cruz River, particularly the San Rafael valley, Danson distinguished at least five types of sites: temporary camps, small villages marked by stone house rings, sherd areas perhaps representing somewhat larger

ARCHAEOLOGY: SONORA

villages, compounds, and caves. Some indications of Cochise-like material were discovered, but most of the ruins have pottery and date to between A.D. 500 and 1400. Cultural affiliations are with the Hohokam, Salado, and Trincheras complexes (Danson, 1941, pp. 51-53). Additional data on this northeast section come from Arroyo el Concho, a tributary of the Bavispe River (Lister, 1958, pp. 4 1 57). A number of caves along this arroyo contain cliff dwellings similar to those of Cave Valley, Chihuahua (ibid., pp. 8-40), 35 miles to the east. In the four caves investigated the cliff dwellings are small pueblo-like structures of rock, wood, and adobe. Many have T-shaped doors. One of the caves (ibid., pl. 12) contains the circular base of a granary of the type known in Olla Cave in Chihuahua (ibid., pp. 29-38). Pottery from Mogollon 4 or early Mogollon 5 (A.D. 900-1100) was found inside the cliff dwellings; pottery from below the dwellings was assignable to late Mogollon 3 (A.D. 900) (ibid., pp. 118-19). In their surveys throughout this region, both Bandelier (1892) and Lumholtz (1902) recorded numerous ruins but not in sufficient detail for use in interpretation of the archaeology. Amsden also surveyed parts of this area and located a number of sites. Among his inferences based on accumulating knowledge in surrounding regions was the existence of a peripheral development of Casas Grandes culture in northeastern Sonora (Μ. Amsden, 1928, pp. 44-46). Brand (1935, fig. 1) also noted this extension of Casas Grandes. Several cultural influences have come into contact in this region. Danson (1941, pp. 51-53) noted the presence of Hohokam, Salado, and Trincheras traits. Lister (1958, pp. 41-57) assigns the cliff dwellings and immediately underlying deposits to the Mogollon culture. Amsden (1928, pp. 4446) and Brand (1935, fig. 1) suggest Casas Grandes influence. Although the interplay of these diverse cultural elements has not

yet been worked out, apparently one result was a blend. An interesting but poorly understood pottery development here is perhaps a reflection of the cultural blending. The best known types are Santa Cruz Polychrome and Babocomari Polychrome. Babocomari Polychrome (DiPeso, 1951, pp. 123-29) has a brown ware paste covered with a white or cream slip. Geometric designs are painted in red and black. Both bowl and jar forms occur. Mica flakes are common surface features. Santa Cruz Polychrome (DiPeso, 1956, p. 333) differs only in its lack of mica and in the presence of a crackled slip. Both types may have developed out of Canelo Brown-on-yellow (Danson, 1941, p. 38), which possibly represents a regional variety of Trincheras Purple-on-red, as a result of influence from the great center of polychrome manufacture in the Casas Grandes region of Chihuahua (Sayles, 1936a). This blend of pottery traits suggests acculturation between the people of Casas Grandes and those of the Trincheras region. The significance of Hohokam, Salado, and Mogollon traits here remains to be determined. SOUTHERN SONORA

Somewhat less information is available for southern Sonora than for northern. Much of what is known comes from a survey by Ekholm (1939, 1940, 1942), who located sites in the vicinity of Huatabampo, near a former channel of the Rio Mayo. As there were apparently no surface features at these sites, interpretations have to be based on collections of artifacts, which included red ware, plain ware, and incised decorated pottery. Also present were ceramic figurines, which bore similarities to those found at Snaketown in southern Arizona in pre-A.D. 900 contexts (Ekholm, 1939, p. 10), slab metates, and overhanging manos (Ekholm, 1940, p. 326). Certain of these traits, including the red ware and the mano and metate forms, have 35

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

parallels at Guasave in northern Sinaloa (Ekholm, 1942). Most of the work at Guasave was done in a burial mound, where deep in the center one burial contained red ware vessels similar to those from the Huatabampo sites (Ekholm, 1939, pp. 910). None of the vessels from this burial resemble those from higher levels in the mound, which implies a temporal difference. On the basis of this stratigraphic evidence and other indications Ekholm (1942, p. 77) suggested the presence in southern Sonora and northern Sinaloa of a Huatabampo complex, which preceded the extension of Mesoamerican culture into this region. On a horizon generally after A.D. 1100, similarities have been noted between the Huatabampo complex and the cultural manifestations of southern Arizona (Ekholm, 1940, p. 326). Among these similarities are extremely close resemblances between the Huatabampo and Guasave red wares and such Southwestern types as Gila Red, Salt Red (Haury, 1945a, pp. 80-100), and Valshni Red (Haury, 1950, p. 17). Another Huatabampo trait which appears in late sites of southern Arizona is the overhanging mano (ibid.). The presence of these traits in southern Arizona indicates, at least, extensive trade. Additional data on the archaeology of southern Sonora come from the vicinity of Yecora, in the mountainous country near the Chihuahua border, where caves have produced mummified human remains. One of these mummies is on display at the University Museum at the University of Sonora in Hermosillo. Cultural associations remain to be worked out, but the presence of artifacts, some perishable, in the caves promises significant information. Also known in the Yecora area is La Cueva Pintada, whose walls are decorated with zoomorphic and geometric paintings in red and black (Wasley and Officer, 1959).

36

SUMMARY

The earliest archaeological remains from Sonora are referable to the prepottery Llano and Cochise complexes. The Llano complex is represented by two Clovis points from the Sonoran coast, the Cochise complex by remains from scattered locations on the coast and in northern and southern parts of the state. Artifacts from most of the Cochise sites, which have been reported, are similar to San Pedro stage Cochise remains from southern Arizona, although there is a suggestion of earlier Cochise remains in southern Sonora. Later remains from coastal Sonora are known from shell middens, which extend from the Punta Peñasco region to the southern boundary and also into Sinaloa. Cultural provinces along this coastal region remain to be delineated, but there are suggestions of differences between the areas around Punta Peñasco, Bahia Kino, and the Rio Mayo. The association of some shell middens with former shorelines offers interesting possibilities for stratigraphic work in the coastal region. Northwestern Sonora is the area of the Trincheras culture, perhaps the best-known archaeological complex within the state. This culture is characterized by terraced hillside sites and Trincheras Purple-on-red pottery. The complex dates between A.D. 800 and 1100, and is related to the Desert Hohokam culture of southern Arizona. Northeastern Sonora includes traits from a number of archaeological complexes including Trincheras, Casas Grandes, Hohokam, Mogollon, and Salado. Apparently, one result of the contiguity of these complexes was a blend, which is manifested in such pottery types as Babocomari Polychrome, Santa Cruz Polychrome, and Nogales Polychrome. The pottery complexes in this area date between A.D. 500 and 1450.

ARCHAEOLOGY: SONORA

Southern Sonora, the least-known part of the state, includes the Huatabampo complex, whose most distinctive trait is a red ware resembling redwares in northern Sinaloa and others in southern Arizona. No dates have been assigned to the Huatabampo complex, but it is probably late, as the similarities with Southwestern pottery are with types which occur after A.D. 1100.

The archaeological complexes within the state have their closest similarities with those of the Southwest, and not with the more elaborate developments of Mesoamerica. The northernmost extension of the Mesoamerican development apparently stopped at the Rio Fuerte in Sinaloa, and did not enter Sonora (see also Noguera, 1958, p. 26).

REFERENCES Amsden, M., 1928 Aveleyra, 1962 Bandelier, 1890b, 1892 Beals, 1932, 1942 Brand, 1938 Danson, 1941 Dice, 1943 DiPeso, 1951, 1955, 1956 Dockstader, 1961 Ekholm, 1939, 1940, 1942 Ezell, 1954 Fay, 1953, 1955, 1956a, 1956b, 1957a, 1957b, 1958, 1959 Gifford, 1946 Gladwin and Gladwin, 1929 Gladwin, Haury, Sayles, and Gladwin, 1937 Haury, 1945a, 1945b, 1950, 1960, 1962 , Sayles, and Wasley, 1959 Hayden, 1956 Hinton, 1955 Holzkamper, 1956

Ives, 1963a, 1963b Jennings, 1957 Johnson, A. E., 1960, 1963 kelley, 1956 Lehmer, 1949 Lister, 1958 Lumholtz, 1902, 1912 Manje, 1954 Noguera, 1958 Owen, 1956 Pfefferkorn, 1949 Roberts, 1945 Saner and Brand, 1931 Sayles, 1936a and Antevs, 1941 Schroeder, 1952 Sellards, 1952 Spicer, 1962 Wasley and Officer, 1959 Woodbury, R. B., 1961 Woodward, 1936

37

3. Archaeology and Ethnohistory of Lower California

WILLIAM

T

HE PENINSULA of Baja California is one of the most isolated regions of the New World, a fact which obtrudes in any aspect of the biotic landscape, including that of man. Part of an arid desert region, it is itself a desert, with all the inherent difficulties of movement, propagation, and survival. These restrictions guard its plant and animal inhabitants from outside disturbance. Thus in the realm of human activities old life ways characteristic of other areas to the north in earlier times have been preserved past their point of disappearance elsewhere, chiefly because these social and technological systems were not subjected to competition. Cultural change was a slow process. The importance of Baja California to Mesoamerica lies in the fact that, by its isolation, it has preserved, even into historic times, materials which were undoubtedly prevalent elsewhere on the mainland in early times. Since its only mainland connection is with southern California, the peninsula expectedly reflects northern prehistory. The most feasible comparative approach to the prehistory and ethnohistory of the peninsula lies through the archaeological and ethno-

38

C. MASSEY

graphic data directly to the north; although there are cultural data from the peninsula which can not be understood in these terms, many of the details and the over-all picture do have this context. Geographical and Geological Background The peninsula of Baja California joins with the mainland of North America immediately south of the international border (fig. 1), along which line the peninsula is about 140 miles wide. From here it trends south-eastward for 800 miles, paralleling the mainland Mexican coast. It is separated from the Sonora-Sinaloa coast by the islandstudded Gulf of California, which extends 600 miles from the mouth of the Colorado River to a point opposite Cabo San Lucas. The gulf itself averages 100 miles wide, although Baja California and Sonora are within 40 miles of each other at latitude approximately 29 degrees N., where Islas Tiburon and Angel de La Guarda shrink the distance even more. The peninsula tapers southward, having a coastline marked by numerous bays, lagoons, and inlets. The largest bay is Bahia de Sebastian Vizcaino, a bight on the Pacific

FIG. 1—BAJA CALIFORNIA, MEXICO

FIG. 2—BAJA CALIFORNIA TERRAIN, α, Volcanic cones, Las Tres Virgenes, northwest of Santa Rosalia, b, Mature granitic landscape, c, Western edge of the Magdalena Plains, d, Eroded granite country near dry Lake Chapala, showing Washingtonia palms, cirio, and cardon cactus.

coast where the land juts westward to Punta Santa Eugenia. Farther south on the Pacific is Bahia de La Magdalena, enclosed by its offshore islands. The gulf coast contains many small bays as well as the larger Bahia de La Paz and Bahia de La Concepcion. Although there are many offshore islands on both coasts, those in the gulf are more numerous. In the central gulf, Angel de La Guarda is one of a series of "steppingstones" across to Tiburon and the Sonora coast. Southward are San Marcos and Carmen. At Bahia de La Paz, three major islands enclose the bay: San Jose, Espíritu Santo, and Cerralvo. None of the gulf islands has running water, but fresh water occasionally occurs in tanks (tinajas) in protected places. On the Pacific side the major islands are Margarita and Magdalena, en40

closing Bahia de La Magdalena, and Isla Cedros northwest of Punta Santa Eugenia. The latter island is high and rocky, and has fresh water. Baja California is geologically complex because of its position on the unstable edge of the continent. It has experienced major faulting, vulcanism, and frequent submergence and emergence. The physiography is therefore mountainous, irregular, and broken. Basically the peninsula is a westwardtilted fault block, with steep slopes on the gulf side of the ranges as far south as the Isthmus of La Paz, beyond which lies the major high range of the Sierra de La Laguna, on the Pacific side of the cape region. In the northern area the Sierra San Pedro Martir rises to elevations of over 10,000

ARCHAEOLOGY & ETHNOHISTORY: LOWER CALIFORNIA

feet. These granites descend to a general plateau from the 30th to the 28th parallel in the central and gulf coastal areas, finally disappearing beneath younger sheet flows of lavas. But apparently the same granites and associated metamorphics reappear in the well-elevated Sierra de La Laguna, far to the south. Equally outstanding are the Tertiary and Quaternary sheet flows of lavas which lie in layers hundreds of feet thick in central and south-central Baja California. On the gulf coast the crest of the Sierra de La Giganta is 5,754 feet above the sea. Cenozoic vulcanism is visible as far north as San Quintín Bay on the Pacific coast of the peninsula and as far south as the southeastern edge of La Paz Bay. The Tres Vírgenes volcanoes which rise as spectacular cones above the level of the Comondu Formation to the west of Santa Rosalia, on the gulf coast (fig. 2,a), are reminders of recent activity. In post-Comondu time there were several periods when the sea encroached on the land. In the cape region are marine Pliocene deposits around the Sierra de La Trinidad north and east of San Jose del Cabo, at elevations of 1400 feet; they are also found on the inner edge of the Magdalena Plains and scattered intermittently south and east of Punta Santa Eugenia. The Pleistocene is marked in major degree on the peninsula by the alternating relations of land and sea. According to Beal (1948, pp. 118-19), there was an emergence of the peninsula followed by a "Great Submergence." During the latter phase the sea level may well have stood 1600 feet higher than at present. It may be that at this time the plants and animals were subjected to a crisis from which few vertebrate animals escaped. Subsequently, in Late Pleistocene times, there was a reemergence to the present conditions. Doubtless the Late Pleistocene fauna such as the large land mammals, bison, mammoth, and camel entered the peninsula under geologically modern conditions. The climate

was considerably more moist than at present; such Pleistocene terrestrial animals are found in situations indicating a wetter period (Massey, 1947, p. 352). 1 Convectional rainfall may occur in Baja California, especially in the south. The main patterns of rainfall are latitudinal, the northern frontier experiencing the southern extension of winter cyclonic cold fronts, and the cape region receiving the tropical and subtropical extension of late summer rains. The central portions of the peninsula may feel little or none of these effects during any year. Temperatures anywhere on the peninsula occasionally dip to freezing in winter. Summers are hot, although the cold California Current ameliorates conditions on the west coast. Humid warm conditions are oppressive on the gulf coast in summer. Surface water is scarce. Although arroyos are numerous, they usually run only briefly with water, which soon disappears into sand or evaporates. Only in the northwest are there a few streams of permanent water; to the south, in the central volcanic region, several large springs pour out, and mark centers of permanent habitation. These include: San Ignacio, Mulege, and Comondu, as well as San Bartolo and Todos Santos in the cape region. Elsewhere water can be obtained by digging shallow wells in arroyo bottoms. Flora and Fauna: Life Zones Although Baja California is chiefly a desert, its great extent in latitude, its varied surfaces and elevations, and its lengthy coast line contribute to diversity among its plants and animals. The fauna is clearly derived from the north beyond the peninsula, presumably in Late Pleistocene and Recent times. The flora is more varied in origin (fig. 3). Many plants, particularly those on the northwest coast, have their closest rela1

The geological summary is a synthesis of personal observations and the work of Beal (1948).

41

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

FIG. 3—LIFE ZONES OF BAJA CALIFORNIA. After E. W. Nelson.

tives in southern California, but the central and southern desert vegetation is derived mainly by way of the Colorado Desert corridor in the northeast. The cape region exhibits vegetation closely allied to that of the higher elevations of Sinaloa and southern Sonora across the gulf (Nelson, 1921, pp. 103-32). The Spanish discoverers and the missionaries of Baja California were struck by the variety and abundance of animal life, both on the peninsula and in the Pacific and gulf waters. Much of it served as food for the aboriginal populations. The large game animals included the mule deer (Odocoileus hemionis eremicus) still found the length of the peninsula, and the pronghorn antelope 42

(Antilocapra americana peninsularis) which formerly had a similar distribution. The mountain sheep (Ovis canadensis cremnobates) is narrowly restricted today, but in native times it ranged from the cape northward. Besides these are foxes, badgers, coyotes, mountain lions, cottontails, jack rabbits, a variety of mice (including the kangaroo mouse), and squirrels. Reptiles are represented by many genera. There are even a few amphibians on the peninsula. The coasts and seas of Baja California abound with a great diversity of mollusks, fish, and marine mammals. Perhaps the waters and littoral of the warmer gulf are more abundant; from an aboriginal economic viewpoint the greatest accumulations of shell heaps and debris are found on the gulf side. In part the vegetation indicates certain recent connections of the peninsula with the mainland. There are two life zones in northern Baja California: the Upper Sonoran on the northwest coast and the Transition in the higher elevations of the San Pedro Martir Mountains. The lowlands of this area contain the familiar flora of southern California; the higher elevations have good stands of Jeffrey, yellow, and lodgepole pines. The lower western margins of the mountains have various oaks. On the east side of the San Pedro Martir Mountains, from the mouth of the Colorado River south to the vicinity of San Ignacio, the peninsula lies within the Lower Sonoran Zone, extending down the higher elevations south of San Ignacio almost to La Paz Bay. The predominant vegetation is that of the deserts of the Southwest and Sonora. Many varieties of cactus abound. Along watercourses mesquite, paloverde, and, locally, palo blanco are common. Yucca, agave, copal, ironwood, cholla, ocotillo, and creosote bush are familiar plants of the American deserts. In the central deserts of Baja California, however, two plants unique to it give the peninsula a totally distinctive landscape: the tall cirio (Idria columnaris)

ARCHAEOLOGY & ETHNOHISTORY: LOWER CALIFORNIA

and the elephant tree (Pachycormus discolor). The watercourses of the northern stretches of the central desert have sporadic growths of fan palms. Farther south, mesquite, palo blanco, and paloverde grow in the courses of intermittent streams. In the southern cape region and occasionally north to San Ignacio are characteristic mimosas, cassias, and erythrinas of the Arid Tropical Zone. These are mingled with many of the plants of the Lower Sonoran Zone. In the cape region proper an abundance of vegetative types distinguishes the area: fan palms, wild figs, pitahayas, torotes, cardones, lomboy (Jatropha spp.), and mangroves in favorable coastal stretches. At higher elevations in the Sierra de La Laguna and elsewhere in the cape region there are isolated areas of the Upper Sonoran Zone. ARCHAEOLOGY

Northern Gateway to Baja California Most plant and animal forms entered the cul-de-sac of the peninsula through the northern gateway, which roughly coincides with the present international boundary. There is no reason to believe that man did otherwise. The cultural and physical antecedents of peninsular populations must be recorded in the archaeology of southern California and southwestern Arizona, but neither the prehistoric artifacts and skeletal remains nor the ethnographic traits and languages of Baja California would be expected to coincide precisely with those of the mainland. The geography suggests that the older material found archaeologically at the gateway would be represented widely southward on the peninsula. Late horizons in the north may find no representation on the peninsula, and therefore the archaeological and ethnographic materials of southern Baja California would be early in the north. The vertical stratigraphy in southern California sites would duplicate to some degree the horizontal "strata" of the peninsula of Baja California (Massey, 1961b). There

should be a time lag on the peninsula; artifacts typical of early horizons in the north may have occurred over long periods in the south—in some cases up to historic times. The peninsula is more simplified in prehistoric times than most areas in North America because it was subjected mainly to cultural and biological pressures and innovations from only one direction, or to internal development within local cultures. Along the southern California gateway there seems to have been a distinction between the prehistoric cultures of the Pacific coast and those east of the Peninsular Range in the desert portions of San Diego County. However, the San Dieguito culture is cited as a common denominator throughout the gateway area. The general distinction between the coast and the desert area lies in the differentiation, beginning with the La Jolla culture on the coast, between the marine-oriented coastal peoples and the hunting Amargosa (Pinto-Gypsum) peoples of the desert. Along the same line north of the peninsula, peoples placed in the Yuman sequence began practicing agriculture and making pottery from about A.D. 700 to the present (M. J. Rogers, 1958, p. 20). Everywhere in Baja California pioneer surveys and widely separated excavations were designed to solve particular problems. In the following examination of archaeological cultures on the peninsula, manifestations known to have specific connections with those in the northern gateway area will be considered first, followed by those for which we have no such connection. Our present confusion results from shifting terminology and lack of diagnostic implements or traits for many of the lithic complexes (Carter, 1957, pp. 174-80; Hunt, 1960, p. 64). San Dieguito Complex (Playa Complex) The earliest known complex in southern California which might have found its way into the peninsula is the San Dieguito I—111, tentatively dated at Ventana Cave to about 8000 B.C. (Haury, 1950a, p. 335), although 43

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

currently (M. J. Rogers, 1958, p. 20) from 2000 to 1000 B.C. This culture has little representation in any of the material in Baja California, so far largely confined to the northern border region (Treganza, 1942, 1947). The plano-convex scrapers, choppers, blades, and even the projectile points of San Dieguito III times may have gone unrecognized; however, artifacts, particularly of phase III, have been found in blowouts of the lowest strand lines on dry Lake Chapala (Massey, 1947, p. 354). South of Lake Chapala no sites or assemblies of these artifacts have been noted, but in the Castaldí Collection,2 projectile points of late San Dieguito (Playa II) types occur mainly in central and north-central Baja California (Massey, n.d.,a). The complex is poorly represented, and, to date, is limited to the north. Some artifacts and assemblages of a different nature, with attributable antiquity, may be considered here. Arnold (1957, pp. 250-53) has described the lithic tools of an assemblage characterized as "Elongate-Biface," heavily patinated, well-shaped elongate and discoidal scrapers and choppers. There are also large bifacial picklike percussion-flaked tools, on sites believed to antedate the other two assemblies of artifacts in the Chapala area. Similar implements are found at quarry sites on the south side of La Paz Bay (Massey, 1947, p. 348), but have little to suggest antiquity. Of interest is a single fluted point, without archaeological context, from near San Ignacio, most closely resembling the Clovis type (Aschmann, 1952a, pp. 262-63). None of these artifacts in Baja California 2 Padre César Castaldí, a modern Jesuit, made extensive archaeological collections from 1905, when his assignment on the peninsula began, until his death in 1946. The artifacts were derived primarily from ranches in the vicinity of Mulege, although material from Calmalli to Cabo San Lucas is included. The majority of the artifacts were identified by locality. In 1953 the collection was photographed, measured, and subsequently analyzed. The manuscript is now being illustrated for publication (Massey, n.d.,a).

44

have appeared in association with extinct animals, although bison, mammoth, horse, and camel bones have been found in localized Late Pleistocene formations under conditions suggesting the presence of man (Massey, 1947, p. 352). La Jolla Complex Following San Dieguito on the coast of southern California, contiguous to the peninsula, a twofold sequence has been suggested for a marine-oriented people named for La Jolla. Archaeological remains are few and generalized; the culture throughout is characterized by negative traits. There were burials in habitation midden sites, unaccompanied by artifacts, and in the late phase there were burial areas. The only typical artifacts seem to have been the shallow basin grinding stone and a one-hand cobble hand stone. In La Jolla II there were more chipped and flaked stone implements. The time span is believed to be from 1000 B.C. to A.D. 1450 on the San Diego County littoral, when the ancestors of the historic Diegueño occupied the area (M. J. Rogers, 1945, pp. 171-73). Rogers (ibid., p. 173) has noted coastal middens of this culture as far south as Ensenada; sites attributable to this culture reach at least as far south as Rosario. But enormous midden accumulations, particularly on the gulf coast, run the length of the peninsula, as well as occasional ones on the Pacific in favorable locations. None of them can be assigned to the La Jolla culture without further examination. Rogers has hypothesized (1945, p. 194) that historic groups (Pericú and others) in southern Baja California were La Jollans pushed into the area by Yumans after 1450, but archaeological, ethnographic, and linguistic data contradict this (Massey, 1947, p. 357; 1949, pp. 30305). Pinto Basin Complex (Amargosa I) The Pinto complex contains percussionflaked chopping tools, keeled scrapers, and

ARCHAEOLOGY & ETHNOHISTORY: LOWER CALIFORNIA

bifacially flaked leaf-shaped points and blades, as well as the Pinto Basin points which are considered diagnostic (fig. 4,a; Amsden, 1935, pp. 33-51). These heavy percussion-flaked points have five variations on a basic type (M. J. Rogers, 1939, p. 54, pl. 13) which is stemmed, with rounded shoulders and a concave base; the stem itself may be parallel-sided or expanding. Edge serrations are common. This type of projectile point, widely distributed in the western United States (Lister, 1953a, p. 265), is one of the best represented in Baja California. Lithic tools and projectile points belonging to the Pinto complex have been recognized at dry Lake Chapala (Massey, 1947, p. 354). The "scraper-plane assemblage" described for old levels of the lake basin apparently is part of this category, with the certain addition there of grinding stones (Arnold, 1957, pp. 253-61, pl. 18). No sites with this complex have been reported north of Lake Chapala on the peninsula, but typical Pinto Basin projectile points have been widely noted to the south. From an analysis of the Castaldi Collection and others, the type makes up 13 per cent of 1181 surface finds of projectile points between Calmalli in central Baja California and Cape San Lucas. It is most frequent on the Magdalena Plain and around the shores of La Paz Bay, and in significant numbers both north and south of these areas (Massey, n.d.,a; field notes). None of these points has been found in post-Spanish context. On the basis of the diagnostic points, this complex appears to be widespread and probably derived from the eastern desert region of southern California. Gypsum Cave Type Points (Amargosa II) Projectile points of Gypsum Cave type (Harrington, 1933, fig. 19) are frequent from San Ignacio south to the cape region (fig. 4,b). They make up about 5 per cent of surface collections of points in the area (Massey, n.d.,a; field notes). The main con-

FIG. 4—PROJECTILE POINTS, BAJA CALIFORNIA, a, Pinto Basin type. b, Gypsum Cave type.

centration seems to be on the gulf coast around Loreto and Mulege. As with the Pinto Basin points, with which they frequently occur, the Gypsum Cave type is not found on historic sites. The two types together often form a surface veneer on shell middens in the south. In addition to the true Gypsum Cave type, a series of tanged projectile points and blades occurs in central and south Baja Cali45

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

blades of tuff, were discovered when a road was cut through a deposit buried by 4 feet of alluvium (fig. 6; Massey, 1947, p. 350; 1955). Amargosa (Amargosa

III)

At present the Amargosa culture, as originally outlined by M. J. Rogers (1939, pp. 61-65), is represented only by a few projectile points in the Castaldí Collection. As these are derived from sites in the southern part of the peninsula, it seems that this culture had little influence in Baja California, and that the points may have another source. For the peninsula as a culture area, externally derived materials and sequences later than Gypsum (Amargosa II) have little meaning. The northern region, above the 30th parallel, shared in the late developments of southern California. Surveys of the northwest coast of Baja California, and tests at Cerrito Blanco and at Santa Catalina

FIG. 5—LA BLADES

PAZ

POINTS

AND

LORETO

fornia even more often. The points, designated La Paz points, are characterized by a slender triangular blade portion, sharp oblique shoulders, and a pointed tang. Occasionally the shoulders are rounded. The heavier group, known as Loreto blades, have a broad leaf-shaped blade portion, rounded oblique or rounded shoulders, and a broad convex tang. Both appear to have been locally derived from the Gypsum Cave type, and are more finely flaked with a good deal of pressure retouching, especially on the stem (fig. 5; Lorenzo, 1959; Massey, n.d.,a). Edges frequently are beveled. At one site (BC 82) in El Valle Perdido in the mountains of the cape region, both of these types, together with leaf-shaped 46

FIG. 6—POINTS AND BLADES, EL VALLE PERDIDO

ARCHAEOLOGY & ETHNOHISTORY: LOWER CALIFORNIA

FIG. 7—BURIALS, BAJA CALIFORNIA. a,b, Secondary burials, Piedra Gorda. c,d, Cerro Cuevoso. Note ceremonial object in d.

Mission, demonstrate the presence of Tizon Brown pottery at least as far south as Rosario Mission (Massey and Tuohy, n.d.,c; McKusick and Gilman, 1959). Potsherds are found occasionally south of the 30th parallel, but they are either late and near Spanish Mission sites, or near the 30th parallel and probably the result of trade or contact with California Yumans to the north (Massey, 1947, p. 355). The influence of ceramicbearing Yumans on the peninsula apparently stopped short of the 30th parallel. Yumanspeakers (Peninsula Yumans) to the south are derived from an earlier preceramic period. Two of these pre-Yuman I cultures are known in detail. Las Palmas Culture This culture spread throughout the mountainous cape region from the Isthmus of La Paz to Cape San Lucas. Burial caves are also recorded from the larger offshore gulf

islands (Cerralvo, Espíritu Santo, San Jose) in the same latitude. Funerary customs of the Las Palmas culture were known by the turn of the century (ten Kate, 1884; Belding, 1885; Diguet, 1899, 1905; Rivet, 1909). The human remains and associated artifacts were uniformly—and erroneously—attributed to the extinct Pericú Indians. Chief interest in these finds centered around the morphology and indices of the crania. The frame of the hyper-dolichocranic high-vaulted skulls long obscured the important archaeological notes of the early investigators. Definition of the culture rests on the excavation of three key caves on or near Bahia de Las Palmas, and on the examination of many other sites, most of which had been vandalized. The three caves, all small and used exclusively for interment, are: Cerro Cuevoso Cave (BC 75), excavated in 1947 (Massey, 1947, pp. 348-49; 1955), 47

FIG. 8—LAS PALMAS MATERIAL CULTURE, BATA CALIFORNIA, a, Ground-stone basin, possibly a grinding stone, b, Awl. c, "Spatula." d, Oyster-shell ornaments, e, Oliveüa beads. f, Ceremonial objects ( ? ) . g, Dart-thrower, h, Lark's-head netting, i, Coiled basketry, either native or traded. j, Sewn palm-bark container.

FIG. 9—COMONDU MATERIAL CULTURE, BAJA CALIFORNIA, a, Ground-stone shallow basin grinding stone, b, Ground-stone pipe, c, Chipped-stone projectile points; height of first one, 1.5 cm. d, Bone awl. e, Haliotis ornament. f, Olivella beads, g, Cane whistle, h, Wooden pitahaya hook, i, Flint chopper. j, Bull roarer ( ? ) . k, Arrow. l, Coiled basket, m, Square-knot netting, n, Tie-twined matting, o, Leather sandal, p, Tump band, q, Human-hair cape, r, Woven sandal, s, Decorated object.

ARCHAEOLOGY & ETHNOHISTORY: LOWER CALIFORNIA

and two caves on Bahia de Las Palmas, the Punta Pescadero site (BC 111) and the Piedra Gorda site (BC 114; Massey, 1955). These caves were alike in containing burials of people of Las Palmas culture. Although each cave did contain at least one unique artifact type, most artifacts are duplicates. Burial practices of the people using these caves consisted chiefly of depositing bound bundles of the defleshed bones of the dead, which had been painted with red ochre (fig. 7). These secondary burials were usually covered with deer skins or palm fronds and accompanied by artifacts. Thirty burials were recovered. In each cave there was at least one primary burial; BC 111 and BC 114 contained two. Primary adult burials were tightly flexed (fig. 7,c); one infant was extended. The ochre-painted secondary bundle burials, unique in western North America as a practice, confirm the earlier descriptions of ten Kate, Belding, and Diguet. Diguet (1899, p. 43) also noted an extended adult burial on nearby Espíritu Santo Island. The only stone artifacts in association with Las Palmas burials were crude percussion-flaked core chopping tools, discovered in the deposit at Cerro Cuevoso (BC 75). This scarcity of stonework balks any attempt to establish connections with the lithic assemblages of La Paz, Loreto, Pinto, and Gypsum points and associated artifacts in many open sites in the cape region. Such crude choppers are found only on open coastal middens. The diagnostic artifacts of Las Palmas culture include a number of unusual items (fig. 8). Four round-shafted wooden dartthrowers with hooks carved above the distal end and single bark-loop grips were recovered with one burial at Cerro Cuevoso (Massey, 1961a). Netting, made exclusively with lark's-head knots, was found at all three caves. Fragments of similar netting were also found at a number of looted sites (Massey, 1955). Another ubiquitous and remarkable artifact type is an oval container

about 75 cm. long, made of sewn palm bark reinforced around the edge with a flexible branch. This is the only type of container mentioned for the historic tribes of the cape region, and provides one of the links with the archaeology. However, at Piedra Gorda (BC 114) two coiled baskets made on a single-rod foundation with simple uninterlocked stitches were found on the skull of Burial 1. These may well be trade pieces, although they differ in wall treatment from those known directly to the north on the peninsula (Massey, 1955). Another unusual but characteristic artifact of the Las Palmas caves is a lozengeshaped hardwood piece, either flat or plano-convex, with a shark-tooth inset in black resin at one end. These well-polished pieces, about 35 cm. long, may be ceremonial or magical artifacts. Nothing comparable has been found on the peninsula or to the north. Olivella shell beads with ground spires were frequent. Circular oyster-shell (Pinctada mazatlanica) ornaments with serrated edges were burial accompaniments. Lozenge-shaped ornaments of oyster shell carved to represent fish were found with an infant burial at Punta Pescadero (BC 111). Spatulate bone pins, about 35 cm. long, and straight bone awls are also found. In summary, a number of unusual artifact traits in this funerary aspect of Las Palmas culture distinguish this culture from others in western North America. Since many of them are perishable and since the cape region is the end of the cul-de-sac, they may be traits which have since disappeared from the archaeological record of the mainland—as known thus far—to the north. The relationship of these burials to the open sites of the south is problematical. A single excavation of an open dune site at Los Frailes (BC 69) provides evidence of a definite relationship between Las Palmas culture and the coastal middens. The items in common are the circular oyster-shell 49

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

ornaments, the stone choppers, and the occasional shells in the cave deposits. Burials in the midden at Los Frailes were primary and, where ascertainable, flexed. The cultural relationship of the burial caves and rock shelters and their funerary culture, the middens, and the open sites of the interior with their lithic hunting artifacts (which alone have a meaningful distribution elsewhere on the peninsula) is unknown. Comondu

Culture

In central Baja California, from at least as far north as Bahia de Los Angeles to slightly south of Comondu, there seems to be a common late culture which is both prehistoric and historic (Massey and Tuohy, n.d.,a) (fig. 9). The Comondu culture is known from two excavations in the Sierra de La Giganta Mountains between Comondu and Loreto; Metate Cave (BC 100) and Caguama Cave (BC 57) (fig. 1). Identical and diagnostic artifacts have been recovered from other caves and noted in the Castaldi Collection from the Mulege-Loreto area (Massey, n.d.,a). In the north, at Bahia de Los Angeles, the same materials have been identified (Massey and Osborne, 1961), and are doubtless to be found in the intervening area. Since artifacts of this culture occur together with those of the Spanish Mission period in the two caves in the Sierra de La Giganta, they can be identified with the historic Laymon Indians; by extension Comondu was the widespread culture of the historic Peninsula Yumans of the central peninsula. Shallow basin and flat milling stones (fig. 9,a) were found in quantities, along with occasional single-hand cobble manos. Seeds ground on these milling stones had been parched in shallow basketry trays made by close-coiling on a bundle foundation, a type known prehistorically and historically in southern California (Tuohy and Massey, n.d.). 50

FIG. 10—PETROGLYPHS, BAJA CALIFORNIA. a, San Borjito (after Dahlgren and Romero, 1951). b, Cape Region (after Diguet, 1899).

Also within the economic complex in the Giganta caves are tiny triangular obsidian and tuff arrowpoints, frequently attached to hardwood foreshafts and cane shafts. Usually the points have serrated edges (fig. 9,c, left). They are totally different from the larger Pinto-Gypsum points found on many open sites. A gathering aspect of the culture can be inferred from "pitahaya hooks" (fig. 9,h), which are a peninsular counterpart of the

ARCHAEOLOGY & ETHNOHISTORY: LOWER CALIFORNIA

"saguaro hooks" common in the desert Southwest (Driver and Massey, 1957, p. 214, Map 33). Made by lashing pointed hardwood pieces at an acute angle to cane shafts, such hooks were used for gathering the fruit of the pitahaya and other cactus. Complete carrying nets and hairnets, as well as fragments, have been found at Bahia de Los Angeles, in the Giganta caves, and in the Castaldí Collection, all made by the unique square-knot netting technique (fig. 9,m). This is distinct from the lark'shead netting of the Las Palmas culture of the south and from netting reported elsewhere in western North America (Massey and Osborne, 1961). In addition, numerous fragments of cordage, woven tump-bands (fig. 9,p), and tubular stone pipes (fig. 9,b) were common to all sites bearing the Comondu culture. The upper levels of the caves in the Sierra de La Giganta contained mission-period maize (Nickerson, 1953, p. 86) and Lagenaria fragments (Whitaker, 1957), as well as bits of porcelain, pottery, and iron spikes. In the burial cave at Los Angeles Bay, Edward Palmer found several Comondu culture traits, including the square-knot netting, two kinds of tubular stone pipes, and a woven tump-band, all identical to the Giganta cave specimens. Besides, a series of artifacts accompanying the burials is reported historically for the entire Peninsula Yuman area. They include the human-hair capes or masks and cane whistles, usually associated with shamans. Burials at Los Angeles Bay were extended (Massey and Osborne, 1961). In summary, the currently known archaeology of Baja California contains several distinct components. The Pinto-Gypsum (Amargosa I and II) is generally widespread from about the 30th parallel south to the cape. The Las Palmas culture is known only in the cape region and adjacent gulf islands. It has suspected, but undemonstrated, affiliations with the Pinto-Gypsum complex, and has definite associations with

the coastal midden cultures. Throughout central Baja California the Comondu culture is known to be that of the historic Indians and to have prehistoric phases as well. It apparently is unrelated to the PintoGypsum culture, which also occurs in the same area. Petroglyphs Petroglyphs occur the length of the peninsula, painted forms much more commonly than pecked. Zones in Baja California can be clearly delimited on the basis of distribution and occurrence of design elements, many of which are definitely localized. Some abstract elements and styles are ubiquitous, such as zigzags, parallel lines, circles connected by a line, chevrons connected by a line (Massey and Tuohy, n.d.,6). As usual the cape region is unique. There isolated rocks are embellished with realistic, solidly painted animals, mostly fish, deer, and rabbits. Solid drawings of hands with missing fingers are known only from the south (fig. 10,b; Massey, field notes). The most unusual paintings on the peninsula are restricted to the Mulege-Comondu area, where groups of humans with headdresses, body paintings, and upraised arms are depicted in caves and rock shelters. One of the most spectacular multicolored groups in North American cave art is the series of humans, many pierced by arrows or spears, painted in the cave of San Borjita west of Mulege (fig. 10,α). Typologically these most closely resemble portrayals of humans at Basketmaker sites and in Val Verde County, Texas (Diguet, 1899; Dahlgren and Romero, 1951, pp. 171-78). ETHNOHISTORY

Language

Distributions

Two linguistic families are represented on the peninsula: the Yuman and the Guaicurian (Massey, 1949; Aschmann, 1959, pp. 52-57). The latter was spoken in the south51

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

FIG. 11—LINGUISTIC RELATIONSHIPS IN BAJA CALIFORNIA. (After Massey, 1947.)

ern third of Baja California, and its speakers were shut off from the rest of North America by the Yuman-speaking peoples (fig. 11). The rancherías or lineages comprising the Guaicurian-speaking group occupied the Magdalena Plains and the Isthmus of La Paz as the Guaicura and Callejue. In the cape region two distinct languages were spoken: Huchiti and Pericú. The best known of the Huchiti-speaking groups was the Cora, which held the gulf coast from La Paz Bay to the southern end of Las Palmas Bay, the area where most of the so-called Pericú materials have been found. In the extreme south, embracing the area 52

around Cape San Lucas and the gulf islands north to Santa Catalina, were the Pericú. North of the Guaicurian area, and well beyond the present international border, were groups speaking Yuman languages. These were linked in distribution with peoples of Yuman speech in southern California and Arizona. On the peninsula proper there was a major division between the California Yumans (Diegueño, Kiliwa, Paipai [Akwa'ala], Kamia, and Ñakipa) and the Peninsula Yumans ("Cochimí"), which included the following languages: Borjeño, Ignacieño, Cadegomeño, Laymon, MonquiDidiu (Cochimí). Differentiation was marked between the Yuman languages of southern California and northern Baja California, and the Peninsula Yumans, but slight among the dialects or languages within the peninsula grouping. Missionaries in the 18th century had little language difficulty from Loreto northward to about the 30th parallel, but beyond that neither they nor their interpreters could understand the Indians of the California group (Massey, 1949, pp. 293-95). Distinctions among languages in the Guaicurian family appear to have been even more pronounced, although the data are less precise than for the Yumans of the north. Three markedly different languages were Guaicura proper, Huchiti, and Pericú. The linguistic distributions in Baja California demonstrate the familiar stratification noted by Kirchhoff (1942). The Guaicurian languages were hemmed in to the south by the Peninsula Yumans, who themselves were cut off from the north by the California Yumans. This is not to be thought of as caused by physical pressure by the Peninsula Yumans on the Guaicurians but quite the reverse; the Pericú and other southern tribesmen were much more aggressive and held a more attractive and ecologically superior area for the general peninsula hunting-and-gathering economy. During the Spanish period they frequently attacked

ARCHAEOLOGY & ETHNOHISTORY: LOWER CALIFORNIA

the nearby Yumans and even other Guaicurians (Massey, 1949, p. 281). The Guaicurians must represent an early population in Baja California which was shut in to the south by at least two migrations of Yuman-speakers, both of whom differed markedly from one another. The extra-peninsular linguistic affiliation of the Guaicurian family is unknown. Ethnography in Baja California For the field ethnologist the peninsula presents a sterile area. At present only a few score Indians live in the mountains of the north, near the border, of either the Paipai (Akwa'ala) or Kiliwa groups (Gifford and Lowie, 1928; Meigs, 1939). In the early 17th century, prior to extensive contact with the Spanish and their missions, the Indians of the peninsula numbered in the thousands. Careful scrutiny of the Spanish documents leads to the conclusion that the aboriginal peninsula held 47,965 Indians (more or less) in 55,000 to 60,000 square miles (Gerhard, n.d.; Aschmann, 1959, Table 7, p. 178; Cook, 1937; Meigs, 1935, p. 140). Kroeber (1939, pp. 178-79), on the contrary, believed that any figure approaching this is inconceivable. Baja California is without parallel on the continent for ethnographic reconstruction. From first Spanish contact with the natives in 1534 to the end of the 18th century documentation of the native cultures is unbroken. Most of the relevant material was written by the missionaries. One of the most famous ethnographic descriptions of all times is Father Baegert's account of the Guaicura Indians at Mission San Luis Gonzaga in his Nachrichten von der amerikanischen Halbinsel Californien of 1772. Two other major compilations relate to the peninsula: the Storia della California (1789) by Father Francisco Clavigero, and the Noticia de La California (1757) of which the documents were collected by Father Miguel Venegas and edited by Andres Burriel. These two books have well-organ-

ized ethnographic sections based on contemporary descriptions, as well as hundreds of documents which provide precise data by region and locality. Basic Cultural Adjustments in Baja California A great deal of basic similarity existed among the cultural groups in this area. Aboriginal subsistence was gained by hunting, fishing, and gathering throughout the peninsula. Only the Cocopa, as a River Yuman tribe—but technically within Baja California —practiced agriculture. Peninsular peoples gave varying importance to fishing and hunting and showed marked differences in skills and equipment. In hunting and warfare the bow and arrow were universal. These seem to have been longer in the Guaicurian area, with bow lengths of 180-210 cm. reported, and arrows of 137 cm. The arrows were either foreshafted with a small stone point or bone; unpointed arrows were known in the south (W. Rogers, 1928, p. 231; Shelvocke, 1814, p. 7; Baegert, 1942, p. 86; Ulloa, 1925, pp. 38, 40). Northward in the Yumanspeaking area bows were shorter (Clavigero, 1937, p. 99; Meigs, 1939, p. 30). Virtually universal on the peninsula was the use of the tule balsa and the doublebladed paddle (Heizer and Massey, 1953, p. 290). The distribution is continuous to southern California and extends to the Seri of the Sonoran coast. Throughout the peninsula men customarily were nude; women wore aprons, usually two-piece with the rear half of deerskin and the front of cut cane (carrizo) tubes strung on cordage (Massey, n.d.,b). Among the Pericú of the cape region women wore a skirt of bark and grass (W. Rogers, 1928, p. 229; Venegas and Burriel, 1943, 1:79; Shelvocke, 1814, p. 5; Clavigero, 1937, p. 96). Skin capes were also universal (Massey, n.d.,b). Other items of peninsula-wide distribution include: carrying nets, skin or animal53

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

bladder containers for water (except for the northern pottery-making area), fire-making equipment for fire-drilling, and grinding stones and hand stones (Massey, n.d.,b). Shamans held a place among bands throughout the peninsula; they were of both sexes, although men predominated. The most common curing method was to suck out disease-causing objects through stone or reed tubes (Massey, n.d.,b). The socio-political unit was the band, or patrilineage in Baja California. The band or ranchería was composed of several families probably related through descent (Massey, n.d.,b; Aschmann, 1959, p. 122). Each band exploited a definite territory and the biological family was the economic unit of exploitation. Such rancherías were small and contained 50-200 persons. Political leadership was minimal and occasional (Massey, n.d.,b). Culture Restricted to the Yuman Area In the northern two-thirds of Baja California, the area held by Yuman-speakers, a number of specific traits and complexes were not part of the cultural inventory of the southern Guaicurian peoples. Several of these items are of economic importance, but most have religious significance. This is equally true when the Peninsula Yumans are differentiated from the California Yumans. Among the economic traits are: curved "rabbit sticks" (Clavigero, 1937, p. 100; Simpson, 1938, p. 19; Sales, 1794, 1:48; Krmpotic, 1923, p. 126), hook-and-line fishing with cactus-thorn and/or tortoiseshell hooks (Wagner, 1925, pp. 25, 52; Bolton, 1916, p. 73; Wagner, 1924, p. 339; Sales, 1794, 1:48), nets used for fishing (Bolton, 1916, p. 59; Venegas and Burriel, 1943, 1:81). Coiled basketry trays were another common feature for gathering small seed as well as for toasting them with coals (Aschmann, 1952b; 1959, pp. 62-63; Massey, n.d.,b). They were made by the technique of close54

coiling on a bundle foundation (Tuohy and Massey, n.d.). Skin blankets made by tying strips of rabbit, deer, fox, seal, and other skins have a similar distribution, although they were rare in Baja California (Aschmann, 1959, p. 107; Massey, n.d.,b). A number of religious and ceremonial traits and complexes had a pan-Yuman distribution on the peninsula. These include the mourning anniversary, the use of human-hair capes, and the use of tubular stone and/or pottery pipes by shamans. Sacred carved wooden tablets and idols were apparently known to all Yumans. Periodic mourning ceremonies and anniversary mourning of the dead are typically Californian (Kroeber, 1925, pp. 859-61) and are reported from southern Peninsula Yuman territory (Clavigero, 1937, p. 111). A similar ceremony was held for an individual recently deceased among the Diegueño and Kiliwa of the 18th century and recent times (Sales, 1794, 1:76-78; Meigs, 1939, pp. 5059). Common to the area were first-fruits ceremonies (Bolton, 1936, pp. 175-77; Sales, 1794, 1:116-21; Clavigero, 1937, pp. 11011). This links with similar practices in southern California (Driver and Massey, 1957, Map 65). Human-hair capes and wiglike masks were an invariable part of the shaman's paraphernalia (Meigs, 1939, pp. 50-59; Aschmann, 1959, pp. 113-15; Sales, 1794, 1:76-80; Bolton, 1936, p. 175); they have also been recovered archaeologically (Massey and Osborne, 1961, pp. 349-50). The capes were reported by Baegert (1942, p. 123), possibly for the Guaicura, but by no other authority for the area south of the Yuman-speakers. Tubular stone pipes (chacuacos) were especially associated with shamans who used them in curing both for smoking and for sucking out disease-causing objects (Venegas and Burriel, 1943, 1:93, 95; Clavigero, 1937, p. 115). Archaeologically they are restricted to the Yuman area on the peninsula (Massey, n.d.,a), and have been

ARCHAEOLOGY & ETHNOHISTORY: LOWER CALIFORNIA

recovered with other shamanistic paraphernalia at Bahia de Los Angeles (Massey and Osborne, 1961, pp. 341-42). Reed tubes had a similar function. Tubular clay pipes are reported sporadically north of San Ignacio (Ulloa, 1925, p. 60; Sales, 1794, 1:107). Throughout the Yuman-speaking area ranchería chiefs or shaman chiefs had anthropomorphic statues or idols carved of wood, or made piecemeal of twigs and branches and dressed with hair and clothing. They were used in first-fruits ceremonies in the south, and in the north on a variety of sacred occasions (Bolton, 1936, pp. 175-76; Ducrue, 1765, p. 91; Krmpotic, 1923, pp. 106-07; Linck, n.d., p. 213). Similar idols have been reported among the modern Kiliwa and Paipai (Meigs, 1939, pp. 54-55). Both in the documents and in the historic archaeology the traits link the Yumanspeaking peoples of the peninsula together culturally and exclude the southern tribes of Pericú and Guaicura. Cultural Distinction between California and Peninsula Yumans The distinction between the California and the Peninsula Yumans is more than linguistic, for certain cultural traits and complexes were not held in common. These are chiefly ceremonial or religious but also material. This distinction is a critical area of ethnohistorical interpretation. Kroeber has described the peninsula properly in terms of a quantitative cultural rise from south to north (1939, pp. 43-44). Kirchhoff, on the other hand, has stressed the cultural stratification from north to south (1942). Both are correct in their own terms. The answer lies in the details and reflects divergent histories (Massey, 1961b). The Diegueño and the Kiliwa, representing the California Yumans, possessed traits either entirely foreign to the Peninsula Yumans or partially known to those near the zone of contact along the 30th parallel. The major items include the stone mortar

(Aschmann, 1959, p. 64; Massey, n.d.,b); slings (Sales, 1794, 1:42); the dog (Massey, n.d.,b); pottery (Massey, 1947, p. 355); and the sweat house (Aschmann, 1959, pp. 109-10; Meigs, 1939, pp. 63-64). On the other hand, the Peninsula Yumans possessed traits and complexes, chiefly religious, not found among the California Yumans at all. The most unusual aspect of religious difference between the Peninsula and California Yumans lies in the presence of god impersonation, male secret societies, and tree decoration in central Baja California. Kroeber has already commented on the unusual distribution of god impersonation and male secret societies in the southwest, central California, and Baja California (1932, pp. 414-15). Tree decoration, which appeared in the description of the same firstfruits ceremony of the Peninsular Yumans by Admiral Atondo in 1684, is unique (Clavigero, 1937, pp. 110-11; Bolton, 1936, pp. 174-77). Interwoven with these traits are two other features of religious life which involved more than a single ranchería: food and deerskin distribution and foot-races. Both features occurred at interband festivals, which might incorporate other ceremonial elements (Aschmann, 1959, pp. 127-30; Massey, n.d.,b; Clavigero, 1937, pp. 110-11; Venegas and Burriel, 1943, 1:96; Bolton, 1936, p. 176). The custom of swallowing chunks of meat tied to cord and then pulled from the stomach and passed to another individual was peculiar to the Peninsula Yuman area. This custom (maroma) was frequently mentioned by the Spaniards, who, although they regarded it as a response to the scarcity of food, noted that it was a ceremonial practice in which the participants sat in a circle and passed the meat from person to person (Aschmann, 1959, pp. 94-95; Massey, n.d.,b). These cultural practices follow the demarcation between the California and the Peninsula Yumans indicated by the linguistic 55

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

FIG. 12—ARCHAEOLOGICAL AND LINGUISTIC CORRELATIONS IN BAJA CALIFORNIA

differences (Massey, 1961b). The two have divergent ethnohistories. Guaicurian Area The Guaicurian-speakers lacked all the traits described above but possessed some unique customs of their own. Spears with fire-hardened tips were used for fishing, hunting, and warfare (Picolo, 1919, 2:59; Ascensión, 1916, pp. 59, 113; Ulloa, 1925, pp. 38-40). Among the Pericú the dart-thrower with fire-hardened darts was still current during the early 17th century (Massey, 1961a). It is important to note that spears were the only known fishing implements; hooks-and-lines, and nets were not in use in historic times (Massey, n.d.,b). Another unique aspect of material culture was the lack of basketry in historic times in the cape region. Oval-shaped bark containers were made of "staves similar to those of barrels," about 45 cm. long and 4 or 5 fingers wide, sewed together (Clavigero, 1937, p. 98). They have been recovered 56

archaeologically in the Las Palmas materials. The famous second harvest of the pitahaya was not confined to the Guaicurian area, but it was especially typical of that region. This was the custom of gathering the desiccated seeds of the pitahaya from excrement and reducing them to flour by grinding (Massey, n.d.,b). The tribes of the cape region differed markedly from the Guaicura in being more warlike, aggressive, and better integrated politically. In these respects they stand apart from all peninsular groups (Massey, 1949, pp. 280-81, 304). So fierce and unrelenting was their warfare that they disrupted the establishment of missions between Loreto and La Paz for years. Eventually the Pericú Revolt erupted in 1734. For two years the groups of the cape threatened the Spanish holdings in Baja California. These attacks and massacres were without parallel elsewhere in either Upper or Baja California. Here, they can be

ARCHAEOLOGY & ETHNOHISTORY: LOWER CALIFORNIA

correlated with the more bountiful environment, greater native population, and closer political integration. SUMMARY OF THE CULTURE HISTORY

Although details of culture-history interpretation of the peninsula will change with future data, major outlines are recognizable (fig. 12). Ethnohistorical interpretation is facilitated by the position of Baja California as a cul-de-sac, chiefly accessible from the north to the relatively unsophisticated peoples who became distributed "stratigraphically." The historic stratification of language and culture is the most salient feature of the historic ethnography but would be equally valid on any aboriginal time level. Early peoples could have entered the peninsula from the north any time after the occupation of the northern gateway in southern California. The earliest hunting immigrants probably arrived under a wetter climatic regimen than the present, when the dry central desert is a considerable barrier for land-oriented people. These peoples could have come when the large Late Pleistocene and Early Recent mammals (mammoth, horse, bison, camel) were still present. Peoples with a marine orientation could have entered any time from southern California. In Baja California archaeological materials which can be identified with certainty begin with the Pinto and Gypsum Cave complexes. These are known on the peninsula from the Colorado Desert east of the San Pedro Martir mountains south into the cape region. The entrance time could have been about 7500 B.C. (Haury, 1950a, p. 335) with a subsequent time lag down the peninsula resulting in the possible survival

of the cultures into historic times in the far south. The record of any pre-Pinto culture on the peninsula is not yet definite but is possible. The Las Palmas culture of the cape region may have begun in its known form prior to the Pinto-Gypsum diffusion or migration, or it may be a later aspect of PintoGypsum in the area, or it may have been a more recent (but still pre-Yuman) invasion of Baja California which came to rest in the south. Its known representation is clearly that of the Guaicurian-speaking cape people (Huchiti-Pericu). The unusual artifact inventory of both the historic tribes and the Las Palmas culture (such as dart-throwers, absence of basketry, lack of fishing equipment), coupled with the famous dolichocranic physical type, forces the conclusion that this is a relic of an ancient past in western North America. Subsequent to the stand of the Guaicurians in the south but prior to Yuman I (A.D. 700) at the northern gateway, the bearers of the Comondu culture entered the peninsula. Since this culture is that of the Peninsula Yumans in historic times, presumably the ancestral people were of Yuman speech. Only the historic phase, however, should be so identified. The more definite separation of Peninsula Yuman languages at the southern edge of their distribution on the peninsula indicates a long time for the development of their differentiation in situ (Massey, 1949, p. 304). The Yuman-speakers did not delay their invasion of the peninsula until after A.D. 1450 (cf. M. J. Rogers, 1945, p. 194). It is very doubtful that even the California Yumans of the northern peninsula, who belong anthropologically with southern California, entered their area that late.

57

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

REFERENCES Amsden, C. Α., 1935 Arnold, 1957 Ascensión, 1916 Aschmann, 1952a, 1952b, 1959 Baegert, 1942 Beal, 1948 Belding, 1885 Bolton, 1916, 1919, 1936 Campbell and Campbell, 1935 Carter, 1957 Clavigero, 1937 Cook, 1937 Dahlgren and Romero, 1951 Diguet, 1899, 1905 Documentos . . . Mexico, 1853-57 Driver and Massey, 1957 Ducrue, 1765 Gerhard, n.d. Gifford and Lowie, 1928 Harrington, 1933 Haury, 1950a Heizer and Lemert, 1947 and Massey, 1953 Hunt, 1960 Kerr, 1814 Kirchhoff, 1942 Krmpotic, 1923

58

Kroeber, 1925, 1932, 1939 Linck, n.d. Lister, 1953a Lorenzo, 1959 McKusick and Gilman, 1959 Massey, n.d., a,b, 1947, 1949, 1955, 1961a, 1961b and Osborne, 1961 and Tuohy, n.d., a,b,c Meigs, 1935, 1939 Nelson, 1921 Nickerson, 1953 O'Crouley, n.d. Picolo, 1919 Rivet, 1909 Rogers, M. J., 1939, 1945, 1958 Rogers, W., 1928 Sales, 1794 Shelvocke, 1814 Simpson, 1938 ten Kate, 1884 Treganza, 1942, 1947 Tuohy and Massey, n.d. Ulloa, 1925 Venegas and Burriel, 1943 Wagner, 1924, 1925 Whitaker, 1957

4. Archaic Cultures Adjacent to the Northeastern Frontiers of Mesoamerica1

WALTER

EYOND the northeast frontiers of Mesoamerica lies a desert land. Its cultures, carried by nomadic hunters and gatherers variously called "barbarians," "wild tribes," Chichimecs, are centered on the north Mexican states of Coahuila, Nuevo Leon, and Tamaulipas, extending into adjacent parts of Mexico and Texas (fig. 1). Except for the pertinent parts of Texas, the area lies roughly south of the Rio Conchos (Chihuahua) and the Rio Grande, north of the city of San Luis Potosi, and between the Gulf of Mexico and the eastern skirts of the Sierra Madre Occidental. The southern part of Tamaulipas, south of the Soto La Marina River, was occupied during the later prehistoric periods by obviously Mesoamerican cultures, including the Huastec; therefore, in this region I shall discuss only the more ancient, non-Mesoamerican, or pre-Mesoamerican cultural manifestations. Physiographically, this vast area may be divided into three major provinces: the coastal plain of the Gulf of Mexico, the Sierra Madre Oriental and its outliers, and 1

Published by permission of the Secretary of the Smithsonian Institution.

W.

TAYLOR

the central Mexican plateau (Tatum, 1931; W. W. Porter, 1932). All these provinces extend across the Rio Grande into Texas, the adjacent portions of which, as has been said about the Trans-Pecos region specifically (W. P. Taylor et al, n. d., p. 2), thus "represent a northward extension of topographical forms and biological resources primarily Mexican in relationships and character." This condition also prevailed in cultural matters. The coastal plain, beginning with the extensive dunes which line the coastal estuaries, rises slowly from the Gulf to the front ranges of the Sierra Madre in Mexico and to the Balcones Escarpment in Texas. In Mexico, this smooth, monotonous topography is broken only by the isolated mountains of the Sierra de Tamaulipas-Sierra San Carlos and by the strike-ridge of the Ceja Madre in the vicinity of Nuevo Laredo, Tamaulipas. In Texas, there are no mountains at all, except localized irregularities hardly deserving the name. Other than the coastal rivers of Texas, only the Rio Grande and a few short watercourses in coastal Tamaulipas drain the area. Again except in Texas, where rivers constituted barriers, 59

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

drainage lines are short, shallow, and ephemeral. They are not salient features of the landscape, nor were they significant factors in the aboriginal culture ecology. The biota is Tamaulipan, characterized by a flora consisting largely of mesquite and acacia, below and between which grow thorny shrubs and succulents (Goldman, 1951, p. 425). The Sierra Madre province encompasses the eastern cordillera and its front ranges. Its southern extensions consist of a series of north-south ranges separated by narrow valleys difficult of access. In the north the ranges shift direction toward the northwest and separate to form broader valleys often occupied by playa basins. The life zones range from Lower Austral to Canadian, characterized by floral contexts of mesquiteacacia, then oak-pine, piñon-madroño, and finally fir-aspen (Goldman, 1951, p. 429). The drainage, when not interior, flows east onto the coastal plain and into the Rio Grande. The mountains, mostly of limestone, contain many caverns, caves, and shelters which were used by the aboriginal populations. The central Mexican plateau is an arid region of playa basins and abrupt, isolated mountain masses. It has a typical basin-andrange topography with some of the higher mountains reaching 9000-10,000 feet. It is included within the Chihuahua-Zacatecas biotic province (Goldman, 1951, p. 421) and mainly within the Lower Austral life zone, exhibiting a flora of mesquite, acacia, ocotillo, creosote bush, and many other desert shrubs and succulents. Some regions are Upper Austral and, in the higher elevations, reach the pines and oaks of the Transition zone. For the most part, the mountains are of limestone and contain many protected, habitable sites. The amount of culture-historical information from this immense area is small. The aboriginal population, sparse to begin with and not accustomed to political or religious domination or a sedentary way of life, did 60

not take kindly to mission reduction or the encomienda system. Decimated by disease and fighting, the Indians disappeared rapidly after the establishment of Spanish rule, and none endured long enough to provide ethnographic information in any detail or extent. Archival sources are scarce at best, and what data they do contain are often colored by religious and evangelical motives which reduced their coverage and throw suspicion upon their objectivity and truth (Martínez del Río, 1954; Alessio Robles, 1927, 1938; Beals, 1932). Ethnographic evidence being nil and archival materials few and faulty, the archaeological record is little better. Fieldwork has been sparse in a huge area where cultural differences appear to have been small (although perhaps no less significant therefore) and where much more detailed work than usual is required to delineate the meaningful and important differences. At present writing, only two professional archaeologists have made stratigraphic excavations in the area; surface surveys and collections from unstratified burial sites have been made by a few professionals and a handful of nonprofessionals; and there has been considerable looting by many people. More serious, publication has been neglected; with a few exceptions, reports are not complete enough for more than general purposes. I here refer primarily to the major (i.e., Mexican) sector of the area: more professional and usable work (as well as more looting) has been done in the related regions north of the Rio Grande. Even there, however, controlled excavations have not been numerous, and some of the resulting publications are not suitable for the refined cultural analysis required by the nature of the aboriginal remains and the cultures they represent. CULTURE AND CULTURE-SEQUENCE IN COAHUILA

This article will be developed from an archaeological base in the state of Coahuila,

ARCHAIC CULTURES: NORTHEASTERN FRONTIERS

FIG. 1—NORTHEASTERN MEXICO AND ADJACENT TEXAS (mountain masses representational).

Mexico, because (1) I know the material here better than elsewhere; (2) there is more material in stratigraphic placement than for any other sector of the northeastern frontier; (3) there seem to be cultural relationships between Coahuila and virtually all the other sectors of the frontier; and (4) the material has been studied in sufficient detail to provide information on what may be called "microvariations," through which small cultural and chronological differences may be defined and compared with some expectation of significant results. Because all variations and their meanings cannot be discussed here I shall have to make many statements for which documen-

tation is unpublished. A similar restriction affects the making of detailed concordances between the archaeological and the ethnohistorical data; therefore, the pertinent ethnohistorical sources are listed in the references, and the archaeological traits for which reasonably acceptable ethnohistorical counterparts exist are starred in the text. From the earliest times of which we have knowledge to the latest there was a cultural continuum in Coahuila, belonging to what has been called the Desert culture (Jennings, 1956, pp. 70-72; W. W. Taylor, 1956, pp. 129-220). Many classes and types of artifacts are the same from bottom to top in the stratigraphic sequence. Variations 61

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

occur and relative frequencies change, but without doubt it is a single cultural tradition throughout its approximately 10,000 years. Within this continuum, however, we can distinguish three complexes. These are not to be thought of as "cultures" or separable entities in any partitive, ethnic sense but merely chronologically separable parts of the total inventory. In brief there was a single "culture" which lasted from bottom to top in our stratified deposits: at the beginning of this range were certain artifact types which later disappeared, and, toward the end, a new series of types were incorporated into the total inventory. These changes, early and late, may have been induced by outside influences or have been endogenetic, but the important fact is that they occurred within a single cultural tradition. In addition to these three complexes found in stratigraphic context, three others have been recognized but have yielded very little comparative, and even less stratigraphic, information. Cienegas Complex The Cienegas complex is known from three sites. Two have been excavated stratigraphically—Frightful Cave (CM-68) and Fat Burro Cave (CM-24); the third site (CM-65) had been vandalized and produced only spoil-pile materials. All three are in canyons opening into the Cuatro Cienegas Basin in central Coahuila (fig. 1). The complex consists of a congeries of traits which formed a small part of the total cultural corpus characteristic of the earliest human occupation of which we have knowledge in Coahuila. Figure 2 illustrates examples of some types included in this complex: wads of human hair,* tucked beneath boulders and cobbles of fall-rock on the bottom of Frightful Cave below the earliest cultural deposits, hair cut at regular intervals of about one month; rattlesnake rattles, either cut or pulled off and the growing matrix carefully 62

FIG. 2—CIENEGAS COMPLEX ARTIFACTS. Left to right and top to bottom: wad of human hair, rattlesnake rattle, agave scuffer-sandal, fiber cross, narrow plaited band, shell of Humboldtiana montezuma Pilsbry, twill-pad sandal. Frightful Cave and CM-65. (Photo, Smithsonian Institution.)

removed; agave scuffer-sandals, some from the bottom level of Frightful Cave gave a radiocarbon date of 6125±450 B.C. (Crane and Griffin, 1958b, p. 1120); narrow plaited bands; shells of Humboldtiana montezuma Pilsbry; twill-pad sandals, some from the bottom level of Frightful Cave gave a radiocarbon date of 5345±400 B.C. (Crane, 1956, p. 669). It is possible that certain types not shown may also pertain to this complex:

ARCHAIC CULTURES: NORTHEASTERN FRONTIERS

round self-pointed dart foreshafts (fig. 10), choppers of limestone (fig. 7), spatulate bone awls (fig. 25), agave needles (fig. 23), scored sotol (?) bases or "buttons," the use of elk,* antelope,* coatimundi, bison,* and bear (including grizzly). By the time the cultural refuse in Frightful Cave had reached a height of 50 cm. above the sterile cave floor, these artifacts were no longer being manufactured, although a few examples, evidently "brought up" in the deposits from their original locations, were found at higher levels. Other artifacts, not of the Cienegas complex but from the bottom level, produced radiocarbon dates of 6915±350 and 6068±350 B.C. (Crane, 1956, p. 669). However, artifacts from the middle level, 50-100 cm. above the cave floor in Frightful Cave, yielded radiocarbon dates of 7585±550 and 7345±400 B.C. (Crane and Griffin, 1958a, p. 1104), i.e., earlier than artifacts from the bottom level. In view of the nature of the cave deposits and the isolated and out-oforder character of these two dates, it is probable that the objects had been "brought up" from an earlier level and that they indicate dates for the bottom level rather than the middle and, thus, for both the Cienegas and Coahuila complexes, i.e., for the cultural matrix in Coahuila at its earliest presently known date. From the general run of these dates, it seems that the types of the Cienegas complex were abandoned sometime between 5000 and 4000 B.C. Coahuila Complex The Coahuila complex was the major cultural matrix in central and northern Coahuila from the earliest times to the latest known stratified deposit. It forms the greater portion of the cultural corpus of which, however, the other recognized complexes were also parts. It is known in sites extending from the Rio Grande to the northern edges of the Laguna District and from the front ranges of the Sierra Madre Oriental on the east to the Coahuila-Chihuahua

border on the west. Four of these sites have been stratigraphically excavated: Frightful Cave, Fat Burro Cave, Nopal Shelter (CM28), and CM-37 which was unrewarding and soon abandoned. The remaining are mortuary sites without stratigraphy or sites in which only unrecorded excavations were conducted or surface collections made. The earliest known dates range between 7600 and 7300 B.C.; the latest radiocarbon date is A.D. 185±250 from the top level of Frightful Cave (Crane and Griffin, 1958b, p. 1120). The latest cross-cultural, comparative date is in the 12th century, based on the finding of sherds of El Paso Brown pottery on the surface of several sites which produced surface finds of Coahuila complex materials. From archival sources, however, it seems very probable that the Coahuila complex endured until the arrival of the Spaniards (de León, 1909; Beals, 1932; Martínez del Río, 1954; W. W. Taylor, 1956, p. 231). The Coahuila complex was not static. Its forms underwent variation, even though the classes and types of artifacts remained notably constant. These microvariations in time create a picture of cultural change which has both depth and breadth. Partly responsible for this was an environmental change, a gradual desiccation affecting both the natural habitat and its human occupants (Gilmore, 1947, p. 163, L. Johnson, 1960, p. 170). It bore upon aboriginal population particularly by way of changing its cultural ecology and making subsistence less secure. The result seems to have been a loss of cultural integration and stability and a consequent increase in group nomadism and typological variation, together with a decline in craftsmanship. It also had the effect of shifting the balance of subsistence from animal to plant foods, of increasing the use of fibrous desert plants at the expense of woody plants, and possibly of increasing ceremonialism concerned with the dead. Toward the very end, it brought increasing cultural contacts with neighboring peoples 63

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

and an influx of material imports or foreign ideas, or both. Throughout the temporal span of the Coahuila complex, the people, pursuing a nomadic, hunting-and-gathering way of life,* occupied caves and shelters, but it is evident that the utilization of such sites was occasional and that most of the living was carried on in the open.* An exception to this is the lower level of Frightful Cave, where floors prepared by the use of water suggest a more permanently and formally occupied habitation. Generally, however, when they did take shelter, the people seemed to choose particular sites which were used over and over again, whereas they made little or no use of others which today seem equally attractive. The preferred locations were near the mouths of canyons or within relatively easy access of the pediment slopes and the alluvial fans (locally called "monte") and the margins of the playa basins, i.e., places where plant and animal resources were the most accessible in variety.* This exercise of choice implies very little population pressure, and this in turn implies a deterrent to population growth. Although the Spaniards were horrified at the food of the aborigines, it seems hardly possible that scarcity was the cause of the small population; for people accustomed to eat what the north Mexican desert offers, there is really no scarcity of food (W. W. Taylor and González Rul, 1960). There is considerable abundance of a localized and seasonal sort* whose exploitation required only mobility and free access to large tracts of land. The ethnocentric reactions of the Spanish were directed against the quality and kind of native food; they do not mention that the people were starving or that they worked very hard to obtain what they did eat. De León (1909, p. 40) says that they were improvident and concerned only with the day's food, not that of the morrow. To the Europeans the diet seemed inadequate more from custom than from any real 64

scarcity of resources. What, then, kept the population low and the people occupying only a relatively few of the available sites? I suggest two answers to this question or, more exactly, a single subsistence dilemma with two points working somewhat at cross purposes: (1) the extremely low nutritive value of the principal dietary resource, the wild plant foods which required a tremendous amount of far-ranging in order to provide a bare subsistence, and (2) water, whose scarcity and localization put restrictive, i.e., just the opposite, pressures on the aborigines. Even in the more humid, early days, Coahuila was an arid to semiarid desert. The few running streams made water scarce and stringently localized. People could gather food and hunt almost anywhere, but had to return to a known and dependable supply of water. Furthermore, "dependability" would mean being able to count on finding the water available. This in turn would mean either fighting to maintain the availability or establishing some sort of social and/or political control. Since the availability of so vital a resource could hardly have been subject, on a day-in-dayout basis, to the vagaries and fortunes of fighting, it seems most probable that there would have developed an accepted and somehow controlled territoriality based on recognized rights to water, much like the territoriality based on hunting rights in other parts of the world. The seasonal character of the major plant resources and the need for abundant supply because of their low quality would have made these territories large and mobility within them of prime importance. This mobility, wideranging but tied to water supply, I have called "tethered nomadism" (W. W. Taylor, 1964). The radius of these waterbound wanderings would have been dependent on the types and amount of the seasonal food resources and on the amount and distribution of water resources. Since water in northeastern Mexico is almost entirely confined to the mountain masses (except the Rio

ARCHAIC CULTURES: NORTHEASTERN FRONTIERS

FIG. 3—EARLY AND MIDDLE COAHUILA COMPLEX POINTS. Top row: Jora points from Fat Burro and Frightful caves. Middle row: Fragua points from Fat Burro Cave and Nopal Shelter. Bottom row left: Duran points from Fat Burro Cave, CM-31, and CM-32. Bottom row right: Gobernadora points from Nopal Shelter, Frightful Cave, and CM-32. (Photo, Wyatt Davis, for Smithsonian Institution.)

Grande and its few perennial tributaries), groups of people were probably localized around these features, occupying the canyon mouths and the skirting monte within easy reach of water and from which they could more easily exploit the surrounding lands: mountain, canyon, monte, and flat. Experience in modern Coahuila shows that this pattern, with only minor modifications due to more elaborate technology, prevails today: people live in settlements around the fringes of the mountain blocks and know their own mountain and its immediate surroundings but are relatively ignorant and often fearful of other mountains and other people clustered around them. Evidence from the Coahuila complex suggests that there were changes in the intensity of nomadism during the course of its existence. At first, during and just after the Cienegas complex, the culture was relatively sedentary and localized, and it

showed internal evidence of being well integrated and stable. This was probably the classic period of "tethered nomadism." By the middle level of Frightful Cave, possibly owing to pressures caused by increasing desiccation, cultural integration and stability appear to have weakened. This is inferred from a noticeably growing heterogeneity of cultural (artifactual) form combined with a degeneration of craftsmanship. Increased nomadism is attested by a greater number of occupied sites and an extension of their distribution throughout much of Coahuila. Also some of the lithic types from this middle period point to considerable cultural contact with outside areas. Finally, toward the end of the Coahuila complex, with the appearance of the Jora complex (discussed below), cultural integration seems to have been restored, evidently under the aegis of foreign influence and on a basis of "ranging," rather than "tethered," nomadism. 65

FIG. 4—EARLY AND M I D D L E COAHUILA COMPLEX POINTS. Top row and left three of middle row: Espantosa points from Frightful Cave. Middle row, fourth: unnamed point from Fat Burro Cave. Middle row, others and foreshafted: Socorro points from Fat Burro Cave. Bottom row left: Socorro points from Fat Burro Cave, heavy subtype; two Socorro points. Bottom row right: two unnamed forms. (Photo, Wyatt Davis, for Smithsonian Institution.)

FIG. 5— M I D D L E AND LATE COAHUILA COMPLEX POINTS. From Fat Burro Cave, Nopal Shelter, Frightful Cave, CM-31, CM-32. (Photo, Wyatt Davis, for Smithsonian Institution.)

ARCHAIC CULTURES: NORTHEASTERN FRONTIERS

In the Coahuila complex there are three major industries based on material of manufacture: wood, plant fiber, and stone. Those based on bone, antler, hair, fur, hide, feathers, shell, minerals, seeds, and other plant products are very minor. Even the stone industry is a minor one in the early stages, although it becomes important in later ones. The ratio of stone to wood to fiber throughout Frightful Cave is approximately 1:6:26, these figures representing the average numbers of artifacts of the respective materials per cubic meter of deposit excavated within the site. In other sites, because of conditions of preservation, special circumstances of occupation, and probably later date, stonework is very much more abundant in relation to the other categories of material. All things considered, Frightful Cave gives the most realistic picture of aboriginal culture in ancient Coahuila. STONE. Viewed as a whole, stonework appears to increase in quantity but to become less formalized and less internally integrated as time goes on. In the blade industry at the beginning, a few bifacial types, formalized and consistent, were quantitatively dominant, and bifacially chipped artifacts of a single form had many uses. In later times, artifacts of a single form and evidently of a single use were made by both bifacial and unifacial techniques. In other words, an original mode of form-technique with multiple use gave way to a mode of form-use with multiple technique. Further, at any given point of time the numerically dominant types of blades were consistently produced by pressure flaking; the same types, when coming in or going out of fashion, were made by both pressure and percussion, the latter often being the more common. Gradually, unifacial blades in considerable typological diversity began to compete with the bifacial. Later, the numbers of unifacial types decreased although the total frequency was maintained; and still later, the proportional representation of unifacial and bifacial types began to de-

crease, whereas unretouched flakes with no typological consistency at all showed a marked proportional and absolute increase. These facts, it seems to me, indicate a definite tendency toward progressive formalization (already accomplished in bifacial blades at the moment of our first knowledge) and then a deformalization through time within the bifacial-unifacial blade categories. This implies a definite fluctuation in the typological conceptualization of the artisan himself. The blade industry in general shows a loss of craftsmanship, although possibly not utility, in the progression from bifacial to unifacial to utilized flake artifacts. Points also show a definite formality at the start and a progressive deformalization evidenced by a proliferation of forms, virtually untypable, in the later epochs (with an ultimate re-formalization at a much later date in Jora complex times as described below). The earliest types are oval, "laurel leaf" in shape (Espantosa and Fragua points, figs. 3 and 4; affinis Lerma, Refugio, Abasolo per Suhm and Krieger, 1954). Overlapping these, but with somewhat later chronological position, is a single type having a large, contracting stem, frequently serrated, and with strong barbs (Jora point, fig. 3). Following these, and again overlapping somewhat, come a heterogeneous lot of notched and stemmed points, only a very few of which are enough alike to warrant being placed in types (figs. 3, 4, 5). Quantitatively, stone points are notably scarce in all stratigraphic levels, but the absolute frequency increases in time. One more thing about the points: at first, shape is consistent, while length, width, and therefore weight vary greatly; later, shape loses its consistency, and the other attributes tend toward stability, although they still have considerable range. These observations suggest that the early points, undoubtedly used for tipping atlatl darts, could be of varying sizes, although in the beginning certain shapes were adhered 67

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

FIG. 6—COAHUILA AND JORA COMPLEX ARTIFACTS. Top row: flake points from Nopal Shelter and Frightful Cave; thinned base triangular point from Nopal Shelter (Taylor Thinned base ? ) ; drill from Frightful Cave; four core "finger choppers" from Fat Burro Cave and Nopal Shelter. Middle row right: core scraper from Fat Burro Cave. Middle and bottom rows: limestone bars from Fat Burro Cave, Nopal Shelter, and Frightful Cave. (Photo, Wyatt Davis, for Smithsonian Institution.)

to with regularity. Later, possibly when the atlatl was giving way to the bow, points had to conform to a more rigid standard of weight and therefore size, but shape, for some reason, was no longer of as much concern to the artisans, possibly because, as seen in other analyses, the culture in general was becoming less formalized and probably less integrated, craftsmanship was breaking down, and distracting outside influences were on the increase. Metates of the slab variety have a wide distribution, although they are not numerous. This suggests either that pounding and milling were not primary food-processing techniques or that these large and bulky artifacts were "brought up" in the deposits and represent cumulative use over long spans of time. Our present impression is that both these factors pertain, although 68

other evidence shows that food preparation on metates was relatively less in the early epochs of the Coahuila complex. Some metates are pitted, indicating pounding of hard objects such as bones* or the walnuts* whose remains are found in quantity in the early levels. Basin metates have not been found in excavated sites and seem to have a northerly and westerly distribution in Coahuila; we therefore believe them relatively late. Striations on metates indicate that manos were used in a longitudinal, straight movement. The manos are mostly arroyo cobbles of limestone; basalt and other stones are obviously foreign and generally late in the Cuatro Cienegas Basin sites. The majority of manos are pitted. There are no true rocker manos, although all manos are of the small, "one-hand" kind. Cores are not common even in sites, such

ARCHAIC CULTURES: NORTHEASTERN FRONTIERS

as Fat Burro Cave and Nopal Shelter, which were certainly chipping centers. This scarcity may be due to the fact that cores ended their careers as fire-rock, abundant everywhere. Only one artifact that could possibly have been used as a drill has been found in our work; it was in the bottom level of Frightful Cave (fig. 6). Although we did not specifically search for burins (Epstein's finding of such implements in Texas [1960b] postdated our laboratory work), none were identified, and my feeling is that none were present. Only seven choppers were found, an inexplicably small number in view of the quantities of wood and fibrous plants processed by the aborigines; they are of limestone and very haphazardly made (fig. 7). Of the six stone ornaments discovered (fig. 8), one from Fat Burro Cave is of selenite; the others are of the ever-present, ever-used, dull, local limestone. One was painted and another was lustrous black. Although unusual materials were known and a desire for color is in (slight) evidence, the aborigines apparently did not have much interest in ornament, color (the arroyos are full of brightly colored stones), shininess (calcite and selenite were known and are common in the local rocks), or even in decoration. The few ornaments found are early. Bedrock mortars occur at 11 sites, mostly in the northern and western sectors of the state. Only two were in sheltered sites; obviously proximity to habitations of any sort, sheltered or otherwise, was not a prerequisite. Nearness to the habitat of plant foods, i.e., the monte, seems to have been a more impelling factor. Mortar holes are most commonly found in groups, and in many instances many or all of the holes are of equal utility, i.e., their depths are not unusably great, nor are they nascent. We therefore infer that many of the holes were used at the same time and that, if groups of women could gather to process foodstuffs in mortar holes away from habitation sites, either they did so under armed guard

FIG. 7—LIMESTONE CHOPPERS. Coahuila complex. From Fat Burro and Frightful Caves. (Photo, Smithsonian Institution.)

or the threat of attack was nil. We believe the latter to have been the case. Again the conjunctives point to some sort of social and/or political control, as with water supply. Rock grooves, the so-called "sharpening grooves" found in the Trans-Pecos of Texas (V. J. Smith, 1938, p. 222), have been found in 12 sites (fig. 9). What they represent cannot be argued at this time, but they certainly were not used for sharpening any implement so far found in Coahuila; in fact, their nature would seem to preclude them from sharpening anything. Rock midden circles, the so-called "mescaleros," have been found at seven locations in central and northern Coahuila in contexts of the Coahuila complex; this distribution does not take into account the 16 found along the Rio Grande in the area of the Diablo 69

FIG. 8—ORNAMENTS. Coahuila, Jora, and Mayran complexes. Top to bottom, left to right: Acacia seed beads on fiber cord,* Canis latrans bone beads, Antilocapra americana dewclaw on thong, black limestone pendant, Odocoileus or Antilocapra hoof-covers on fiber cord, limestone button or pendant painted red, Spondylus princeps bead, engraved limestone pendant, limestone button or bead, bone and seed beads, Lampsilis siliquoidea pendant, three strangle-groove bone beads of Lepus californicus, tubular bead of serpulid marine worm (often erroneously called Vermetus in the literature), bone bead, Lampsilis siliquoidea pendant, selenite button or pendant, dense bone button. (Photo, Smithsonian Institution.)

ARCHAIC CULTURES: NORTHEASTERN FRONTIERS

Reservoir (W. W. Taylor and González Rul, 1960). Their distribution is definitely northern (and also western in central Coahuila), geographically associated with the Rio Grande and its approaches. Limestone bars were found in Frightful Cave, Fat Burro Cave, and in Nopal Shelter, three of the four stratigraphically excavated sites (fig. 6). Whatever their aboriginal use, they were not natural in the sites but had been brought there by human agency. They bear no paint but many show signs of use, mostly rubbing but some pounding. Discolorations suggest grease, soot, blood, hematite, and/or some combination of these. Although three of the five in Frightful Cave came from the bottom level (the other two were top level), evidence from the other sites indicates them to be relatively late. Similar objects have been found along the Rio Grande in northern Coahuila and Texas (Taylor and González Rul, 1960; Epstein, 1960a, pp. 99-100; Archaeological Salvage Program, 1958, pp. 22-23). WOOD. The Coahuila complex was primarily concerned with wood and fiber and only secondarily with stone and other materials. Much of the stonework seems to have served for processing wood and fiber. In the bottom level of Frightful Cave, for example, only two notched dart foreshafts (for stone points) were found, as against seven self-pointed wooden foreshafts. Also in the bottom level, there were 725 fiber artifacts, 273 wood, and only 46 stone. Notched foreshafts increased from bottom to top, conjoining with the increase in stone projectile points to strengthen the context (fig. 10). The seven self-pointed wood foreshafts from the bottom level are long, heavy, and round, often with some of the bark still adhering; the other four (from the top level) are faceted to a quadrangular section, are much shorter and lighter, and retain no bark (fig. 10). The one bunt foreshaft, very large, was found in the bottom level (fig. 10). Shaft wrenches are late in Frightful

FIG. 9—POLISHED LIMESTONE W I T H GROOVES. Often called grooves." From CM-70a. (Photo, Institution.)

BOULDER "sharpening Smithsonian

Cave, the only place where they have been found (fig. 11). Atlatls are of two types: (1) the mixed or Mexican variety with both groove and engaging-hook is late in time; (2) the "male" type, represented by three examples, was fashioned from a hardwood limb with a natural fork formed into a large and powerful hook. These were found one in each of the three levels in Frightful Cave (fig. 12). Grooved clubs* were present in quantity and were of the three-groove, "southern" variety (Heizer, 1942); they ranged from bottom to top and were quite surely associated with bows as well as with atlatls (Kellar, 1955, p. 307). Only six digging sticks were found. There were 19 specimens of fire tongs for use in handling the hot rocks for stone-boiling (fig. 11). The netting reel (fig. 11) is unique in the literature as far as can be ascertained. Two notched sticks, surely musical rasps,* came from the middle and top levels of Frightful Cave. Hearths for fire-drills were scarce, probably because used ones had been 71

FIG. 10—DART FORESHAFTS. Coahuila complex. Top row: notched and bunt. Bottom four rows: self-pointed—top two rectangular, bottom two round with bark. From Fat Burro and Frightful Caves. (Photo, Smithsonian Institution.)

FIG. 11—WOODEN ARTIFACTS. Coahuila, Jora, and Mayran complexes. Top: fire-rock tongs. Middle: netting reel, three pegs, miniature cradle from infant burial probably of Mayran complex in Fat Burro Cave, two withe loops, split-twig loop of Jora complex, fire hearth, shaft wrench. (Photo, Smithsonian Institution.)

72

ARCHAIC CULTURES: NORTHEASTERN FRONTIERS

thrown into the fire when discarded (fig. 13). Drilling was done in later times with the hearth placed across the body, but in early times with the hearth in line away from the body. Rubbed and scorched sticks show that fire was also made merely by rubbing two sticks together.* Whether what we call "burial sticks" (fig. 14) belong with the Coahuila complex or with the Jora complex is uncertain; they have been found only in burial sites, lying loose on the surface with other cultural objects for which assignment to one or the other complex is undecided. They range from 420 to 1390 mm. in length and are generally made of the flowering stalk of Agave lechuguilla (?) or Yucca sp. (?). Their larger ends have been modified as strangle-grooved, eyed, end-flattened, fiber-wrapped, or combinations of these. They pertain unquestionably to the burial complex. Another frequent component of this complex is the stick arc, clearly either the frame of a carrying net* or a cradle. In all excavated sites were quantities of cut sticks, finished and unfinished pegs, and signs of a great amount of wood working (fig. 11). FIBER. Fiber is by far the most abundant material-of-manufacture in the Coahuila complex. Among the artifacts from Frightful Cave there is over 20 times as much fiber as stone, and over four times as much as wood. These figures pertain to manufactured artifacts only, not to fiber "matrix" items such as quids, grass-lined cache pits, food and manufacturing refuse, all of which were very common in the deposits. Twisted-fiber cordage (mostly of Agave* and Hesperaloe) is the most abundant fiber artifact. From Frightful Cave were recovered 1193 pieces having a total length of 185.74 m.; there is less, both proportionately and absolutely, in Fat Burro Cave. Z-twist is most common, approximately 10 times S-twist. Three- and four-ply cordage is extremely scarce. Sandals are the next most numerous fiber

FIG. 12—ATLATLS. Coahuila complex, a, Roundshafted male type, early Coahuila complex. b,c, Two Mexican or "mixed"-type hooks, d, Engraved section of a shaft, e, Proximal end. f, Hook end of a broken but complete Mexican type. From Frightful Cave and CM-73. (Photo, Smithsonian Institution. )

artifact. In Frightful Cave there were 959 of them. The other sites did not produce nearly as many, even allowing for the differences in cubic meters excavated. In Frightful Cave twill-pad sandals (fig. 15) came from the bottom level. They are early in Fat Burro Cave also. Checker-pad sandals (figs. 15, 16) are a bit later, sewed sandals (fig. 16) later still, and braided ones 73

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

Cave the incidence of coiled basketry, proportional to other fiber categories, increases from bottom to middle to top (less than 10 per cent, around 30 per cent, and just over 60 per cent of the category, respectively); the same percentage sequence holds for five of its attributes: bundle foundation, half-rod foundation, split stitch, and each direction of stitch slant. This consistency is conjoining and supports the validity of the several distributions. Split stitch is present in 85 per cent of all coiled basketry in Frightful Cave; 63 per cent is of half-rod foundation, 31 per cent bundle, 6 per cent whole rod. The proportions in Fat Burro Cave are 76 per cent bundle, 14 per cent half-rod, and 8 per cent whole rod. These data suggest that bundle foundation is a relatively late trait and that whole rod is early. Interlocking stitch is both rare (only five specimens) and definitely early. Counterclockwise stitch on the work surface is superseded by clockwise. Convex work surface is slightly later than that of concave. In plaited matting, both the twill and the checker techniques are present from bottom

FIG. 13—SPLIT-TWIG LOOPS. Jora complex. From Fat Burro and Frightful caves. (Photo, Smithsonian Institution.)

(fig. 17) are top level. Two-warp plaited sandals are found from bottom to top, but the three-warp variety is top level only. The seven types of sandal ties apparently have significant chronological proveniences but cannot be discussed here. Coiled basketry appears from bottom to top in both Frightful and Fat Burro caves. Its greatest frequency does not come until the top level of the former and the middle level of the latter (fig. 18). In Frightful 74

FIG. 14—BURIAL STICKS. Complex uncertain, probably Coahuila. Only larger end illustrated to show varieties of working. From CM-59d, CM64, CM-73. (Photo, Smithsonian Institution.)

ARCHAIC CULTURES: NORTHEASTERN FRONTIERS

FIG. 15—SANDALS. Top: Coahuila complex twowarp, left mostly Hesperaloe funifera in "crude" state, right Agave lechuguilla decorticated. Bottom: Cienegas complex, twill pad of two subtypes. From Frightful and Fat Burro caves and CM-65. (Photo, Smithsonian Institution.)

FIG. 16—PLAITED BAG SELVAGE AND SANDALS. Top left: selvage. Top right: checkerpad sandal, Cienegas complex. Bottom left: sewed sandal, late Coahuila and Mayran complexes. Bottom right: only fishtail sandal in collection, Coahuila or Jora complex. From Fat Burro and Frightful caves. (Photo, Smithsonian Institution.)

to top, but checker equals twill only in the top level of Frightful Cave (fig. 19). Knotless netting ("coil without foundation") is present throughout the stratigraphic column (fig. 20). Knotted netting is late in Frightful Cave. Twined textiles appear early and in major proportion, seem to diminish, and then come on again late in the sequence (figs. 21, 22). Virtually identical percentage progressions are seen in hard twining (mats, bags, baskets, fringes) and in soft

twining (aprons, robes). In both hard and soft twining most of the early specimens have their weft elements slanting from u p right to down-left, whereas in later times the reverse slant is more common. Rosettes (fig. 23), radiocarbon dated at 1275±350 B.C. (Crane and Griffin, 1958b, p. 1120), appear, from bottom to top in Frightful Cave, as 2 per cent, 24 per cent, 63 per cent; they have been found in only one other site, CM-37, in Cave Canyon near 75

FIG. 17—BRAIDED SANDALS. Late Coahuila, Jora, or Mayran complex. From Fat Burro and Frightful caves. (Photo, Wyatt Davis, for Smithsonian Institution.)

FIG. 18—COILED BASKETRY. Mayran complex, from CM-79. Diameter of smallest: 10 cm; diameter of top right (with human mandible): ca. 29 cm. (Photo, Smithsonian Institution.)

76

ARCHAIC CULTURES: NORTHEASTERN FRONTIERS

FIG. 19—PLAITED MATTING. Top: checker, Coahuila complex, from Frightful Cave, 43 by 33 cm. Bottom left: twill with red paint outlining woven design, Mayran complex, from infant burial in Fat Burro Cave. Bottom right: twill with woven design, Mayran complex, from CM-74. (Photo, Smithsonian Institution.)

Fat Burro Cave and Nopal Shelter. Not counting those on sandals, there were 1761 knots of all kinds recovered. Of these, 91 per cent are comprised of square and overhand knots; the rest included granny, figure-eight, slip, sheet bend, half-hitch, and one double-fold knot. In Frightful Cave more than 95 per cent of all knots were found in the front, in the fire-rock area, evidently left there during food preparation. The sheet bends were distributed middle and back, the half-hitches front and middle. Vertically, knots follow quite closely the general trend of all fiber artifacts from the site; but sheet bends are concentrated in the middle level, halfhitches in the top level. A series of scari-

fiers* (fig. 24) is significant in view of the fact that the archives refer to tattooing and bloodletting, for both of which tasks these artifacts would be admirably suited. BONE. Bone awls and antler artifacts (fig. 25) have a "normal" vertical distribution in Frightful Cave, i.e., similar to that of the quantitatively major categories. The awls are definitely located front and back, i.e., in the fire-rock and waste areas. Not a single ulna awl has been found in Coahuila. Reduced cannon-bone awls are later, whereas those not modified are early. No bone beads were found in Frightful Cave, indicating that the considerable numbers found in other sites, particularly burial sites, are to be considered late and possibly as custom77

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

FIG. 20—NETTING. Top left: crude fiber knotted (sheet bends) bag, Coahuila complex, from Frightful Cave. Top right: knotless, Mayran complex, from CM-79. Bottom left: knotless bag, 8 cm. deep, 12 cm. long. Bottom center: knotted (double half-hitch or clove hitch), Coahuila complex, from Frightful Cave. Bottom right: knotless bag, Mayran complex, from CM-79. (Photo, Smithsonian Institution.)

ary grave goods. Antler was used for both flakers and flaking anvils. One set of deer antlers* was found in a burial cave with two deer mandibles crossed and tied into the branches; from the accompanying material this specimen is thought to belong to the latter part of the Coahuila complex or to the Jora complex (fig. 26). The archives speak of the use of deer skulls in ritual (García Torres, 1856, p. 83). The concentration of miscellaneous bone remnants, especially deer, is early and diminishes in the upper levels of Frightful Cave. Notable is the proportionally great number of mandibles of many species of animal; 78

the quantity would indicate that this is a cultural phenomenon possibly connected with the practice of breaking up, hence destroying, marrow and other large bones that could be powdered and eaten.* Rodent mandibles were bound for reinforcement and used probably as gravers. OTHER MATERIALS. Shell* and minerals

are extremely scarce, although a considerable amount of hematite* was found in Fat Burro Cave. Fur and processed hide* were quite common in Frightful and Fat Burro caves; fur cord was made and evidently woven into textiles.* Featherwork* was present but rare. Human feces were ap-

FIG. 2 1 — T W I N E D MATS AND CONTAINERS. Coahuila complex, from Frightful Cave. Top row left: 34.5 cm. long. (Photo, Smithsonian Institution.)

FIG. 22—BAG AND APRON. Coahuila complex, from Frightful Cave. Left: twill plaited bag of two fabrics sewed together* with thong, 34 cm. deep. Right: twined apron of soft fiber with thong (?) waist belt. (Photo, Smithsonian Institution.)

FIG. 23—FIBER ARTIFACTS. Coahuila complex, except three. All from Frightful Cave. Top row: Agave needles, Cienegas complex; two scarifiers. Second row: scarifier of single A. lechuguilla leaf. Third row: small plaited pad or pillow; narrow plaited band, Cienegas complex. Fourth row: plaited "ears." Fifth row: completed rosette; fiber cross, Cienegas complex; unfinished rosette. Bottom row: self-wrapped strips. (Photo, Smithsonian Institution. )

FIG. 24—SCARIFIERS. All Coahuila complex and all from Frightful Cave, except small one with rodent teeth second from right in top row, which is possibly Jora complex and from Fat Burro Cave. Two left: a single Agave lechuguilla leaf turned on itself and bound; others are of Opuntia spines lineally arranged in various ways, including twined weave in right three. (Photo, Wyatt Davis, for Smithsonian Institution.)

pallingly common in all sites where preservation permitted; natural functions seem to have been performed without regard for modesty, sanitation, or probably even the normal routine within an occupied habitation site.* Quids of fiber,* the sucked-out remnants of roasted succulents used for food, were everywhere and in great quantity, especially frequent in the fire-rock areas (W. W. Taylor, 1948, p. 172). Quantities of the narcotic mescal bean, Sophora secundiflora, were found in Frightful Cave; this shrub grows in the canyon today but the presence of the beans in the cave surely represents purposeful collecting by humans (Campbell, 1958). Serpulid marine worm tubes (Protula superba?) fashioned into beads were found in CM-64, a burial site; these have also been found in the Hueco Caves (Cosgrove, 1947, p. 152), caves in northern Chihuahua (Sayles, 1936b, pl. 16), the Mogollon and Harris villages (Haury,

1936b, pl. 19, fig. 30), the colonial period at Snaketown (Gladwin et al, 1937, pl. 113), upper Rio Fuerte in southern Chihuahua (Zingg, 1940, p. 25), San Cayetano and Babocomari villages of southern Arizona (DiPeso, 1951, p. 190; 1956, p. 100), at Santa Ana near Zape, Durango (Mason, personal communication). In the published reports, they have generally been identified as Vermetus, an erroneous identification which obscures the fact that the presently known locus of these animals is the Pacific Ocean, specifically off the Santa Barbara coast of southern California. Jora Complex The Jora complex consists of a number of traits which were part of the late cultural corpus in Coahuila. It is not a separate ethnic entity but rather a congeries of chronologically significant traits within a cultural continuum, the body of which we have 81

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

FIG. 25—BONE ARTIFACTS. Left: plastron of Kinosternon sonoriense, edges smoothed, from Fat Burro Cave; awl of earlier (?) type; awl of later (?) type; two spatulate awls, Cienegas complex ( ? ) ; antler flaker; antler flaking anvil from Fat Burro Cave. Lower left: rodent jaw bound as graver ( ? ) , from Fat Burro Cave. All Coahuila complex and Frightful Cave unless otherwise stated. (Photo, Smithsonian Institution.)

been calling the Coahuila complex. Possibly some of these traits are of foreign origin, specifically from the region of La Junta de los Rios (Conchos-Rio Grande confluence). The traits so far recognized are: small projectile points obviously for use on arrows* (fig. 27), small self-pointed wooden foreshafts also for arrows,* split-twig loops (fig. 11), small snub-nosed flake scrapers sometimes notched, use of predominantly lightcolored chert for chipped implements, use of basalt and sandstone for manos, basin metates and rocker manos, bedrock mortar holes, rock midden circles, petroglyphs, ceramics. Only the first seven traits have been found in stratigraphic position; the rest are surface finds placed (sometimes uncertainly) in the complex on the basis of seriation and association. 82

The types of small points are identical, individually and collectively, with points already known from Texas, particularly from the central Texas and Bravo Valley aspects but also from certain east and south Texas foci such as Henrietta, Frankston, Galveston Bay, and Rockport (Jelks, Davis, and Sturgis, 1960; Suhm and Krieger, 1954). Split-twig loops have been found in the Trans-Pecos, and at least one specimen was on exhibit at the Sul Ross State College Museum in Alpine, Texas, in 1940. The center of distribution of the small notched end scrapers seems to be farther south, possibly in the state of San Luis Potosí (Beatriz Braniff, personal communication); Aveleyra found several in the Laguna District (Aveleyra et al, 1956, p. 75, figs. 7, 9), but the Coahuila specimens seem to be the feather

ARCHAIC CULTURES: NORTHEASTERN FRONTIERS

edge of the range (W. W. Taylor and González Rul, 1960). Basin metates, basalt manos, bedrock mortar holes, rock midden circles all seem to be more common in the northern and western sectors of Coahuila, i.e., near Rio Grande and Trans-Pecos of Texas, particularly the Conchos-Rio Grande confluence. In this region there are exposures of the types of extrusive rocks used to make these manos. Ceramics are scarce in Coahuila and all are surface material, including the two sherds found in the Candelaria burial cave (Bernal in Aveleyra et ah, 1956, p. 205). Pottery is of several kinds, only two of which are known at the present time to have relationships elsewhere: El Paso Brown typical of the Jornada Branch (Lehmer, 1948, p. 94), this particular variety dating from the 12th century (Mera, letter of April 8, 1942), and another rather generic, undated type, both unpainted and painted red, which has been identified as relating to the brown-ware sites of southern New Mexico (Mera, ibid.; Jennings and Neumann, 1940, p. 6). There is no way at present to associate this pottery definitely with any other cultural material from Coahuila since it has not been found in situ; however, its distribution and the nature of the sites and associated surface material point to its connection with the Jora complex, and possibly with the late Coahuila complex. The Jora complex, to judge from the very little we know about it, seems to represent a reconstitution and reintegration of culture in (northern) Coahuila. This time, however, outside influences are definitely indicated, unlike the Cienegas complex, for the origin of whose obvious stability and integration we have no signs at all. Mayran Complex This material comprises that found by Aveleyra in Candelaria and Paila caves (Aveleyra et ah, 1956) and that of the Palmer "mummies" and much of the material recovered by the nonprofessionals

FIG. 26—ANTLERS. Odocoileus antlers with mandibles tied across; probably late Coahuila complex, CM-74. (Photo, Smithsonian Institution.)

working out of Torreon and Saltillo (Studley, 1884; McVaugh, 1956, p. 80, 133 ff.; Barragán, Cárdenas, and Valdés, 1960). An infant burial attributable to the Mayran complex was found in Fat Burro Cave, and a disturbed burial cave (CM-79) with multiple interments and Mayran grave goods was salvaged in west-central Coahuila. Characteristic of it are: elaborate textiles of netting cloth and loom-woven material including cotton (Barragán, Cárdenas, and Valdés, 1960, fig. 30), large boldly chipped triangular knives which are often hafted* (fig. 28; Aveleyra et ah, 1956, pls. 12-16), snub-nosed flake scrapers and small projectile points of Jora complex or Bravo Valley aspect affiliations (Aveleyra et ah, 1956, pls. 4, 7, 9, 11; Barragán, Cárdenas, and Valdés, 1960, fig. 18), a variety of notched and stemmed medium-sized points of Coahuila complex affinities (Aveleyra et ah, 1956, pls. 1, 2), elaborate bone and shell bead 83

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

FIG. 27—POINTS O F T H E JORA COMPLEX. Top row and second row left: Sierra Madera (Toyah?). Second row right: Cienegas. Third row: Nopal (Perdiz?). Last three third row right: Ojo (Clifton?). Fourth row: El Muerto (Fresno?). Bottom row: various untyped. From Fat Burro Cave, Nopal Shelter, CM-32, and CM-37. (Photo, Wyatt Davis, for Smithsonian Institution.)

"flowers" (ibid., pls. 22-24), bow and arrow, twig-frame net baskets (ibid., p. 153, pl. 46; Barragán, Cárdenas, and Valdés, 1960, fig. 25), and an extensive mortuary complex involving bundle burial of whole cadavers (Barragán, Cárdenas, and Valdés, 1960, fig. 29) with accompanying grave goods enveloped in textiles and entombed in large numbers probably over a considerable time in limestone caves (often solution cracks), not habitation or even habitable locations. Since there are, in addition to the projectile points, a number of accompanying traits which belong to the Coahuila complex, it is possible that this mortuary complex has some antiquity, although it certainly endured quite late, possibly even as late as the end of the 16th or beginning of the 17th centuries, if we can credit the identity of the two potsherds found in Can84

delaria Cave and those from a mission site in the Laguna District (Bernal in Aveleyra et al., 1956, p. 205). A number of peyote buttons* strung on a cord were found as part of the Mayran complex grave goods in the burial cave, CM-79. Coastal Plain Complex No excavation has been done in the coastal plain province of Coahuila. Material attributed to this complex comes entirely from surface collecting (Miillerried, 1934). Other surface collections which appear to be typologically like those found on the coastal plain (fig. 29; W. W. Taylor, personal notes; Dudley Jackson, personal notes; Kirk Bryan, personal notes) have been made in the interior regions of Coahuila. On the other hand, these collections are quite similar to some from the Laguna Dis-

ARCHAIC CULTURES: NORTHEASTERN FRONTIERS

trict which include Mayran complex types as well. One thing appears certain: that this large, boldly chipped stonework, especially the large triangular blades with obvious affinities to Tortugas points of the Falcon reservoir and Tamaulipas areas, is not characteristic of the Coahuila complex. Whether this material represents a cultural or a chronological difference or both is a problem at the present time. My present hunch is that the Coastal Plain complex material is generically related to the Coahuila complex and contemporaneous with it during the last pre-Spanish and earliest conquest periods. The archives tell us that Indians traveled back and forth throughout northern Mexico from Zacatecas and San Luis Potosí into Texas and from the shores of the Gulf of Mexico to the mines of Chihuahua. Kelley (1955) has shown the effects of such movement in one particular instance, and it is not difficult to envisage a broad diffusion of material objects, and possibly also of ideas, from one end of the area to another. Thus the appearance in the west of these artifacts of more eastern affiliation is not as unexpected as might be supposed. COAHUILA SKELETAL MATERIAL. Skeletal material is not common in Coahuila, and, with the exception of that found with Mayran complex artifacts, its cultural associations are very uncertain. On the sterile gravels at the bottom of Frightful Cave, but in a pit excavated from approximately 50 cm. above, was the burial of an aged female dressed in loincloth and G-string, wrapped in a robe of twined vegetal fiber, and wearing a pair of two-warp sandals; she was lying flexed on a bed of rocks and twigs; her face had been covered by pricklypear pads, and a plaited mat had been tucked around and under her head. This was an early or early middle Coahuila complex burial, one of only three known to have been found in situ in occupation sites in Coahuila. Serological tests on desiccated tissue (W. W. Taylor and Boyd, 1943; Boyd, 1959) showed that this person had

FIG. 28—STONE ARTIFACTS O F T H E MAYRAN COMPLEX. From CM-88 and CM-89, Palmer Coll., Peabody Museum, Harvard University. (Photo, Wyatt Davis, for Smithsonian Institution. )

blood type B; with one example from site CM-59b in Coahuila (W. W. Taylor and Boyd, 1943), one from the Big Bend of Texas (Boyd and Boyd, 1937; Boyd, 1950, p. 249), and examples from the Paracas burials of coastal Peru (Boyd and Boyd, 1937), this is the only test on pre-European American Indian tissue that has shown blood type B. An infant burial in Fat Burro Cave had with it a miniature cradle, two nock ends of arrows and two arrow shafts, a string of cervid hoof-covers, a ball of cotton cord, a bone-bead necklace, traces of typical Mayran complex netting, a painted mat of twill plaiting, a coiled basket, several leather strips; the whole was wrapped in the tanned hide of what was apparently a mountain lion (Felis concolor). Most burials, however, were placed in small niches or shelters at some distance from habitation sites. Some seem to have 85

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

FIG. 29—COASTAL PLAIN, JORA, AND COAHUILA COMPLEX ARTIFACTS. Top two rows: typical surface-collected material from Coastal Plain complex sites in western Coahuila, CM-40, -41b, -49, -54, -50, -58. Bottom row: Jora and Coahuila complex material from central Coahuila, CM-55, CM-62. (Photo, Wyatt Davis, for Smithsonian Institution.)

been secondary bundle burials; many were enveloped in matting and covered with piles of stone. Grave goods were scarce, sometimes absent. Several types of artifacts were found consistently with such burials: burial sticks, cradles and/or net carrying frames, strangle-groove bone beads of jackrabbit bone, strung matting of marsh plants (Cyperaceae, Typha), plaited matting. In the Laguna District, the type locality for the Mayran complex, burial caves are large and contain many interments, evidently representing long use. The mortuary complex is elaborate, although the burial sticks so characteristic of the northern region have not been found. From this and the rather frequent evidence for the bow and arrow, it appears that these burials and, hence, the Mayran complex are late in Coahuila. 86

Except for the work of Studley (1884) and Hooton (1930, p. 233 ff.), almost no study has been done on human remains in Coahuila—or in northeast Mexico as a whole. From these scanty data and from the few measurements taken on other Coahuila skeletal materials, it is apparent that the population was dolichocephalic to hyperdolichocephalic (one cranial index of 64), of small stature, with small cranial capacity, orthognathous, leptoprosopic, and mesorrhine. They are quite closely comparable to the Texas Coast, the Abilene, and the Trans-Pecos series (Woodbury and Woodbury, 1935, p. 35 ff.), the Arizona and Pecos Basket Maker series (Hooton, 1930, p. 233 ff.), to the Pericu-Lagoa Santa type of Baja California and Brazil respectively (Woodbury and Woodbury, 1935, p. 43), and thus

ARCHAIC CULTURES: NORTHEASTERN FRONTIERS

also to the two "Cochise" skulls from Ventana Cave (Gable, 1950, p. 513). COAHUILA PARIETAL ART. Pictographs in Coahuila are not uncommon, but their cultural affiliations are obscure. Geometric figures in red and yellow seem to be the earliest; later, black and white are added, and hand prints appear. In the latest group zoomorphic representations include figures of horsemen in European, specifically ecclesiastical, dress. The horses are of the "Plains type"; the paintings probably were made by Indians from Texas, possibly the Comanche or Lipan who raided into Coahuila after 1780. None of the typical "Pecos River" pictographs has been found, with the exception of one series reported from near the Rio Grande (H. C. Taylor, 1948). Petroglyphs are quite common in the Laguna District (Carl Compton, personal communication; W. W. Taylor, personal notes; Barragán, Cárdenas, and Valdés, 1960, figs. 1416), but only one series has been found in the rest of the state. The former series may be connected with the Mayran complex and the latter with the Jora complex, although any assignment is highly speculative. CULTURE AND CULTURE-SEQUENCE IN TAMAULIPAS

MacNeish has set up three series of cultural phases in Tamaulipas (fig. 30). In the southern part of the state, only the earliest material, up to Almagre times, is typologically Archaic, but in the north this material continues into the historic period. There are certain differences between the Tamaulipas inventories and those from other regions of northeast Mexico and Texas; the cultural manifestations had generic relationships but, by the time of our first information, had developed distinctive subgroupings. Although full documentation is not possible here, the following judgments are based on detailed analysis of published and unpublished material. At the present writing, MacNeish's cul-

tural sequence from the Sierra de Tamaulipas (1958) is basic to a picture of this entire eastern region, including Tamaulipas, parts of Nuevo Leon and southern Texas, and probably eastern Coahuila. In regard to MacNeish's earliest or Diablo assemblage, I do not believe that its one distinctive specimen and its other 10 specimens, which can be typologically duplicated in the succeeding assemblage, warrant designation as a separate cultural entity. At the moment, it is more realistic to regard it as an early and meagerly represented component of the succeeding phase. On the other hand, the differences between the Lerma and Nogales phases deserve more consideration, since the latter has 10 stonework types not present in the former. The pre-Laguna (i.e., pre-Mesoamerican) sequence as a whole, beginning with Nogales and ending with Almagre, is essentially a single cultural continuum or tradition with only minor modifications, some patently due to differential preservation and some representing variations, perhaps, on a widespread, long-enduring, but basically single way of life. I am definitely skeptical that seven additional stone types (out of 50) comprising a mere 33 specimens (out of 1039), over a period estimated at 3500 years, constitute sufficient evidence for the designation of three distinct cultural entities or "phases." The finding of evidence for agriculture in the middle of La Perra level does not really alter the case, as MacNeish appears to believe (1960, p. 593). First, this is undoubtedly a function of differential preservation and, second, the addition of domesticated plants, according to MacNeish himself (1958, pp. 146, 201), does not seem to have made much difference in the aboriginal way of life. The rest of the cultural inventory shows no significant change. The real change comes in the following assemblage, the Laguna phase, in which definitely Mesoamerican traits appear in southern Tamaulipas: intensive agricul87

FIG. 30—Table of cultural phases and complexes, dates, and climate in northeastern Mexico (adapted with modifications and additions from MacNeish, 1958, Table 30, p. 192).

ARCHAIC CULTURES: NORTHEASTERN FRONTIERS

ture, permanent occupation in villages with plazas, house and temple platforms, ceramics. The Tamaulipan continuum represents a way of life much like that of Coahuila, i.e., largely nomadic with only sporadic, possibly seasonal, use of sheltered sites. In one component of Almagre times there is evidence for wattle-and-daub houses. Except for the perishable material recovered from a single component, the cultural inventory from Diablo to Almagre consists almost entirely of stonework, with a very few specimens of bone, antler, and shell. Projectile points followed generally the same sequence as in Coahuila (and to some extent in Trans-Pecos Texas): leaf-shaped points first, then contracting stem, then various medium- to large-stemmed and notched shapes, and finally small-stemmed and notched types. In Tamaulipas, however, following the leaf-shape points and contemporaneous with the first contracting stems, there is a class of basically triangular points (which has been split into three types on the shape of the base: Abasolo, Nogales, and Tortugas). This general group is numerically dominant from Nogales times onward, continuing as one of the most characteristic shapes until the Mesoamerican period in the south and the historic period in the north. It constitutes one of the major differences between stone artifacts from Tamaulipas (and possibly from the entire northeastern region of Mexico) and the early complexes of Coahuila. But, by the time of the Mayran and Coastal Plain complexes, these triangular types had reached Coahuila and serve to interrelate many of the late cultural levels throughout northeast Mexico and southern Texas. In southern Tamaulipas, the non-Mesoamerican sequence does not extend to small-point times, but in the north these types are part of the late, but still Archaic, inventory. Notable are the evidences for domesticated plants. Since these specimens of squash (Cucúrbita pepo) and maize (Nal-Tel, types A and Β)

come from deposits radiocarbon dated at about 2490 B.C, they constitute one of the earliest examples of agriculture in the Americas (Mangelsdorf, MacNeish, Galinat, 1956). Faunal and floral waste reflect environmental changes (fig. 30) and indicate subsistence patterns which appear to involve hunting, a small amount of agriculture by at least La Perra times, and a great amount of collecting wild plants. For the Sierra Madre of southwestern Tamaulipas, MacNeish set up another eight phases, some of which he says are like phases in the Sierra de Tamaulipas, whereas others fill gaps in that sequence. The finds in the Sierra Madre produced considerable information, particularly in the way of perishable materials. Domestic gourds, squash (C. pepo), chile, and possibly runner beans were recovered from the earliest assemblage, the Infiernillo, which yielded radiocarbon dates of 6585 and 6245 B.C. (Crane and Griffin, 1958a). It is notable that no maize was found in this context; Bat Cave corn turned up later in the Flacco assemblage, dating around 1800 B.C. (MacNeish, 1958, p. 194). However, MacNeish says that the people from beginning to end of the pre-Mesoamerican sequence were largely collectors despite their knowledge of agriculture. In fact, contrary to what MacNeish appears to believe, the cultural inventory seems to be remarkably similar to that in the Sierra de Tamaulipas, particularly in stonework but in other categories as well, if due allowance is made for expectable local variation, the differential occurrence of perishable materials, and the differences in inventory completeness which certainly must be assumed. I see no reason to isolate these congeries or to give them names different from those of the Sierra de Tamaulipas, at least according to our present knowledge. It would seem much better to set up two groupings: Diablo and Lerma in one, Nogales-Ocampo-La Perra-Almagre-Flacco-Guerro in the other. Where to put 89

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

Infiernillo is a problem which might be resolved by further excavation, but my present hunch is that it belongs with the second group because of its obvious affiliations with Nogales and its greater resemblance to La Perra than to Lerma. However, some new traits do appear, which indicate cultural relationships with other regions of northern Mexico and Texas: net baskets or "purses" on a rod foundation from an Infiernillo context are like those in the Mayran complex of Coahuila; cotton, possibly from the Flacco and definitely from the Guerro, has also been found in the Mayran complex; the quantitative dominance of split-stitch and bundle foundation in coiled basketry is relatively late, as in the Coahuila complex. It should also be noted that the common bean and possibly moschata squash were found in the Guerro assemblage, from which also came Bat Cave type of maize. For northern Tamaulipas and adjacent southernmost Texas, MacNeish set up another series of cultural groupings. One phase he identifies with the Nogales phase of the Sierra de Tamaulipas; five others he calls "complexes" because of their relatively uncertain character. Again I feel that this separation and naming of congeries, on the basis of so little real difference in their inventories and in view of the small collections and the nature of the sites from which they were collected, is highly suspect. This is even more so when MacNeish begins to look upon them and to treat them as culture-historical realities, even actual sociopolitical units, talking of them, for example, as "the Catan people" (1958, p. 183). Eleven out of 13 Nogales types are also found in the succeeding Repelo complex, i.e., only two "Nogales" types (consisting of ten specimens) are not in both assemblages. Further, when it realized that there are only 61 specimens from all the Nogales phase sites in this northern region, the difference seems too small and the possibility that purely sampling factors are responsible 90

seems too great to justify the separation, at least at present. MacNeish says that the early assemblages are inland and represent hunting-andcollecting peoples in small to large groups having an essentially nomadic habit. The later people occupied the coastal zone, often camping on the dunes and relying considerably more on seafood. On a number of these later sites, pottery of Huastec type has been found, suggesting either that the people themselves were of southern affiliations or that there was trade between the less elaborated cultures of the north and their more developed neighbors to the south. MacNeish infers that the later population was small and divided into small nomadic bands which had widespread contacts with other groups in Texas and along the Rio Grande for an appreciable distance upstream from its delta. Late in 1945, J. T. Hughes made a "highway survey" along the road between Matamoros and Ciudad Victoria, Tamaulipas (J. T. Hughes, 1947). He discovered 11 sites and made lithic collections which are now at the University of Texas. MacNeish, in his report on the Sierra de Tamaulipas, used Hughes' material. South of Reynosa, Tamaulipas, Antonieta Espejo, of the Instituto Nacional de Antropología e Historia, Mexico, has found a series of sites at the juncture of the Conchos (Tamaulipas) and Lorenzo rivers (personal communication, 1961). One of these produced true Langtry and Shumla points. Another site yielded many large, boldly chipped stone blades typical of the Coastal Plain complex of Coahuila and the material from the Falcon reservoir. In 1950 and again in 1952, Luis Aveleyra, of the same institution, made surface collections on the Mexican side of the Falcon reservoir in Tamaulipas downstream from Nuevo Laredo (Aveleyra, 1951; Rubín de la Borbolla and Aveleyra, 1953). More or less at the same time, on the United States

ARCHAIC CULTURES: NORTHEASTERN FRONTIERS

side of the river, archaeologists from the U.S. National Park Service and the University of Texas excavated several sites (Cason, 1952; Krieger and Hughes, 1950). The material is characterized by rather large, boldly chipped blades and triangular points. There are also points very much resembling Fragua, Refugio, Abasolo, Almagre, Langtry, and Catan, and there are a few small points after the fashion of Brownsville and Starr types (Suhm and Krieger, 1954; MacNeish, 1958, fig. 50). There is a radiocarbon date of about 2700 B.C. for the Falcon focus (Suhm and Krieger, 1954, p. 565). CULTURE AND CULTURE-SEQUENCE IN TEXAS

A thorough review of the archaeological culture-history of Texas by Jelks, Davis, and Sturgis (1960) contains an excellent and lengthy bibliography classified by region, county, culture, and topic. Here I shall summarize this material, with other data from northeastern Mexico, to point up cultural and chronological interrelationships. A study of the inventories of the various classificatory units established for the Archaic of Texas shows that a certain basic way of life was typical over a wide area and a long period of time and was also related to a similar pattern to the south in Mexico. In earlier publications (1956, 1961), I have said that I believed these manifestations to be derived from a single, basic culture, specifically from some form of the Desert culture. More recent study has further indicated that, within the area embracing Texas and north Mexico, there were at least two phrasings of this basic culture, both of which quite certainly stemmed from the common ancestor but each of which, by the period of our first dependable information, had gone its own way to such an extent as to develop what were essentially two second-level basic cultures. One of these two phrasings has been named the Balcones phase (Kelley, 1947, p. 99; 1959). This includes the Edwards

Plateau aspect, the Aransas focus, and the unpublished Morhiss focus (Kelley, personal communication). Later manifestations, both on the Gulf coast and inland, seem to be descendants and mixtures of these earlier cultures and, since they are certainly not Mesoamerican, must be considered as pertaining to the Archaic cultures being discussed here. The other phrasing has a provenience largely Mexican. I am calling it the Frontera phase. The recognition of this second-level basic culture seems warranted because a quantitatively significant number of traits characteristic of the Balcones phase are not present, or only rarely so, in the inventories of Frontera phase sites. In the latter, for example, although there is much fire-rock in both sheltered and open sites, there are no great burnt-rock mounds such as are typical of the Edwards Plateau aspect of the Balcones phase in Texas. Basin metates are present only in the later Frontera material and are considered the result of stimulus diffusion from the north, since they appear in company with other northern traits. The typical Balcones phase projectile points— Pedernales, Taylor, Baird, Nolan, Montell, Frio, Bulverde, and others (Kelley, 1947, p. 104; 1959, fig. 2, legend; Suhm and Krieger, 1954, p. 108)—are not found at all in Coahuila, or only in extremely rare instances quite certainly as intrusives. There is a complete absence or a very noticeable scarcity of such common Balcones implements as stone drills, hand axes, choppers, gravers, Clear Fork gouges, off-center stem knives, picks, boatstone, egg-shaped stone pipes, engraved stone tablets, net sinkers, hones, projectile shaft abraders, antler sockets, shell "hoes," and stone-lined pit burials. Balanced against these negative indications, however, are many positive and basic resemblances: the subsistence economy was undoubtedly very similar or identical, occupation patterns were the same and, by inference, socio-political organization must 91

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

have been quite similar, basic implements were the same (their differences largely in stylistic variation), methods and materials of manufacture were the same, and the progressive developments of technology ran parallel in the two phrasings. If we had more perishable materials from Balcones sites, the similarities would probably be even more pronounced. In this scheme of two generically related but specifically distinct "basic" cultural traditions, the Big Bend aspect (Suhm and Krieger, 1954, p. 52), the Monte aspect (Kelley, 1947, p. 104, note 26; 1959, p. 285), and MacNeish's Tamaulipas material present problems, not serious but requiring mention. The Pecos River focus, the earliest in the Big Bend aspect, is obviously a very specialized and localized development (Epstein, 1960a5 p. 140; W. W. Taylor and González Rul, 1960). It has many unique traits and many others characteristic of both the Balcones and Frontera phases. Without much more information and analysis, there is little hope of resolving this uncertainty. The subsequent Chisos focus presents somewhat the same problem, although in its later assemblages the outside influences are both obvious and strong, especially in the appearance of domestic plants. At present, I would place the Pecos River focus as an extreme and geographically very restricted variant of the Balcones phase and the Chisos focus as a variant of the Hueco phase, obviously another Desert culture but with affiliations to the north (Lehmer, 1960, p. 127; Suhm and Krieger, 1954, p. 31). MacNeish's material from Tamaulipas and the Monte aspect along the Rio Grande (Falcon and Mier foci) also seem to have a basis in the Desert culture, but to which of its two local phrasings they should be attributed is uncertain. Despite an individualized stone industry, particularly a great emphasis on large, boldly chipped blades and medium to large triangular points (characteristics not found in early Coahuila complexes), they appear to be more closely 92

related to the Frontera phase, mostly because they lack the common types of the Balcones phase and because those central Texas types which they do have are obviously imports. Furthermore, they do not have the burnt-rock mounds. On a quantitative basis, both in type and frequency, their cultural inventory is much more like Frontera than Balcones assemblages. Our scant information on similarities in perishable materials points in the same direction. Finally, despite the fact that early assemblages in Coahuila lack the large, boldly chipped blades and the triangular points, these types do appear in the Coahuila as a unit in the Mayran and in the Coastal Plain complexes. There is no such "package deal" appearance in Texas. A strong tradition of cultural interchange between Tamaulipas and Coahuila seems to have existed both early (from west to east?) and late (from east to west?), but not so strong a one between the Balcones phase cultures and the Frontera phase cultures of either Coahuila or Tamaulipas. However, the highly tentative nature of these assignments must be emphasized. CULTURE AND CULTURE-SEQUENCE IN PERIPHERAL AREAS IN MEXICO

For the state of Nuevo Leon, we have only the work of J. F. Epstein, of the University of Texas, who conducted a survey in the northern part of the state in the summer of 1960. His findings have not been published but are known through personal communication and a preliminary, mimeographed report. The material has strong similarities with that of the Monte aspect to the north and east and with the Coastal Plain complex of Coahuila. After a little more work in eastern Coahuila, Nuevo Leon, and northern Tamaulipas it will probably be shown that the Coastal Plain complex of Coahuila, most of Epstein's material, and that found by Espejo south of Reynosa all belong in the Monte aspect, along with several of MacNeish's assemblages from inland, north-

ARCHAIC CULTURES: NORTHEASTERN FRONTIERS

ern Tamaulipas. In addition, Epstein found interesting stonework definitely paleolithic in technique and appearance. Like all his finds in Nuevo Leon, these artifacts were strictly surface material, and so chronological placement cannot yet be made, nor can cultural interrelationships be suggested because of the uniqueness of the types. In San Luis Potosi, a considerable amount of work has been done by nonprofessionals and some survey by Beatriz Braniff, of the Instituto Nacional de Antropología e Historia, Mexico. There has been no professional publication of materials pertaining to the Archaic. The collections which I have seen and which consist entirely of stonework resemble those of the Coahuila and Jora complexes. One notable feature is the large number of notched snub-nosed flake scrapers, slightly variant but generically like the notched and unnotched ones from the Jora and Mayran complexes in Coahuila. In Zacatecas and Durango, J. A. Mason has done survey and a little test excavation (1936, 1937). Some of his material from the upper Nasas drainage, including Zape, is the same as some from the Mayran and middle to late Coahuila complexes, specifically sewed sandals, scored sotol (?) buttons, contracting-stem and multiplenotch stone projectile points, marine worm tube beads, El Paso Brown (?) pottery. J. C. Kelley has also done survey, but only incidentally, in sites that could be associated with the Archaic. In eastern Chihuahua, Paul Reiter conducted a field school of the University of New Mexico in 1947, locating sites with Archaic material on ancient lake terraces in the vicinity of Jimenez. Considerable numbers of contracting-stem points were similar to, but somewhat variant from, those of the early Coahuila complex; other types were certainly the same generic style and period as the middle and late Coahuila complex points having stems and notches. Along the Rio Conchos (Chihuahua) from

the Junta de los Rios upstream almost to Chihuahua City, J. C. Kelley, during the course of fieldwork directed toward later sites, found Archaic materials in surface collections. Although this material has not been studied in detail, it seems to bear resemblances to material from Trans-Pecos Texas and northern and northwestern Coahuila. Kirk Bryan also found this type of stonework in western Coahuila and on the Coahuila-Chihuahua border near the mining town of Sierra Mojada, Coahuila. Ethnohistory and Language Extended, narrative accounts of aboriginal cultures in this area during historic times have been published by Beals (1932), Alessio Robles (1938), Martínez del Río (1954), and MacNeish (1958). Little new material can be added to the picture. The way of life described in the archival sources is very similar, sometimes identical, to that which can be inferred from the archaeological record as far back as we have information. This means that for the last 10,000 years at least, there was little or no significant culture change and that an essentially Archaic way of life, established in arid North America during the first years of human occupation, endured with no major modification until it was destroyed by pressures of an alien force. Recently (1961) I discussed my reasons for believing that there was linguistic as well as other cultural continuity in northeastern Mexico and southern Texas. Spanish archival sources give evidence that throughout much of this area dialects of Coahuiltecan were spoken by a numberless multitude of small groups. This language belongs to the Hokaltecan group of Hokan-Siouan. Until recently (Swadesh, 1959) Tamaulipecan was also said to belong to Hokaltecan (F. Johnson, 1940)—and this would make considerable cultural sense—but Swadesh's attribution of Tamaulipecan to Utaztecan may not be completely off, because there is evidence (Martínez del Río, 1954) that at 93

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

least some of the Laguneros or Irritilas of southern Coahuila and quite certainly the Zacatecos, who occupied that area in historic times, spoke Utaztecan. It is also probable that the Utaztecan, Tepehuan, and Tarahumar lived much closer to the Laguna District and thus to the desert

peoples of northeastern Mexico than they do today. Taken with the recognized continuity in other aspects of culture in northeastern Mexico and southern Texas, these data suggest that the people of the prehistoric cultures also spoke Hokaltecan and Utaztecan.

REFERENCES Alegre, 1841 Alessio Robles, 1927, 1938 Archaeological Salvage . . . Texas Arlegui, 1737 Aveleyra, 1951 , Maldonado-Koerdell, and Martínez del Río, 1956 Barragán, Cárdenas, and Valdés, 1960 Beals, 1932 Bosque, 1675 Boyd, 1950, 1959 and Boyd, 1937 Campbell, 1958 Cano, 1873 Cason, 1952 Cosgrove, 1947 Crane, 1956 and Griffin, 1958a, 1958b DiPeso, 1951, 1956 Epstein, 1960a, 1960b Espinosa, I. F., 1746 Gable, 1950 García Torres, 1856 Gilmore, 1947 Gladwin, Haury, Sayles, and Gladwin, 1937 Goldman, 1951 González, 1885 Hackett, 1923-37 Haury, 1936b Heizer, 1942 Hooton, 1930 Hughes, J. T., 1947 Jelks, Davis, and Sturgis, 1960 Jennings, 1956

94

and Neumann, 1940 Johnson, F., 1940 Johnson, L., 1960 Kellar, 1955 Kelley, 1947, 1955, 1959 Krieger and Hughes, 1950 Lehmer, 1948, 1960 León, 1909 MacNeish, 1958, 1960 McVaugh, 1956 Mangelsdorf, MacNeish, and Galinat, 1956 Martínez del Río, 1954 Mason, 1936, 1937 Moto y Escobar, 1940 Müllerried, 1934 Pérez de Ribas, 1645 Porter, W. W., 1932 Portillo, 1886, 1897 Rubín de la Borbolla and Aveleyra, 1953 Sayles, 1936b Smith, V. J., 1938 Studley, 1884 Suhm and Krieger, 1954 Swadesh, 1959 Tatum, 1931 Taylor, Η. C., 1948 Taylor, W. P., McDougall, Presnall, and Schmidt, n.d. Taylor, W. W., 1948, 1956, 1961, 1964 and Boyd, 1943 and González Rul, 1960 Woodbury and Woodbury, 1935 Zingg, 1940

5. Mesoamerica and the Southwestern United States

J. CHARLES

HE cultures of the American SouthT west are geographically peripheral

to those of Mesoamerica—and are very nearly geographically contiguous with them. The Southwestern cultures have the general appearance of attenuated Mesoamerican cultures, and the distributional evidence suggests strongly that they are peripheral and reduced copies of Mesoamerican prototypes. Archaeological evidence indicates, however, that they have a respectable antiquity, evolution in situ, and perhaps even independent origins. This enigma has puzzled students of American prehistory for almost a century. Among Southwestern traits especially Mesoamerican in character are: (1) an agricultural economy based on maize-beanssquash agriculture—plus chili and cotton— and, locally, irrigation; (2) permanent houses and villages, with stone and adobe masonry, conventionalized village plans, plazas, and specialized religious structures including platform mounds, kivas, and ball courts; (3) highly developed technology and artistry in stone, bone, shell, ceramics, and textiles; (4) religious art in murals, ceramics, and weaving; (5) highly organ-

KELLEY

ized socio-political structures emphasizing village hegemony and dual religious and secular leadership; (6) an organized priesthood; (7) well-developed ceremonialism involving curing societies, fertility cults, hunting and war cults, astronomical and nature deities, rain ceremonials, masked dances, god impersonation, horned or feathered serpent deities, with associated bird and amphibian representations, astronomical-ceremonial concepts, directional color symbolism, and an organized utilitarian and ceremonial calendar, culture heroes with dual aspects (such as twin war gods), sun worship, new fire and harvest ceremonials, scalp ceremonials, and possible vestiges of human sacrifice. There also are highly specific Mesoamerican elements such as copper bells, mosaic mirrors, and conch-shell trumpets. Regardless of the weight of this evidence, Southwestern archaeologists have tended to explain most of the Southwestern developments in terms of local developments. Thus, considerable evidence for multiple, long, regional, evolutionary, and developmental traditions in ceramics, architecture, decorative art, and ceremonialism has been accumu95

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

lated. The Anasazi plaza and kiva have been explained entirely in terms of local developments; Morris (1927) even suggested the independent local invention of Anasazi ceramics. Nevertheless, certain elements in Southwestern archaeological cultures, especially copper bells and mosaic mirrors, have been recognized as essentially identical with similar items in Mesoamerica. There must have been, consequently, some interchange between the Southwest and Mesoamerica. The problem became one of ascertaining the amount and direction of influence, the sources and corridors of movement, and the temporal correlation of Southwestern and Mesoamerican cultural sequences. The fundamental question was the mechanism of contact between these cultures. At the beginning of work on Southwestern archaeology and ethnology, the existence of strong Mesoamerican influence on Southwestern cultural development appears to have been accepted as a matter of course. Thus, Fewkes (1912) identified structures at Casa Grande as pyramids and compounds and pointed out (1893) that the Hopi Snake Dance resembled closely a recorded Central American ceremonial. But from about 1915 until well into the 1930's Southwestern archaeology largely concentrated on the Anasazi, who plainly showed gradual local developments in architecture, ceramics, and other material culture traits. There was little archaeological evidence of the elaborate religious ceremonialism of the historic Pueblo peoples in pre-Spanish times; the general tendency was to explain the Mesoamerican-oriented ceremonialism as having been introduced during the Spanish conquest of the Southwest under the influence of Catholicism or by Mexican Indians accompanying the conquerors or as nothing more than cultural convergence between Mesoamerican and Southwestern cultures (Parsons, 1928; 1939, vol. 2, chap. 8; Brew, 1940). There developed an archaeological cult more or less formally dedicated 96

to the proposition that Southwestern culture was largely an independent development, characterized by resistance to borrowing rather than by receptivity to diffusion— from Mesoamerica or anywhere else. But the Anasazi represented the Southwestern culture geographically the farthest removed from the Mesoamerican frontier, one which in all probability received its Mesoamerican elements through a Hohokam filter until quite late in its development. When intensive work began in the Hohokam area in southern Arizona in the late 1920's, the question of Mesoamerican influence had to be faced directly, for there were very clear-cut resemblances between Hohokam and Mesoamerican trait lists that could not easily be explained away as "convergence." The immediate result of the discovery—or rediscovery—of significant shared traits between Mesoamerica and the Southwest was Gladwin's utilization (1937, pp. 106-10) of differences in chronologies of the two areas to hypothesize a movement of cultural traits—and peoples—from (or through) the Southwest into Mesoamerica, a theory he later abandoned (1957). But there was also sound gain from the Hohokam work. Vaillant (1932) saw clearly some of the far-reaching resemblances between Southwestern and Mesoamerican ceramics, and other archaeologists also began to revise their thinking. At the same time archaeological work by Sauer and Brand (1932), Isabel Kelly (1938, 1939, 1945), and Ekholm (1939, 1942) in Sinaloa, and by Mason (1937) and Brand (1939) in Durango, for the first time made available data regarding the archaeological cultures of the northwestern periphery of Mesoamerica. Much of this new information was summarized by Ekholm (1940); two years later in the conclusions to his report on Guasave, he presented (1942, pp. 133-36) the first discussion of "Middle American-Southwestern Relationships" to utilize the new data. While agreeing that there was evidence of Mesoamerican influ-

MESOAMERICA & SOUTHWESTERN UNITED STATES

ence on the Southwest, Ekholm concluded that, contrary to earlier expectations, the West Coast corridor did not seem to have been the major channel, suggesting instead that the Zacatecas-Durango region was probably the proper place to search for the Mesoamerican origins of Southwestern culture. World War II temporarily ended field work in the critical area of northern Mesoamerica but studies continued. In August 1943, the Sociedad Mexicana de Antropología sponsored a "Mesa Redonda" in Mexico City in which problems concerned with "El Norte de Mexico y el Sur de Estados Unidos" were considered. Papers by Beals (1944a,b), Brand (1944), Brew (1944), Haury (1944), Kelly (1944), and Marquina (1944) summarized known ethnological and archaeological data regarding Mesoamerican-Southwestern relationships and attempted a rough correlation of the sequences of the two areas. Shortly thereafter Haury (1945b) reviewed and analyzed the entire relationship. So acute was Haury's understanding and treatment of the subject that even today, much new information does little more than expand and elucidate the propositions which he made. From the late 1940's to the present, field work in northern Mexico has been done principally in Nayarit-Zacatecas-DurangoChihuahua. With a flood of new knowledge regarding the northwestern periphery of Mesoamerica, the strength of the Mesoamerican impact upon the Southwest became apparent. In the Southwest itself interest was renewed; specifically Mesoamerican architectural features were identified in both Hohokam (Wasley, 1960) and Anasazi (Ferdon, 1955) cultures. Regional intercultural relationships were examined (Ferdon, 1955; A. S. Johnson, 1958; Kelley, 1960), and the Southwest Seminar held in Santa Fe in 1955 openly accepted the strong dependence of the Southwest, period after period, on Mesoamerican influence (Jennings, 1956).

METHODOLOGY AND BASIC CONCEPTS

Although archaeology, ethnology, linguistics and ethnobotany all contribute data supporting this viewpoint, attention here is focused on the archaeological evidence. Many elements of Puebloan religion and ceremonialism are almost certainly of Mesoamerican origin, and their entrance into the Southwest ties in with the archaeological findings. Certain north Mexican ethnic groups, such as the Huichol, which are usually accepted without question as having Mesoamerican affiliation, nevertheless show so many Southwestern characteristics that Beals (1944a) found it necessary to create the concept of a "Greater Southwest" which would incorporate such groups in the Southwest cultural-ecological area. What, specifically, is the archaeological evidence of Mesoamerican-Southwestern contacts? So far, only copper bells, which have a wide distribution in both Anasazi and Hohokam especially after A.D. 1100, are demonstrably trade items. Mosaic mirrors and shell trumpets of the Hohokam, however, are possible items, and in the future other items almost certainly will be identified as trade goods. Mosaic mirrors, shell trumpets, elaborate turquoise mosaics, and certain pottery decorative patterns and elements probably represent local copies of Mesoamerican artifacts. Architecturally, platform mounds and compounds of the Hohokam and certain intrusive elements in Anasazi architecture belong almost certainly to this category. More dilute Mesoamerican influences can be detected in style changes in ceramic decoration and in other artifact types in various Southwestern cultures, particularly the Hohokam. When such changes parallel those occurring in definite Mesoamerican cultures but are later, they can be used as evidence of sustained contacts over a period of time. Many apparent parallels between the Southwest and Mesoamerica may represent influence of the latter on the former, rather 97

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

than independent but convergent development. Thus, the inward-facing plaza of the Southwest may prove to be merely a local copy of the standard pattern of a court surrounded by platforms and houses characteristic of the cultures of the northwestern frontier of Mesoamerica. The Southwestern "Great Sanctuary" may prove to be the equivalent of the "Hall of Columns" ceremonial structures of the Mesoamerican cultures. The sipapu as a subterranean entrance of the gods from the spirit world may prove to be the equivalent of the Huichol cavity beneath the fire hearth in which an image of the god is kept—and of the Canutillo-Chalchihuites (Mesoamerican) trait of the statue of the god buried in the altar in the center of the plaza. There is also inferential evidence. Both Southwestern and Mesoamerican cosmologies include the concept of world quarters, usually with specific gods or ceremonial attributes assigned to each quarter. When this concept can be identified archaeologically in ceramic design in the northern Mesoamerican cultures—and in the surviving peoples there such as Huichol-Tepecano —and when the first painted pottery arriving in both Mogollon and Hohokam contains this design in simple form, apparently derived from the Mesoamerican ceramic complexes in question—can the inference then be made that the concept itself in all probability entered the Southwest from Mesoamerica at that time, along with the patterns and techniques for decorating pottery? A great number of such inferences can be made regarding MesoamericanSouthwestern relationships, and although these inferences are not subject to direct testing, they provide working hypotheses that may be tested in other ways as new evidence appears. The question of comparative chronologies enters strongly into the question under discussion. A chronology for the Anasazi development, based on dendrochronology, seems fairly firm, but the same cannot be 98

said for the Hohokam or even the Mogollon. Although chronologies for the latter cultures have been developed, principally through cross-dating with Anasazi periods, the chronology accepted for the Hohokam sequence by some investigators differs notably and, in terms of correlations with Mesoamerica, significantly from that used by others. The basic Mesoamerican chronology is being revised constantly and the peripheral Mesoamerican cultures of the northwestern frontier are dated inadequately in terms of weak correlation southward with the central Mesoamerican development and northward with the Southwestern developments. Here and there a few radiocarbon dates reinforce the chronologies, but not definitively so. Only rarely can comparative chronology of the sequences, especially as expressed in actual dates, aid appreciably in interpreting MesoamericanSouthwestern relationships. Indeed, the comparative validity of the various chronologies may depend on an adequate correlation arrived at through other means. It is clear, however, that the basic Mesoamerican development leads in time by a considerable margin that of the Southwest. Moreover, correlation between peripheral Mesoamerican sequences and those of the Southwest may be established on a relative basis, involving the temporal point (but not actual date) of appearance of certain elements relative to still other elements in the regional sequences. The use of this approach and of trial chronologies to be fitted against each other as correlation proceeds by other means may eventually align regional chronologies. The mechanism of contact between the Mesoamerican and Southwestern developments is important in interpreting the evidence of that contact and influence. Large group migration, small group migration, infiltration of individuals, and direct, indirect, and stimulus diffusion have been suggested as mechanisms. Although both Charles C. DiPeso and

MESOAMERICA & SOUTHWESTERN UNITED STATES

Albert H. Schroeder (Schroeder, 1956) appear to favor large group migration, its only clear-cut advocate is Harold Gladwin (1957, pp. 81-94), who derives the Hohokam as an ethnic and cultural group from an unidentified Mesoamerican source through mass migration up the west coast of Mexico. The evidence cited is general, and such a migration is not supported by the archaeological data on western Mexico. DiPeso has developed the concept of a basic (and simple) village-dwelling, agricultural, ceramic group, the Ootam, who either were modified almost overnight into the Hohokam or were locally displaced by them, and who are regarded either as a small power group—culturally if not politically—or as a fairly large migrant group (DiPeso, 1956, pp. 561-68). Later, an Ootam resurgence resulted in the relatively low-level occupation of the Hohokam area by the Pima-Papago in the historic period. If the intrusive elements were regarded as groups of traders, perhaps with military support, and the period of replacement of Ootam, through acculturation, as decades or centuries rather than overnight, this hypothesis would be fairly close to the one I have developed here. But because the Mesoamerican elements introduced into the Ootam (which changed it into the Hohokam) arrived in clusters over some range of time rather than all at once, DiPeso's thesis in its rawest form seems untenable. Edwin F. Ferdon, however, has presented the small-group-migration hypothesis in its purest and most nearly acceptable form (1955). He visualizes either infiltration of the organized Mesoamerican trading groups (the pochteca) into the Southwest or, alternatively (and perhaps in combination), the entrance into the region of a small group of militaristic Toltec nobles, displaced from the central Mesoamerican homeland by the same disruptive forces that destroyed Tula itself. Small group migration, especially through the entrance of traders, is in itself one kind

of diffusion but other kinds of diffusion have been suggested to account for the observed facts of Mesoamerican influence on the Southwestern cultures. Haury (1945b, p. 56) stated: "The traits which I believe to be traceable to Mexico were borne by all methods of cultural transmission, from the diffusion of ideas from group to group to the importation of actual objects probably by traders." Ekholm (1942, p. 136) noted that the later contacts especially "occurred across great areas that were occupied by relatively primitive peoples, and it seems likely that it must have been accomplished in large part by small groups of travelers." Investigations completed during the last decade in northern Mexico, taken with linguistic and ethnographic data, suggest that in preceramic times there was a fairly constant group of food-gathering peoples, essentially Desert culture in affiliation, spread over both northern Mexico and the Southwest. Somewhat later, perhaps by the time of Christ, if not before, there developed along the Sierra Madre Occidental a chain of related cultures and peoples, characteristically simple agriculturists, living in small fairly permanent hamlets, making simple pottery vessels, and in all probability speaking related dialects of UtoAztecan. One of these groups, the Loma San Gabriel culture, extended from southern Durango well into Chihuahua along the Sierra corridor. It existed there throughout the Chalchihuites occupancy of Durango and probably is represented in remnant form today by the Tepehuan people and culture. On archaeological grounds it seems probable that the earliest pottery of the Chalchihuites culture, dating back perhaps as early as A.D. 300, was traded northward along this corridor at least as far as northern Durango—and with the pottery almost certainly all sorts of ideas associated with it and with the culture of its makers. Here was a ready-made cultural corridor along which trade items, descriptions, ideas, and wandering travelers must have moved with 99

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

ease. The end of this corridor lies on the very doorstep of that part of the Southwest in which both Hohokam and Mogollon developed. PRINCIPAL LOCAL OR REGIONAL CULTURES INVOLVED IN MESOAMERICAN-SOUTHWESTERN CONTACTS

What Mesoamerican cultures existed on the northern frontier and which of these may have been the sources of influence on the Southwest at various times? Although much territory is still archaeologically unknown, we do know that the Preclassic northwestern Mesoamerican frontier did not extend far beyond the southern boundaries of Jalisco-Guanajuato. The Chupicuaro culture on the Michoacan-Guanajuato border is the best known of these, representing a very well developed Late Preclassic and Early Classic culture showing strong ties with the Preclassic of the Valley of Mexico (Μ. Ν. Porter, 1956). Ceramics of this culture were apparently diffused far northward into Zacatecas, and the decorative patterns of these traded ceramics almost certainly strongly affected ceramic developments in Zacatecas-Durango, perhaps also as far away as the American Southwest. Classic cultures are known from Colima and Nayarit. If current interpretations are correct, the Chametla culture of southern Sinaloa and the initial stages at least of the Malpaso-Canutillo (La Quemada) and Chalchihuites cultures also belong on a Late Classic horizon. In Durango the later phases of the Chalchihuites culture—Las Joyas through Calera phases—are certainly Early Postclassic in age, as is the Aztatlan horizon in Nayarit-Sinaloa. The Culiacan culture in northern Sinaloa, and related local cultures to the south, continued throughout the Late Postclassic until terminated by the Spanish conquest. Durango appears to have been entirely deserted by Mesoamerican cultures after about A.D. 1350, as was most of Zacatecas, except that in southern Zacatecas and adjacent Jalisco 100

a Mesoamerican culture is known to have survived into conquest times, perhaps as a lineal descendant of the Juchipila-Bolanos cultures known archaeologically from this area and extending back at least to the beginnings of the Canutillo-Chalchihuites cultures. North of the Mesoamerican frontier proper in Durango-Chihuahua and Sonora are the Loma San Gabriel and Huatabampo cultures, respectively, which appear to be intermediate between Southwestern and Mesoamerican patterns. In Southwestern United States the Hohokam was the principal affected culture, although the Mogollon culture was certainly in early contact with Mesoamerica. The Anasazi culture in its earliest periods received Mesoamerican influences primarily through a Hohokam screen, although later, as it expanded into Mogollon and Hohokam territory, it too came to confront directly the Mesoamerican periphery. The Chihuahua culture, long thought to be basically Southwestern with roots in both Anasazi and Mogollon cultures, represents a local development in northwestern Chihuahua. DiPeso's recent large-scale excavations there have shown that in late times this culture came under very strong Mesoamerican influences, the source of which cannot be pinpointed at the moment. The Patayan culture of the Southwest has not been studied in terms of Mesoamerican contacts and will not be given attention here. SUMMARY OF APPARENT MESOAMERICANSOUTHWESTERN CONTACTS, WITH REFERENCE Tο SPECIFIC CULTURES AND PERIODS

This discussion must be viewed for the present as a hypothesis, but one which can be substantiated at many points. The principal outline appears to be well established, leaving details, times, and sources of contacts to be questioned, as well as certain inferential constructs regarding religion and ceremonialism. Basic Southwestern agricultural products and the technology for their production and

MESOAMERICA & SOUTHWESTERN UNITED STATES

utilization were trade-diffused to the Southwest into and through a Desert culture matrix, probably by 2000 B.C., if not earlier. These must have been foodstuffs which required only a short growing season or had other characteristics which made their cultivation possible in new physical environments by peoples previously unskilled in agriculture. Only slowly would such crops come to be of prime importance, serving at first only as auxiliary food supplies in a hunting-gathering economy (see Jennings, 1956, pp. 73-78, 108-10). Perhaps certain simple ceremonials and religious concepts were diffused into the Southwest with the first agricultural know-how, but if so, they cannot be identified with certainty now. Around the beginning of the Christian Era, perhaps a few centuries earlier, certain small groups in southern New Mexico and Arizona came to depend on simple agriculture as a way of life and therefore to be relatively permanent village—or hamlet— dwellers. Coincidentally, the Preclassic frontier of Mesoamerica had moved northward into southern Jalisco-Guanajuato, and emanations from it were probably felt as far as this southern part of the Southwest. Simple ceramic forms in plain wares and polished red wares appeared early in the occupation of the earliest Southwestern villages and almost certainly must be attributed to a Mesoamerican source in the advancing Late Preclassic and Early Classic cultures to the south. It is tempting to see in the intermediate Loma San Gabriel and Huatabampo cultures the immediate source for these innovations, but proof is lacking that these cultures extend far enough back in time to have so served. Clay figurines appeared with the first pottery (Gladwin, Haury, Sayles, and Gladwin, 1937, pp. 23645, 251-53) and probably indicate that certain religious traits have been passed to the Southwest along with ceramics. Not long after the introduction of pottery itself, perhaps by A.D. 100, decorated pottery with virtually identical designs and

layout patterns appeared in both early Mogollon (Georgetown phase) and early Hohokam (Estrella phase) village sites (Gladwin, Haury, Sayles, and Gladwin, 1937, pp. 199-202, 253-54; Jennings, 1956, pp. 81-82). These oldest Southwestern painted wares, San Lorenzo Red-on-Brown and Estrella Red-on-Grey, are simple hemispherical bowls with quadrant interior decoration in broad (and some narrow) lines and nested triangles. Essentially the design layout consists of a broadline cross, or a series of narrow lines forming a cross, dividing the bowl interior into quadrants which are filled with nested triangles. Estrella Redon-Grey also has four sets of nested triangles pendent from the rim exterior. This same design on bowl interiors appears in the Valley of Mexico at Tlatilco in the Middle Preclassic (Piña Chan, 1958, fig. 43,i) and in modified form at Ticoman in the Late Preclassic (Vaillant, 1931, pl. 71,f,i). At Chupicuaro, on the MichoacanGuanajuato border, the same design occurs in Chupicuaro Red-on-Buff Polychrome of the Transitional period (Μ. Ν. Porter, 1956, p. 549, fig. ll,p,q3t), complete with pendent triangles on the exterior. These, however, are tripod bowls; only three of the exterior triangles, placed between the tripod leg bases, are surrounded by an elaborate design of their own which does not appear in the Southwestern wares. The Chupicuaro bowls with this design layout apparently are comparatively rare; they probably date approximately between 100 B.C. and A.D. 100, highly significant dates in terms of the introduction of this layout pattern into the Southwest at about A.D. 100. Batres (1903, pl. 27) illustrated two bowls purportedly from La Quemada, far to the north in Zacatecas, which Μ. Ν. Porter (1956, p. 574) identifies as "without question of Chupicuaro origin." One of these is a specimen of Chupicuaro Red-onBuff Polychrome with the quadrant design discussed above. The rarity of this type at 101

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

Chupicuaro makes its presence at La Quemada, if indeed these specimens did come from there, all the more interesting. More to the point, however, is the fact that a whole series of decorated wares of the early periods of the Malpaso-Canutillo and Chalchihuites cultures in Zacatecas (including La Quemada) and Durango utilize this same motif as the principal layout for the interior decoration of painted bowls. In this series of wares, however, the placement of god representations and astronomical symbols in the layout quadrants indicates quite clearly that this design was an indicator of the world quarters with their respective gods. This concept survives in the symbolism of the Huichol, where the concept as depicted on stone discs and ceremonial basket-plaques is explicitly stated as outlined above (Lumholtz, 1900, pp. 31-32, 166-67, pl. 11, fig. 2). Chupicuaro, or some northern outpost of this culture in Jalisco or Zacatecas, was the source of the first painted wares introduced into the Hohokam and Mogollon, and in view of the ceremonial significance of the pottery layout in northern Mesoamerica and in historic times in the Southwest it may well be that associated concepts of cosmology and ceremonialism were introduced at the same time. Also of interest is the occurrence of slant eyes and "turban"-style hairdress on some figurines of the Estrella phase (Gladwin, Haury, Sayles, and Gladwin, 1937, pl. 204,g,k), for these are characteristics which also appear on Chupicuaro figurines. DEVELOPMENT OF THE HOHOKAM CULTURE, THROUGH MESOAMERICAN ACCULTURATION OF THE OOTAM

Perhaps as early as A.D. 300 or as late as A.D. 500, Mesoamerican cultures were established in western Zacatecas and southern Durango and in southern Sinaloa. There is little indication that the Chametla culture of southern Sinaloa left any specific mark on the Southwest, but there is ample evi102

dence that the Chalchihuites and Canutillo cultures in western Zacatecas and southern Durango strongly affected the late Pioneer, Colonial, and Sedentary periods of the Southwestern Hohokam. Hohokam architecture and village plan, ceremonialism, ceramics, and other artifacts all show strong and incremental influences from Mesoamerica during this time, many, if not all, of which can be traced to a specific Chalchihuites or Canutillo culture source. By the beginning of the Classic Hohokam, when Chalchihuites influence began to wane, the Hohokam had accepted so many Mesoamerican traits that we can accurately speak of the Mesoamerican acculturation of the Hohokam. It is known that the relatively highly evolved Pima-Papago maize was in use in the Chalchihuites culture when it began to affect the Hohokam. It may be that the transmission of this advanced maize to the Hohokam was accompanied by the appropriate ceremonial techniques for growing it. The evolved maize requires more care in planting, cultivating, and harvesting, with especial attention to time of planting because of the much longer growing season needed. The Chalchihuites geometric designs and life forms on Hohokam pottery in the late Pioneer and early Colonial periods are traits which appear on Chalchihuites vessels in a ceremonial context. Chalchihuites ceramic design elements and layout patterns appeared first in the Hohokam (where they were reworked in local style), followed much later by Chalchihuites vessel forms, crudely copied and never popular in Hohokam. Among the ceramic traits apparently transmitted from the Canutillo-Chalchihuites cultures to the Hohokam during this interval were: preoccupation with positiveand negative-painted small animals; many small unit designs; simple and interlocking plain and fringed scrolls in great variety; serrate-border design bands; quadrant bowl interior patterns, plain, and with animals

MESOAMERICA & SOUTHWESTERN UNITED STATES

and/or geometric designs in the quarter panels; circling interior and exterior scroll patterns; rims with hatched and scroll decoration; footed vessels; basket-handled bowls. Many of the scrolls, serrate-bordered bands, positive and negative life forms, small unit designs, and crosshatching appear to have been derived as painted copies of Michilia Red-filled Engraved pottery designs of the Alta Vista phase (A.D. 300-700) of the Chalchihuites culture. Similar painted copies of both the positive and negative designs were made in the following Ayala phase, with loss of the champlevé technique in the new colonies in Durango, so it is possible that some or all of the Hohokam designs in question were derived from this secondary source between A.D. 500 and 700. In general, these ceramic traits appear earlier in Mesoamerica and are components of much richer ceremonial art complexes there, so that the direction of flow seems clearly to have been from Mesoamerica to the American Southwest. In terms of other material traits, the Hohokam shares with the peripheral Mesoamerican cultures such items as stone bowls and mortars, carved and plain; three-quarter-grooved polished stone axes (but also full-grooved axes and mauls in Chalchihuites); troughed metates; "medicine" stones; small pestles; and stone "doughnuts." Stone palettes, so well developed in the Hohokam, are not part of the peripheral Mesoamerican cultures; and although both cultures have mosaic mirrors, the types are different and, so far as now known, are confined to one period (Las Joyas phase, A.D. 700-900) in Chalchihuites. Both cultures share a preoccupation with work in shell, including specific bead and gorget types, and miniature carving in the round of shell and stone. Both share conch-shell trumpets —Chalchihuites has ceramic copies of the shell specimens also—and turquoise mosaics. Many if not all of these traits probably will be traceable to a Mesoamerican origin,

but at the moment the direction of movement of most of them is not certain. Copper does not appear in Chalchihuites until about A.D. 1000 or 1100, or at the same time it appeared in the Hohokam, but whereas only copper bells reached the Hohokam— and the Southwest in general—Chalchihuites had a considerable variety of copper artifacts such as bells, needles, awls, chains, hooks, and pendants. It cannot be demonstrated that the copper bells entering the Southwest came from Chalchihuites, but certainly they were derived from some Mesoamerican source. Architecturally, the picture is not as clear. Ball courts appeared in the Hohokam in early Colonial (Gila Butte phase) times and are regarded as Mesoamerican in origin. However, a small putative ball court at the Schroeder Site and a large I-shaped court at La Quemada are the only ball courts known from the peripheral Mesoamerican cultures; there is little resemblance between these and the Hohokam courts. However, platform mounds of refuse, often mixed with rubble or earth, are characteristic of the Canutillo-Chalchihuites cultures throughout their life spans, and a similar structure has been excavated at the Gatlin Site (Wasley, 1960), attributed to the Sedentary Hohokam, in Arizona. The Mesoamerican features were constructed by building masonry walls and filling between them with refuse, earth, or rubble, topped by a floor of cobblestones covered with an adobe surface layer, as were the walls. The Hohokam mound was surfaced and its sides were confined with a heavy layer of caliche plaster. It had been rebuilt in several stages, as the Mesoamerican structures customarily were. At the same Hohokam site copper bells occurred in relative abundance. The Gatlin Site platform mound certainly is a local copy of a common architectural feature of the peripheral Mesoamerican cultures. In view of their features and construction it seems probable that the large carefully made and oriented refuse mounds 103

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

of the large Hohokam site at Snaketown (Gladwin, Haury, Sayles, and Gladwin, 1937) actually must be regarded as platform mounds or pseudo-pyramids of Mesoamerican inspiration. Canal irrigation became important in the Hohokam in Colonial times and probably was used also in the peripheral Mesoamerican cultures, although its presence there is inferred and has not been demonstrated archaeologically. Cremation of the dead was important in the Hohokam but was uncommon or absent in the peripheral Mesoamerican cultures; it is probably an ancient trait in Hohokam, perhaps carried over from Desert culture burial practice. The specific house types of the Hohokam probably are of Southwestern rather than Mesoamerican origins. The Mesoamerican traits that were intruded upon the small Ootam villages during this period to produce the true Hohokam culture were primarily concerned either with agricultural subsistence or with ceremonialism. This is especially true with many of the pottery designs and layout patterns which derive from a rich religiousceremonial art in the Chalchihuites-Canutillo cultures—an art closely integrated with the religion and ceremonial art of Mesoamerica. Thus there are alligator-monsters; twin deities with astronomical indicators; horned bird-serpents, tigers, coyotes, turtles; world quarters with respective gods; a raised basket-handle, double-headed serpent spanning the world with accompanying sun and star symbols, eagles, dancing twin gods and xolotl or monster-twin figures passing over the span; vessel rim figures below the serpent span in which the xolotls enter a monster mouth; and other elements in the Chalchihuites-Canutillo cultures. Many of the specific life forms, the geometric forms which also represent them, the astronomical symbols, and even the basket-handle heavenly serpent appeared also in Hohokam ceramics. In Hohokam their use seems to have been primarily decorative; neverthe104

less it is probable that Mesoamerican sun, rain, and fertility cults and beliefs also were transferred to the Hohokam from Mesoamerica at this time, probably along with improved agricultural products, as a ceremonial-technical complex. It seems, therefore, that progressively through late Pioneer, Colonial, and Sedentary Hohokam periods, Hohokam agricultural technology, ceremonial architecture, ceremonial ceramic art and ceremonialism in general, and many associated artifacts were shaped into a rough copy of similar items in the peripheral Mesoamerican donor cultures to the south, so that Mesoamerican-acculturated late Sedentary Hohokam was in fact a local and crude Mesoamerican image. MESOAMERICAN DIFFUSION TO THE MOGOLLON AND ANASAZI CULTURES

Superficially at least, the Mogollon culture does not seem to have borrowed as much from Mesoamerica, after the acquisition of agriculture and basic painted pottery, as did the Hohokam. Southwestern archaeologists incline to the belief that Mogollon developments were influenced by both Hohokam and Anasazi. Thus, Haury (1936a, fig. 6, p. 21) attributed scrolls, life forms, and bowl layout patterns in Mogollon ceramics to the Hohokam and the adoption of a white background to the Anasazi. The Southwest Seminar group was impressed with the relative lack of Mesoamerican elements in Mogollon and in explanation developed the hypothesis that the early Mogollon occupied narrow mountain valleys at high altitudes and thus were not as ecologically well situated to borrow the full Mesoamerican cultural patterns as were the Hohokam. In later Mogollon times it was easier to borrow from the nearby Hohokam, already well acculturated to the Mesoamerican way of life, than from the more distant Mesoamerican peripheral cultures (Jennings, 1956, pp. 118-19). Some or all of the ceramic elements noted—specific layout patterns, scrolls, life

MESOAMERICA & SOUTHWESTERN UNITED STATES

forms, and white background—may have been borrowed by the Mogollon directly from the Canutillo-Chalchihuites cultures, rather than from the Hohokam-Anasazi. Surprisingly, there is evidence of such contact in reverse. Otinapa Red-on-White pottery of the Rio Tunal phase (A.D. 950-1150) of the Chalchihuites culture is very close in layout patterns, design elements, pigments, and white background to Three Circle Redon-White of the Mogollon culture, dated somewhat questionably as early as A.D. 800. The relative dates would suggest that the Otinapa type was introduced into the Chalchihuites culture from the Mogollon rather than the reverse. However, since it is in this very Three Circle Red-on-White ware that several of the new traits first appear in the Mogollon, this relationship needs further study. The recent discovery of a wide range of red-on-white wares in the Canutillo culture dating prior to A.D. 900 suggests that there may have been a transfer of the elements in question from these wares to the Mogollon where, following a period of redevelopment, the new ware was reintroduced into the Durango Mesoamerican colonies, in which red-onwhite wares had been lost during the Las Joyas phase. The relative paucity of Mesoamerican influence on the Mogollon may be explained in terms of the hypothesis presented earlier in this paper: that the great mass of new items which were introduced into the Hohokam—and apparently not into the Mogollon—were components of a ritualistic complex accompanying the introduction of new and improved maize. Perhaps the early Mogollon habitat had too short a growing season for these new maize types to be easily adapted there—and the associated complex failed to diffuse there also. Then, with the move of the Mogollon into broad valleys at lower elevations, the new complex, agricultural and ceremonial, may have been adopted belatedly, accounting for the sudden and surprising shift into life forms

and ceremonialism evidenced in Mimbres

pottery.

It seems probable, however, that the Mogollon were constantly in contact with the Mesoamerican outposts, as with the Hohokam, and that their failure to adopt many Mesoamerican elements early in the sequence does not mean that they were ignorant of them. Since some Southwestern archaeologists tend to see the Mogollon as the source for early influences on evolving Anasazi ceramics (Haury, 1936a, pp. 16-17, fig. 6; Jennings, 1956, p. 96), it is highly probable that Mesoamerican elements in Anasazi may have been passed in part through a Mogollon screen. Just as certainly some Mesoamerican elements must have been diffused to developing Anasazi through a Hohokam screen. The distribution of design elements in Anasazi pottery suggests such an explanation, and even more impressive is the discovery by Brew (1946, fig. 101,w) at Alkalai Ridge, Utah, of a unique negative-painted sherd in an early Pueblo I context dated in the 8th century A.D. The design in negative dots and bands on this sherd resembles intrusive negative-painted, dot-decorated sherds found in the Mesoamerican Canutillo culture on a putative A.D. 350-750 horizon. The negative-painted technique demonstrated in this unique specimen is evidence of Mesoamerican influence on early Anasazi. Basic elements of Mesoamerican ceremonialism must also have been diffused to Anasazi during this period through a Hohokam-Mogollon screen. But the studies which would clarify and demonstrate such influences, either in material culture, technology, or ceremonialism have yet to be made or published. From a much later period in the Anasazi development there is ample and specific evidence of Mesoamerican influence (Jennings, 1956, pp. 96-97). On a general late Pueblo II-early Pueblo III horizon, A.D. 1000-1100, and afterward, copper bells of Mesoamerican origin appeared at various localities in the Southwest, including such 105

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

Anasazi strongholds as Pueblo Bonito (Pepper, 1920, pp. 269, 324-25; Judd, 1954, pp. 109, 715, fig. 28) and Aztec (Morris, 1919, p. 100) in New Mexico and in the Mimbres (Bradfield, 1931, p. 124) as well as in the Hohokam area. In addition, cloisonne-decorated pieces of sandstone (mirror backs?) and conch-shell trumpets found at Pueblo Bonito (Pepper, 1920, pp. 53, 177, figs. 13, 77,a; Judd, 1954, pp. 305-06, pl. 82,b) are probably of Mesoamerican origin, illustrating, with the bells, the close relationship of the trade items to ceremonialism. Ferdon (1955) also has called attention to the existence of Mesoamerican architectural features in Chaco Canyon: core masonry, a platform mound with forecourt and stairway with balustrades, a square-column fronted gallery, and circular towers. Examining more closely the material culture of Pueblo Bonito, one finds rock-cut stairways, parrots, a pottery stamp, cylindrical jars, animal and human effigy jars, elaborately painted wooden slabs, and elaborate work in turquoise mosaic including bone inlays (Pepper, 1906; 1920, figs. 42, 43, 45, 65, 71, 86, 101, pls. 2-8; Judd, 1954, figs. 54, 63, 64, pls. 6, 36, 37, 67, 68, 76). The cut-out multipart bird figures found in the Chaco sites certainly belong in this group. Although some of these traits may prove to be local developments, their florescence coincidentally with the appearance of known Mesoamerican traits in Chaco Canyon suggests that most of them may be of Mesoamerican origin. Ferdon (1955) believes that the Mesoamerican architectural features in Chaco Canyon may have been introduced by traveling traders or perhaps by displaced Toltec nobles, in company perhaps with the Quetzalcoatl cult, around A.D. 1050. Although most of the architectural features noted, plus many of the other elements, were known in the Chalchihuites-Canutillo cultures, there is no immediate proof that these cultures survived to such a late date, 106

except in the Guadiana Valley region in central Durango. There in the late Rio Tunal and Calera phases, the Mesoamerican colonies were reduced and unimportant, and it seems probable that the influences must have come from elsewhere, perhaps in part from the Mesoamerican-influenced site of Casas Grandes, recently excavated by Charles DiPeso, or from late Mesoamerican developments in northern Sinaloa. As a guess, Casas Grandes may prove to be the origin of the architectural features and perhaps other items as well, but determination of this point must await DiPeso's studies. However, almost certainly the late developments in the Aztatlan horizon at Guasave in Sinaloa played a major part in somewhat later developments in the Southwest and perhaps in those just discussed as well. PROBABLE LATE MOVEMENT OF A REINVIGORATED

MESOAMERICAN

CEREMONIALISM

UP THE WEST COAST OF NORTHERN MEXICO AND INTO THE AMERICAN SOUTHWEST

Excluding events at Casas Grandes, two principal developments in the culture history of northern Mesoamerica and the Southwest should be noted. In the Southwest itself the principal event was the expansion of the Anasazi southward and westward until by A.D. 1100-1200 they shared a common frontier with the Hohokam and then, according to some Southwestern archaeologists, either conquered the Hohokam or moved in to live with them for more than a century (Haury, 1945a, pp. 204-13; Reed, 1951, pp. 436-37). Whether the Puebloan intruders in Hohokam territory were Sinagua or Salado in origin matters not in this context. What is important here is that this move, together with related Puebloan movements into the Mogollon-Mimbres and into the lower Rio Grande Valley in New Mexico and adjacent Texas (in unholy union with the Mogollon) and Chihuahua, for the first time brought

MESOAMERICA & SOUTHWESTERN UNITED STATES

the Anasazi on a broad front into direct confrontation with the northern periphery of Mesoamerica. The second principal event took place in the Nayarit-Sinaloa coastal strip of northwestern Mexico. There the Aztatlan horizon, beginning perhaps as early as A.D. 750, had by about A.D. 1000 become more or less a pan-Sinaloa culture maintaining surprisingly direct contacts back to Early Postclassic, primarily early Aztec, cultures of the Valley of Mexico. Certainly, by about A.D. 1100— just when Anasazi southwestward expansion brought them virtually to the United States-Mexican boundary region—there was established at Guasave on the Sinaloa River in northern Sinaloa, almost on the Sonora border, a florescent late Aztatlan horizon culture, the Guasave phase (Ekholm, 1942; Kelley and Winters, 1960). Ekholm, who excavated the Guasave site, reported the existence there of many elements of the central Mexican "MixtecaPuebla" cult (1942, pp. 125-32). So strong was this Mixteca-Puebla contribution to the Guasave culture that Ekholm suggested that it must have been introduced by an actual migration of people from central Mexico, not earlier than A.D. 1300 and more probably around A.D. 1350. The central Mesoamerican element in the Guasave phase is strong indeed. To judge from ceramic and mortuary evidence it was exemplified primarily in ceremonialism and ceremonial art and was intimately associated with the northward spread of copper metallurgy, or at least copper artifacts. Ekholm (1958, chart) apparently now favors dating the introduction of this complex into Sinaloa at shortly after A.D. 900. This appears to be more nearly correct; some of the bowl interior decorations in "MixtecaPuebla" style which Ekholm illustrates (i.e., figs. 7,a; 9,b; 6,f,g) appear to be stylized serpent heads closely related to Aztec I specimens from the Valley of Mexico (Covarrubias, 1957, p. 324), and Aztec I

must have been well developed prior to 1200. Regardless of the correctness of the migration hypothesis, it is certain that a strong wave of ceremonialism and ceremonial art—illustrated by elaborate stylized polychrome paintings of serpents, birds, individuals in feather headdress—moved up the west coast of Mexico to the southern boundary of Sonora, almost face to face with the Hohokam, around A.D. 1100-1200. A number of Southwestern artifacts occurring at Guasave (shaft polishers, bilobed beads, general shellwork, perhaps a stone pipe) suggest a composite Hohokam-Anasazi origin; Mesoamerican traits (such as bilobed spindle whorls, overhanging-end manos) appearing in the Classic Hohokam, in the Desert Hohokam Topawa and Sells phases (Haury, 1950, pp. 2-14), suggest a late west coast source or the Guasave phase as their point of origin. The architectural features that are putatively Mesoamerican and new in Hohokam in the Classic period (compounds with house mounds and "great houses") cannot be attributed, with present knowledge, to a Guasave source, since no architectural forms are known for the Aztatlan horizon, or any other Sinaloa culture for that matter. Perhaps they go back to the Mesoamerican influences that shaped Casas Grandes in Chihuahua; the question is an open one. Similarly, it is tempting to an outsider to see the origin of the new polychrome wares of the Salado-Hohokam (Pinto, Tonto, Gila polychromes) in the florescence of polychrome wares in the Guasave phase, but Southwestern archaeologists believe that these have local developmental histories. In effect, it is difficult to demonstrate much strong influence of Guasave on Hohokam— or on Salado. But it is now known that most of Southwestern ceremonialism—or ceremonial art —first appears archaeologically at the very end of Pueblo III or the very beginning of A.D.

107

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

Pueblo IV (about A.D. 1300, or the very culmination of the Guasave ceremonial florescence) and that it has a decidedly Mesoamerican cast. Brew (1944, pp. 24245) has described a complex which includes Jeddito Black-on-Yellow and Sikyatki Polychrome pottery of the Hopi country; many new pottery decorative techniques, colors, design motifs and layout patterns, including life forms and mythological beings; ceremonial mural art, with birds, masked dancers, mythological beings; and other elements which spread rapidly through the Anasazi pueblos at the beginning of Pueblo IV. He suggests, following Beals in part, that the Kachina cult may have been introduced from Mesoamerica at this time as a variant of the Tlaloc cult. This seems very probable, but it is far from the complete picture. Not only do the Sikyatki Polychrome vessels represent a new and flamboyant ceremonial art style, with especial attention given to stylized birds, butterflies, reptiles, horned serpents, animals, human figures, sun symbols, and crosses, but they are associated with other ceremonial elements: pipes, prayer sticks, pottery bell. One highly stylized depiction, identified by Fewkes as a "man-eagle" (1898, p. 683, fig. 272), is very similar, allowing for Puebloan art conventionalizations, to the highly stylized plumed serpent depictions from Guasave, which in turn appear to be specialized depictions of similar subjects from Aztec I depictions, as discussed above. Another more realistic depiction of man-eagle, recognized by the Hopi as Kwataka from their own mythology, shows him eating a small animal (Fewkes, 1898, p. 692, fig. 275, pl. 146,c). Sun symbols, crosses, feathers, and certain circling geometric layout patterns also resemble Guasave examples. The archaeological evidence for the beginning of the Kachina cult at this time, late Pueblo III and early Pueblo IV, seems very good. DiPeso (1950) has described 108

certain finds of painted stone (altar?) slabs, especially those at Point of Pines dated around A.D. 1350-1450, pointing out their relationship to the Kachina cult. Haury reports the wing of a wooden composite bird and several effigy pahos or "proto-kachinas" from Double Butte Cave, in a Classic Hohokam context, and discusses their identification with modern Hopi kachinas (1945a, pp. 198-200, fig. 128). At Hooper Ranch Pueblo, Paul Martin found a Four Mile Polychrome bowl with three painted kachina representations on the interior, in an Anasazi-Mogollon complex dating about A.D. 1200-1375, the specimen itself probably dating about A.D. 1350-1375 (Martin, Rinaldo, et al, 1961, pp. 122-23, 134, figs. 79, 80). In a crypt in the great kiva of the same pueblo, Martin found a painted stone image, which he identifies as a female cult deity related to the underworld and representing a fertility cult associated with the Kachina cult (Martin, Rinaldo, et al, 1962, pp. 69-74, fig. 42). On an earlier horizon, stated to be Pueblo III, near Flagstaff, Arizona, McGregor (1943) excavated at Ridge Ruin the burial of "an early American magician," accompanied by over 600 artifacts including elaborate turquoise mosaics, lignite mirrors, and ceremonial equipment such as "carved, painted and inlaid sticks, awls, small bows, arrows." The ceremonial objects included wooden wands with wing-shaped turquoise inlaid sections, and wands ending in hands and hoofs. There were ear ornaments and noseplugs. In the room above the burial were two parrot skeletons. The kiva murals, which replace Pueblo III and earlier murals of purely geometric nature, are all dated to Pueblo IV, probably in the 14th and 15th centuries, and are especially well represented at sites such as Kuaua, Pottery Mound, and Awatovi (Smith and Ewing, 1952) ruins in (respectively) the Rio Grande, Rio Puerco, and Hopi mesas in New Mexico and Arizona.

MESOAMERICA & SOUTHWESTERN UNITED STATES

These are elaborate polychrome murals, often life-size, on kiva walls, usually with many superimposed layers, representing animals, men and women, mythological beings, and other items. Human figures are often shown in dance costumes, sometimes masked, or in hunt costumes. Ceremonies appear to be represented, and there is considerable symbolism. The murals are closely associated with the ritual of the Kachina and other Puebloan cults, and their appearance at this time suggests large-scale introduction from Mesoamerica of much of the content of these cults, as pointed out by Brew. It should also be noted that there is considerable ethnological evidence for the introduction to the Puebloans at this time of a new and strong plumed- (water-) serpent cult (Fewkes, 1898, pp. 671-72; Parsons, 1939, pp. 184-85), and that there is also evidence in Pueblo mythology, and perhaps even in the Pottery Mound murals, of human sacrifice. Plumed-serpent depictions in Rio Grande glaze wares and in Pajarito Plateau Biscuit Ware may also reflect archaeologically this resurgent plumedserpent cult. Reviewing all of the evidence, we see that Mesoamerican symbolism, ceremonialism, and ceremonial art swept through the Pueblo IV Anasazi like an early Ghost Dance religion, and that the principal center of its dispersal was the Guasave Aztatlan culture of northern Sinaloa, perhaps aided and abetted by the unidentified Mesoamericans who had occupied Casas Grandes, Chihuahua. In the Puebloan Southwest the new Tlaloc and Quetzalcoatl religious and artistic complexes were adapted to longexistent Puebloan styles, and perhaps superimposed on an early, weak infiltration of Mesoamerican ceremonialism concerned with world quarters and the gods thereof, the fire god, the sun god, a twin war god concept associated with an early Quetzal-

coatl concept, and basic rain-fertility cult ideas. DEVELOPMENT OF THE MESOAMERICANSOUTHWESTERN CULTURAL-GEOGRAPHIC HIATUS AND HISTORIC SOUTHWESTERN ISOLATION

Around A.D. 1350-1400, the Guasave phase collapsed and disappeared, leaving the introverted Culiacan culture to the south as the northern frontier of Mesoamerica on the west coast. In the north Mexican central plateau region to the east, the last feeble phase (Calera) of the Chalchihuites culture disappeared at about A.D. 1350 and the northern Mesoamerican frontier there was withdrawn some 450 km. to the south, to a new stand in southern Zacatecas. In the Southwest, the Anasazi push to the south and west ceased and the Puebloan frontier retreated to Zuni, the Hopi towns, and the Piro settlements on the Rio Grande north of El Paso, Texas. The terminal Hohokam and Mogollon-Anasazi cultures either disappeared from Chihuahua, Texas, southern New Mexico and Arizona, and Sonora, or were so culturally reduced as to be almost unrecognizable. Between Anasazi and the northern frontier cultures of Mesoamerica, there developed a geographic and cultural hiatus, occupied only by tribes of considerably lower cultural pressure. Whatever the reasons for the virtual abandonment of the intervening area by groups of higher culture, Southwestern or Mesoamerican, the historical consequence was the virtual isolation of the Southwestern (Anasazi) Puebloans for a century or more. During this period the rich Pueblo IV infusion of Mesoamerican art and religion was reworked into characteristic Puebloan-Southwestern type. Thus the Spanish conquistadors found it in the 16th and 17th centuries, and the anthropologists in the 19th and 20th centuries.

109

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

REFERENCES Batres, 1903 Beals, 1944a, 1944b Bradfield, 1931 Brand, 1939, 1944 Brew, 1940, 1944, 1946 Covarrubias, 1957 DiPeso, 1950, 1956 Ekholm, 1939, 1940, 1942, 1958 Ferdon, 1955 Fewkes, 1893, 1898, 1912 Gladwin, 1937, 1957 , Haury, Sayles, and Gladwin, 1937 Haury, 1936a, 1944, 1945a, 1945b, 1950a, 1950b Jennings, 1956 Johnson, A. S., 1958 Judd, 1954 Kelley, 1960 and Winters, 1960

no

Kelly, 1938, 1939, 1944, 1945 Lumholtz, 1900 McGregor, 1943 Marquina, 1944 Martin, Rinaldo, et al, 1961, 1962 Mason, 1937 Morris, 1919, 1927 Parsons, 1928, 1939 Pepper, 1906, 1920 Piña Chan, R., 1958 Porter, Μ. Ν., 1956 Reed, 1951 Sauer and Brand, 1932 Schroeder, 1956 Smith, W., and Ewing, 1952 Vaillant, 1931, 1932 Wasley, 1960

6. Mesoamerica and the Eastern United States in Prehistoric Times

JAMES B. GRIFFIN

N

OT VERY much that is new will be presented in this article, but I shall refer to the contributions of anthropologists in Mexico and the United States who have contributed their interpretations of the connections between the two areas. I have used the terms "Mesoamerica" and "Mexican" almost interchangeably because almost all the Mesoamerican influences (Mesoamerica as a cultural term) have passed through a Mexican (either areal, or areal-cultural) filter. The cultural traits which originated or first appeared south of the border and then moved north are regarded as having been incorporated into Mexican or Mesoamerican prehistoric complexes before such movement. Descriptions and interpretations of the Mesoamerican cultural groups and levels of development are given elsewhere in this Handbook; those covering most of the cultural units of the eastern United States are given in Griffin (1952a) or in the more recent publications cited in the following discussion. I have not seriously used the term "Formative" to apply to early agricultural peoples of the eastern United States or

to the more intensively agricultural Mississippi complexes. This does not deny a general relationship between the way of life of these people in the United States and the Mesoamerican Formative, for in fact it is quite evident. It is assumed then, and not demonstrated, that this is the case. The term "Formative" as applied to Mesoamerica is valuable since it implies a development out of which the later more complex societies arise. It is also applicable to the Andean area. In the United States, however, neither the term nor the concept has much practical value. The term "Archaic" for the eastern United States has been used as a time period to apply to a way of life represented by several areal variations which are known primarily within that time period. One can also speak of certain hunting-gathering complexes in Mesoamerica as belonging to an "Archaic" cultural stage equivalent to that of the eastern United States in general evolutionary terms, but I have not employed this device. Although I have used the terms "Classic" and "Postclassic" in Mesoamerica, following 111

FIG. 1—CHRONOLOGY OF EASTERN NORTH MEXICAN COMPLEXES. Chart prepared July, 1961.

AMERICA AND

CERTAIN

MESOAMERICA & EASTERN UNITED STATES

certain current nomenclature, I deplore the connotations of these words, for they are not applicable to the Mesoamerican cultural development. They imply the same kind of cultural recession indicated by the terms "Great Pueblo" for Pueblo III and "Regressive Pueblo" for Pueblo IV in the Southwest. In either evolutionary or functional-developmental terms the general level of the Mesoamerican way of life reaches its most advanced stage in the A.D. 1200-1500 period. It is relatively unimportant if certain areas in Mesoamerica have a cultural recession for the same reasons that in the Mediterranean area it is relatively unimportant, in terms of cultural development, that Rome succeeded the Greek cities as the dominant cultural expression. On the chronology chart I have indicated a beginning of village agricultural societies in Mesoamerica between 4000 and 3500 B.C. There is no factual support for this but, given our current knowledge of the age of agriculture in Mexico, I believe that evidence for such early settled villages will eventually appear. The initial appearance of agriculture in the eastern United States is set between 2000 B.C. and the birth of Christ. Question marks are put in to indicate that the exact time is unknown. PALEO-INDIAN PERIOD

There are a number of different cultures in this time span, which ends at 8000 B.C. (in this article) and extends backward an unknown length of time. It is assumed here that the Sandia hunters, whatever their correct age is, had a similar way of life to that of the only well-defined cultural complex, that of the fluted-blade hunters. These hunters spread over most of the United States between 11,000 and 8000 B.C. The Clovis and Folsom fluted blades are recognizably distinct successive forms in the High Plains. Other fluted-blade types are found in the eastern woodlands, where they appear to have an antiquity comparable to those in the west. No one knows where the

technique of fluting first developed. The remarkable similarity of the artifact complex produced by these early American hunters living in many different environments and hunting different kinds of animals suggests a number of things. Their hunting economy was adaptable to many areas. They were able to spread throughout most of North America south of the ice front without much opposition from possible earlier resident groups. The absence of any indication of the acquisition of new tools implies that there were either no earlier resident groups or, if there were, then no cultural exchange. In the eastern United States there is no reason to call them "big game" hunters, for there is no sound evidence of association of any Paleo-Indian or Early Archaic people with "big game" animals. There was a much greater variety of food available in the east than seems to have been the case in the Plains. Although the mastodon was still available up to 60005000 B.C. in the Great Lakes, we have yet to find a clear association. It is difficult to understand the meaning of two radiocarbon dates from the Lewisville, Texas, site which were more than 37,000 years ago, in spite of a fine presentation of the data from the site (Crook and Harris, 1957). This means that the age of the charcoal is back beyond that point. The Clovis fluted point reported from the site is almost certainly the same age as other Clovis forms, which means some 10,0009000 B.C. Fluted blades have been found in northern Mexico and some have been reported in Central America (M. D. Coe, 1960a). There are an increasing number of indications that during the Paleo-Indian period people penetrated into South America and that they reached southern South America by 8000 B.C. We can expect that the future will provide many sites in Mesoamerica of occupations before 8000 B.C. The earliest radiocarbon-dated Mexican finds, about 7700 B.C, illustrating the activity of Paleo-Indians, are in the Valley of Mexico 113

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

on the northern shores of Lake Texcoco in association with extinct elephants (Crane and Griffin, 1960, p. 43). The projectile points with the elephants are, however, apparently related to the post-fluted-blade Piano forms of the western Plains and to the post-fluted-blade forms to the east of the Plains. There are some indications of other cultural complexes in the western United States during the Paleo-Indian period. The distribution of material called the Desert culture, from Oregon into Mexico at a time just before and after 8000 B.C., implies that it has considerable antiquity. The intergradations of this less specialized western hunting and gathering people with the primarily hunting populations is still to be worked out (Jennings, 1956, pp. 69-72). There are no indications of complexes like the Desert culture in the east during the Paleo-Indian period. The Paleo-Indian period in this paper stops at 8000 B.C. because, by and large, the fluted-blade style has disappeared by this date; because the date is often used to separate the Pleistocene from the Recent although it is a completely arbitrary division; and because I have a feeling that the term "Paleo" should be limited to "old" things, and not last indefinitely. Some archaeologists speak of Paleo-Indian up to 3000 B.C. (Webb, 1960, p. 46), and some to as late as 1500 B.C. The terminology adopted in this paper avoids confusing statements such as, "The Archaic is older than the Paleo-Indian," or "The Archaic at this site is as old as the Paleo-Indian." Furthermore, over most of the United States area with which this article is primarily concerned, there are indications that regional adaptations to the local environments with consequent cultural change began shortly after this period. ARCHAIC PERIOD

For the purposes of this article the Archaic period in the eastern United States is placed 114

between 8000 B.C. and 1500 B.C. During this long period there is gradual change from the primarily hunting economy known earlier, to increasing adaptation to local food supplied in the rivers, lakes, mountains, forests, and sea coasts. There are no marked sudden culture changes which would call for the migration of distinctively new groups into the area. Instead there are changes in the slowly evolving projectile point forms and other implements which gradually produce broad regional differences that are nevertheless tied together by common cultural trends. We have no evidence of the physical type of the PaleoIndian in the east, or of the earliest Archaic people. The tendency toward regional stability of populations probably marks the beginning of linguistic differentiation from the earlier, more or less closely related speech patterns of the eastern Paleo-Indian. In the past 10 years there has come into vogue the use of the term "Desert culture" for a loosely defined cultural adaptation to the relatively dry upland environment in the western part of the United States. In some ways it is similar to the Eastern Archaic way of life. The Desert culture people were small bands of wandering seasonal food gatherers, collectors, and hunters. They ate a wide variety of animal and plant foods, and acquired or developed techniques for small seed harvesting and processing. Their best-known habitations were in caves and rock shelters. This culture type was present in the western area before and during the Archaic and seems to have spread into the east through Texas and Oklahoma, and into northern Mexico by 8000-6000 B.C. (W. W. Taylor, 1956). Because of the dryness of the climate the otherwise perishable remains of basketry, nets, mats, cordage, fur cloaks, sandals, atlatls and shafts, wooden clubs, digging stick and quids have been preserved in their sheltered habitations. In open areas where only projectile points, rough stone implements used as choppers, scrapers, planes,

MESOAMERICA & EASTERN UNITED STATES

and other "paleolithic"-appearing tools are found, the cultural assemblage looks quite different. This cultural adaptation apparently spread southward during the Archaic period where elements appear in different combinations in northern and central Mexico, according to the reports of various archaeologists. There are too few reports, however, on only a small amount of material, so that it is difficult to assess this purported spread of the Desert culture. The term "Plano culture" is applied to the hunting adaptation to the Plains area during the Archaic period from 8000 to about 3000 B.C. This adaptation is a direct descendant of the fluted-blade culture and, through time, includes a considerable number of projectile-point types named by archaeologists. There are a few new tool types such as the Cody knife (Dick and Mountain, 1960). These projectiles are primarily after 8000 B.C., and include Plainview, Milnesand, Portales, Agate Basin, Angostura, Scottsbluff, and others as well. The projectiles found with the Mexican mammoths are said to be related to Plano forms, but the long-blade knives struck from a prepared core reported from the Valley of Mexico do not have counterparts at this time period in the north. The wide distribution of Plano forms across southern Texas implies that similar complexes should extend well south into Mexico. The cultural history of Tamaulipas, as reconstructed by MacNeish (1958), gives an outline which in general terms will probably apply to much of northern Mexico and to the central and southern Mexican area as well. Indeed, MacNeish has informed me that he regards the lower levels of the Santa Marta cave sequence in Chiapas, dated at 6770-5360 B.C. (Crane and Griffin, 1961, p. 120), as clearly related to his early hunting-gathering, and earliest agricultural levels in Tamaulipas. As this is written (May, 1961), excavations near Tehuacan, Puebla, by MacNeish (personal communication) and his associates at the Coxcatlan

Cave show that the earliest occupation includes projectile points related to the Lake Texcoco mammoth sites and to Plano forms. This complex is followed in time by projectile points resembling those of the Early Archaic period, along with scraping planes, bifacial choppers, blades and stone bowls, conical pestles and pebble manos. These people were primarily plant collectors like those known in Tamaulipas from 70003000 B.C. In Tamaulipas the culture sequence is clearly related in projectile-point forms to that of the southern part of the United States from the Southwest, across Texas. The projectile-point forms and other stone tools of similar type are also found in the southeastern United States but without the perishable material which helps to distinguish between the Desert culture and the Eastern Archaic complexes. Many of the projectile-point forms in Tamaulipas are directly comparable to ones in Texas and the Southeast and similar type names are used interchangeably. The changes in style, through time, are parallel in the two areas, and the implication is that there was no barrier to communication and diffusion of projectile-point forms and of other ideas in the period 8000-1500 B.C. There is at present very little evidence to indicate where any of the wide variety of projectile-point styles originated, although it is relatively easy to jump to the conclusion that the primary development of many of the forms was in the United States. This is so, because we have more data on this period from north of the Rio Grande than we have from Mexico. The same comment will apply to such artifacts as the gouge, choppers, bone tools, basketry, matting and cordage. There are many artifact forms which appear in the Middle and Late Archaic of eastern North America which either do not move at all into Mexico or do so in a later and roundabout fashion. Among these are the ground and polished atlatl weights such as the bannerstone, boatstone, and bird115

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

stone, various gorget forms, the grooved axe, and the unique use of copper for implements. The closest observed connections, then, between Mesoamerica and the southern United States during the Archaic are to be seen in the adjacent area of northern Mexico and southern Texas. This desert and semidesert area allowed easier movement of people from the Southwest and Texas into Mexico, and vice versa. There was a clear ecological barrier to significant cultural exchanges between the adaptations of the more humid southeast and those of tropical Mexico. The major differentiation between the United States and Mexico during 8000-1500 B.C. is that in Tamaulipas, and farther south, there is a gradual development of food production following the domestication of plants. The initial agricultural activities are presumably simple and provide only part of the food supply. The sequence of domestication, beginning about 6000 B.C., with pumpkins, peppers, gourds, followed by beans and corn silk between 5000 and 3000 B.C., and climaxed by the introduction of corn at 2500 B.C., is so far unique to Tamaulipas but is probably a widespread phenomenon in Mesoamerica. By 1800-1400 B.C., the domesticated plants include gourds, peppers, pumpkins of two varieties, maize, warty squash (Cucurbita moschata), red and yellow beans, and cotton. Small village agricultural populations should be in existence before 3500 B.C. and before pottery appears. The gourd is said to be known in the southeastern Archaic but this is not yet definitely established and the time of its introduction is not known. None of the other plants named are known from the Southeast during the Late Archaic, but it may be that some of the plants were carried into the Southeast but have not yet been identified because of poor conditions of preservation. With agricultural practices firmly established in northern Mexico by 116

4000 B.C. there is little justification for proposals of independent domestication in the eastern United States of such seed plants as amaranth, chenopodium, and sunflowers, which in a real sense were not domesticates. The evidence for it was never sound and such attempts could well be abandoned, along with such terms as "incipient" cultivation to apply to conditions which have not been demonstrated. In this article agriculture is regarded as intentional planting of seeds for food production. In the heart of the Mesoamerican area from 5000-2000 B.C. there can be postulated a gradual increase in the importance of food production with a concomitant population increase. We can also hypothesize an increased attention of the early farmers to the marked seasonal changes and their relationship to the movement of the sun, moon, and stars. The date for the beginning of the Maya calendar, about 3100 B.C, which has been called "hypothetical" (Morley, 1946, p. 284), may be close to the actual formalization of the basic principles of the Mesoamerican calendrical system. The highly complex calendrical system and integration with the developed religious structure are the result of the Late Formative to Postclassic cultural evolution. By the end of the Late Archaic period at 1500 B.C., in the eastern United States, Mesoamerica is already in what has been called a Village or Early Formative cultural stage with production of pottery, dependence on agriculture, and all the other rights and privileges of this level of development. The only pottery in the rest of North America at this time is the fiber-tempered pottery of the far Southeast, which began about 2000 B.C, seems to have had no ceramic ancestor, and left few, if any, descendants. There is no discernible connection between the Mesoamerican and fiber-tempered developments and Woodland pottery, which has not yet appeared in the Mississippi Valley. Some archaeologists have suggested

MESOAMERICA & EASTERN UNITED STATES

that Woodland pottery is a derivative of fiber-tempered ceramics but the suggested transitions can be interpreted in other ways. If up to this time one tends to see a general movement of cultural traits from north to south, this is clearly reversed in the 1500 B.C.-A.D. 1500 span of time, beginning with the introduction of agriculture into the United States. In the remainder of this article we will normally assume that most of the cultural diffusion was from south to north, on well founded, if not datable, evidence. EARLY AMERICAN CULTIVATORS

In the eastern United States we may identify the next period to be discussed here as that of the rise and climax of the Early and Middle Woodland cultures, from ca. 1500 B.c. to A.D. 400. Some of the distinctive characteristics are: burial mounds; ceremonial earthworks of a number of types and a growing emphasis on burial ceremonialism; the Woodland pottery tradition (Griffin, 1945, pp. 243-46); the introduction of agriculture; a marked increase in trade and barter; and the wide distribution of distinctive art styles. The main economic base in the beginning of the Woodland cultures is still hunting, fishing, and gathering, and camps or small villages are the normal community pattern. By the time of the Hopewell culture, however, before and after the birth of Christ, there are indications of villages occupying up to five acres. There are impressive earthwork structures with enclosed village and burial areas up to 100 acres in size; and some of the walled "fortified" hilltop enclosures are from one to four miles in length (Griffin, 1956). The general level of cultural development of such outstanding population concentrations is roughly equivalent to that of part of the Village Formative in Mesoamerica 10001500 years earlier. Some years ago I divided an earlier comparative study into Early and Middle

Woodland (Griffin, 1949) but here we will consider that these two divisions actually represent a strong cultural continuum, with the dividing line between them even more arbitrary than is normal in archaeological interpretations. The cultural traits which seem to link Mesoamerica and the eastern United States in this period are mounds in which burials are placed, a number of varieties of decoration and surface finish of pottery, clay figurines, earspools, the use of prepared cores and flake knives, and agriculture. The suggestion has been made that Adena, one of the distinctive Early Woodland cultures resident only in the central Ohio Valley, was derived from Mesoamerica by a migration of brachycephalic people. This view was originally espoused as a "speculation" by Webb and Snow (1945, pp. 328-35), supported to some degree by Spaulding (1952, 1955), and again asserted by Snow (1957, p. 60). The majority of American archaeologists, however, have supported the view that Adena is primarily an indigenous development, which was stimulated from a number of surrounding areas, and that elements which might show relations with Mesoamerica have gone through such a transformation en route that they are blurred almost beyond recognition (Griffin, 1949, pp. 81, 87). The majority of Adena traits are those resident in the Ohio Valley in the Late Archaic or those which evolve from the Archaic. Snow's reasons for a migration on the basis of physical type seem to melt away when it is known that there are brachycephalic populations in the eastern United States by 3500 B.C, at the Modoc Shelter in Illinois and in the Laurentian population of the Northeast by perhaps 2000-1500 B.C. In a paper read before the Society for American Archaeology and the American Association of Physical Anthropologists in May, 1961, Marshall T. Newman pointed out a consistent trend toward brachycephalization 117

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

through time in prehistoric American populations and indicated that there was no need to look to Mesoamerica to account for the Adena physical type. G. K. Neumann is of the opinion that Late Archaic populations of the southern Great Lakes area form a strong component of the Adena population (personal communication, May, 1960). In an earlier paper it was stated that "nowhere in Mesoamerica is there anything approaching the burial mound complex as it is known in the eastern United States" (Griffin, 1949, p. 87). This statement is still valid. Archaeologists have also never taken very seriously the possibility of the introduction of burial mounds from Asia via the circum-boreal route. There is quite evident now, because of radiocarbon dates, a series of developmental phases of burialmound building in the Ohio and Upper Mississippi Valley. This sequence begins about 800 B.C. with small low mounds covering a relatively few burials, or a small house structure which burned and was covered with earth. Only in Late Adena and in Ohio Hopewell are there the large complex multitomb and multi-burial structures some 70 feet high, and others some 250 feet long, which date about 200 B.C. to A.D. 200. The large earth mound, conical mound, and large multiple octagon at Poverty Point in northeastern Louisiana (Ford and Webb, 1956) are an anomaly in the period about 800-600 B.C. The radiocarbon dates indicate that the conical mound, apparently associated with burial ceremonies, and some part of the octagon were built during this period. It is less certain that the large mound and much of the surface-collected artifact inventory can be assigned to the same time span. There is nothing in Mesoamerica which resembles this complex, and the Poverty Point octagon figure clearly precedes the Ohio Hopewell earthworks by some hundreds of years. Truncated pyramids of earth as substructures were being built in the Gulf Olmec territory at this time, and in other areas of Mesoamerica as 118

well. If the inspiration for earth-mound building did come from Mexico, there is both a wide geographical gap and a considerable difference in the function of the mounds in the two areas. Figurines at Poverty Point are identified as from "refuse deposits" but whether surface or excavated is not mentioned. They have little in common with Mexican figurines except that they represent people, are made by hand, and are made out of clay. With their temporal position estimated at about 800-600 B.C., they are much too early to have been stimulated by Hopewell figurines—and I prefer Mexico. It may also be noted that according to a recent study, crude figurines made their first appearance in the Southwest somewhere around 300200 B.C. (Davis, 1959). The population growth from 1500 B.C. to A.D. 400 in the eastern United States is one of increase both in the number of sites and in the number of people per village. The population increase is best exemplified at Late Adena and Hopewell sites in the north, and at sites contemporary with Hopewell in the Southeast. The majority of American archaeologists interpret this evidence as a reflection of some dependence on agriculture. There is very little sound evidence, however, for this in terms of preserved plant remains (Griffin, 1960b). Maize has been found at the Harness and Turner Hopewell sites in Ohio and at the Renner site near Kansas City. It has also been found at the Knight Village site in Illinois, but there is reason to doubt that it is definitely associated with the Hopewell complex at that site (ibid., pp. 23-24). The maize from the first three sites has not been studied. The Knight maize, according to H. C. Cutler, "probably resembled Guatemala Tropical Flints" and is thus what is known as eastern-type corn normally associated with the Mississippi period. Corn has been found in rock shelters in Kentucky and in Ohio but it is at present difficult to know when it first appeared (Griffin,

MESOAMERICA & EASTERN UNITED STATES

1952a, pp. 357-58). Both the Knight and Renner sites are dated about A.D. 200, but radiocarbon dates for Turner and Harness are not known. The last named sites could begin somewhat earlier, perhaps before the beginning of our era, but certainly are still in existence at the end of Hopewell. An effigy gourd from the Brangenburg Hopewell mound in Calhoun County, Illinois, is clear evidence of the knowledge of Lagenaria (Griffin and Morgan, 1941, p. 42, pl. 45). This site should date about A.D. 100. The gourd from Sterns Creek has been assigned to the post-Hopewell period since the late 1940's, and a date for Sterns Creek at the Walker-Gilmore site (Champe, 1946, pp. 66-75), M-1129, of A.D. 930 ± 150, has recently been computed at the University of Michigan. Neither the gourds nor the sunflower seeds from the Newt Kash Rock Shelter, or any other Kentucky shelter, can be accurately dated although there is presumptive evidence that people occupied these locations before, during, and after the Early Woodland occupation. The gourd or squash found at two Ohio Adena sites cannot be used as evidence of agriculture in Early Woodland, for on cultural evidence the sites are Late Adena, and the radiocarbon dates also suggest they are Late Adena. They are probably not as late, however, as the University of Chicago laboratory dates of A.D. 440 for Cowan Creek, and A.D. 525 for the Florence Mound (Griffin, 1958, pp. 2-5). There are a number of discussions of the spread of maize into the Southwest and into the eastern United States (Griffin, 1949, pp. 83-84; Jones and Fonner, 1954, pp. 10615). At present there are intensive studies being carried on at a number of centers on the development and differentiation of maize. They do not speak the same descriptive or classificatory language and it is difficult for the nonspecialist to interpret the results. In the summer of 1959, Dan F. Morse excavated an Early Woodland site

in Gordon County, Georgia, for the Georgia Historical Society, where he found a group of 30-40 burnt maize cobs. The two largest cobs are 7.5 cm. long and about 2 cm. in diameter. They seem to be similar to cobs at a second Early Woodland site in northwest Georgia. They are markedly dissimilar to the eastern big butt maize found at the Etowah Mississippi site. H. C. Cutler has most of these cobs for study, and radiocarbon dates will soon be available on the Early Woodland maize, which should be about 100 B.C. The marked increase in burial ceremonialism of Early and Middle Woodland is a social pattern which might have been diffused from the south. It is known in a nascent form, however, in the transitional period between the Late Archaic and Early Woodland in the north, particularly in the Great Lakes-Upper St. Lawrence area. The Adena and Hopewell burial complexes are best regarded as accentuations and elaborations of many of the characteristics of the Red Ochre to Red Paint burial practices. The earspool is probably the most certain artifact form to have come from Mesoamerica. It is present on the unique Adena Mound effigy pipe of clear Hopewell style, and perhaps other earspools are associated with some Late Adena sites. The most common type of Hopewell earspool, the spoolshaped form made of copper, does not have a precise Mesoamerican counterpart. Earspools of this type are one of the diagnostic traits of Hopewell although they do not appear in all areas where Hopewellian characteristics are found. Another form of Hopewell earspool (Shetrone, 1926, pp. 249-52) has been called the "pulley" type. This form is now thought to be restricted to the latter part of Hopewell. It resembles earspools in Mesoamerica from Guatemala to the Panuco River on the east, and Sinaloa on the west. A somewhat similar shape is also present in the later Mississippi period, but these would never be confused with the Hopewell forms. 119

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

Another Hopewellian resemblance to a Mesoamerican practice may be seen in the figurines. The Hopewell clay figurines have a bent-knee stance similar to many of the La Venta-Olmec figures. This is coupled with a naturalistic style which helps to give an additional feeling of resemblance. Covarrubias (1954, p. 249) has described the Adena human-effigy pipe as being "made in the early Olmec Mexican tradition." The Adena pipe represents a dwarf figure, as Snow (1957, p. 55) has pointed out, and this is also one of the distinctive features of many of the Olmec figures. It is a significant fact, however, that although Covarrubias was an ardent diffusionist, he did not see other significant parallels or influences between Mesoamerica and Hopewell. The Hopewellian figurines are concentrated in the north from Ohio to Illinois, Missouri, and northeastern Oklahoma. They are not at all common at any site and do not appear to have been produced as part of a fertility complex, as is said to be the case in Mesoamerica. The holding of infants by Hopewell female figurines is apparently an expression of motherhood. The Hopewell style over this large area is strikingly uniform, and the dress and ornamentation are not Mesoamerican. The figurines are a part of Late Hopewell, about A.D. 100-300. In the spring of 1961, a Hopewell figurine bust and head, with a large and striking headdress, was found at the Mandeville site in Georgia (A. R. Kelly, personal communication ). The site runs from Early Swift Creek to the Mississippi period. The Early Swift Creek occupation of the site has been dated about A.D. 1 (M-1042). Other possible resemblances between Hopewell and Olmec are the striking emphasis on greenstone celts shown by their association with burials in northern Hopewellian, Copena, and La Venta. The celt seems to appear in the eastern United States during the Late Archaic. It is known in Middle America by the Early to Middle Formative (W. R. Coe, 1959, pp. 40-42) 120

and is extremely rare in the Southwest. The grooved axe, which is such a prominent tool in the eastern United States between 4000 B.C. and A.D. 1000, did not diffuse into the Mesoamerican region until after about A.D. 700 and then entered western Mexico from the Southwest (Griffin, 1955). One of the developments at La Venta by 800-400 B.C. is the ability to drill stone, which is found on many of the artifacts. This technique is known in the Late Archaic and must go back to at least 4000 B.C. in the eastern United States. A possible connection between Mesoamerica and Hopewellian on the basis of a common emphasis on blade production techniques was postulated some years ago (Griffin, 1949, p. 89). This technique is in Mexico as early as 8000 B.C. and was, for a time, believed to be almost limited to Hopewell in the north. Since then, the Poverty Point microblade industry has been identified and described in the Lower Mississippi Valley during Late Archaic and Early Woodland times (Ford and Webb, 1956). I regard the Poverty Point industry as a rather local response to a particular ecological situation, for this microflint complex does not seem to have originated outside of, or to have spread from, that area. It does not closely resemble Mesoamerican forms, nor is it close to the Arctic Small Tool tradition (MacNeish, 1959), which, by and large in its spread, does not penetrate even as far south as the Great Lakes area. It is possible that the northern Hopewell varieties of core and blade are an indigenous development on an old tradition of flintworking. William R. Coe (1959, pp. 14-15) has made a survey of the production of flake-blades in the Americas as did Kidder (1947) somewhat earlier. Almost all, if not all, the flake-blades in the United States are themselves tools. The Eurasian and Arctic American production has a strong emphasis on the production of flakeblades from which to fashion a variety of burins and other tools.

MESOAMERICA & EASTERN UNITED STATES

By Middle to Late Formative, as represented at La Venta, such traits as cranial deformation, clay stamps, and cat and bird motifs are known in Mexico. There is very little similarity between the feline portrayals of Olmec and Hopewell. The Adena engraved tablets have some designs with motifs that have some analogies in Mesoamerica, such as the hand and eye and the bird of prey. An obsidian core from La Venta has an engraving representing a bird of prey about to strike its quarry. It has a cross within a circle for the bird's eye (Drucker, 1952, p. 170). The Berlin tablet has an engraved bird on one face with a large circular eye divided into four zones (Webb and Baby, 1957, p. 85). Any connection between the two areas on the basis of such decoration is doubtful. Many of the engraved Adena tablets have a design style very close to the Hopewell style, with the implication that they are Late Adena and either just before or contemporary with Hopewell. It is questionable that the Adena tablets were used as stamps. This brings us to the vexing question of connections between the two areas on the basis of a relatively simple technique of pottery decoration called rocker-stamping. The intercontinental American distribution of this technique has been known for at least 20 years. It is spread in the Americas from Peru to Nova Scotia. It also has an even wider distribution in the Old World and much greater antiquity in that area than in the New World. It has been diffused by archaeologists with much greater speed and abandon than the prehistoric Indians could have possibly done. In northern and northwestern South America it is known by about 1000 B.C. in Formative level sites. The radiocarbon age of more than 2000-2500 B.C. (Rubin and Alexander, 1960, p. 181), made on shells from the Valdivia site in Ecuador, seems excessive and in this article is not regarded as the "true" age. It should be closer to 1000 B.C. There now seems to be little doubt of cultural exchange between

the southwestern part of Mesoamerica and northwestern South America on the Formative level. This exchange continued and the trend toward civilization in Nuclear America was a shared development. The rocker-stamp decoration is part of the Middle Formative in Mesoamerica. It is widely distributed but is not often a major decorative technique. The distribution of the technique in Mesoamerica as it is known now is from northwestern Guatemala and British Honduras to northern Veracruz. At Tlatilco and La Venta it is part of the welldeveloped Olmec level of the Formative, and while the recently discovered Ocos phase in Guatemala and the Pit 50 period are said to be earlier on radiocarbon dates from Chiapas (Dixon, 1959), the time period for this decorative style would seem to be from about 1200 B.C. to 600 B.C. This is substantially earlier than the rockerstamped techniques north of the Rio Grande. A Formative site which provides data, if not clarification, about the rocker-stamped pottery is El Trapiche, excavated by García Payón. This site, in the eastern Veracruz coastal area south of Cempoala, was occupied over a long period. During the Round Table Conference in Jalapa, Veracruz, 1951. García Payón showed various conference members a rocked dentate stamp on an incurving-rim bowl and a plain rocker-stamp decoration on a well-polished flat-base bowl. This latter sherd is particularly interesting because it has a twoline concentric circle design cut by three vertical lines (García Payón, 1950, pl. 12, fig. 4). This same design appears on one Marksville vessel from the Crooks site (Ford and Willey, 1940, fig. 32,á), in Illinois Valley Hopewell (Griffin and Morgan, 1941, pl. 57, no. 2) and from Iowa Hopewell (Holmes, 1903, pl. 171,a). Since then, García Payón has obtained more examples of both zoned and unzoned plain rocker stamp, rocked dentate stamp, and dentate stamp. These specimens were so similar to 121

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

Hopewell styles that a group of Mississippi Valley archaeologists who saw the sherds in Mexico City in December, 1959, thought they were from the United States. García Payón has kindly provided me with photographs of these specimens and with their provenience from El Trapiche. He tends to equate these in time with MacNeish's Panuco I level because of a number of resemblances in the ceramics which MacNeish equates with the Lower Formative. MacNeish had no rocker or dentate stamp sherds from his early level at Panuco. In fact, there are none reported from any level in northern Veracruz or Tamaulipas. If the El Trapiche stamped specimens are closely related to those at La Venta and Tlatilco, such a chronological placement as García Payón has suggested would be too early for the latter Mexican sites, and far too early for a direct connection into the southern United States. Whatever the significance of the El Trapiche specimens eventually turns out to be, it will not be evidence of a ceramic diffusion from the United States into Mexico. In the Lower Mississippi Valley, rockerstamped decoration first appears on Tchefuncte ceramics in a crude simple form of over-all body decoration placed on a smooth plain surface. This type of plain surface treatment is dominant along the Gulf Coast from 2000 B.C. to the end of pottery making in the historic period. Tchefuncte pottery may date from about 400 B.C. to A.D. 500 on the basis of radiocarbon dates (Ford and Webb, 1956, pp. 121-24; Griffin, 1958, pp. 22-24). On cultural grounds, Tchefuncte should end by about 100 B.C. Tchefuncte pottery is a curious mélange of design techniques and styles of the earlier fiber-tempered pottery and of northern Mississippi Valley Woodland (Griffin, 1946, pp. 48-49). Tchefuncte Stamped is manufactured by a variety of techniques and only a minor proportion of this type is plain rocker stamped and a small minority is rocker dentate. The zoned 122

stamped style of Hopewellian pottery and of El Trapiche and Ocos does not appear. Tlatilco has little or no rocked dentate examples (Μ. Ν. Porter, 1953, pp. 37-38) and is not zoned in the manner of Ocos. La Venta Coarse Buff and Coarse Brown ware has a plain rocker stamp and is also decorated by incising, punctating, and by engraving (Drucker, 1952, pp. 81-89). These are also the techniques of Tchefuncte pottery. Tchefuncte Incised has line-filled triangular designs and other patterns similar to those of La Venta. Lake Borgne Incised, Alexander Pinched, and Tammany Pinched of the Tchefuncte complex (Ford and Quimby, 1945) have similarities to La Venta punctated designs. The latter site, however, has zoned punctate styles which, as Wauchope noted for several Formative period examples of this decoration, are reminiscent of Weeden Island Punctate of the A.D. 300-1000 period, and of the even later Fort Walton ceramics along the northwest Florida coast (Wauchope, 1950, pp. 225-26; Drucker, 1952, pl. 20). There also occurred at La Venta, "Three Coarse Brown ware sherds, of the entire lot [over 15,000 sherds], show what appear to be textile impressions [they clearly are] . . . . Such a low incidence is probably to be interpreted as indicating accidental or experimental attempts to vary surface treatment that never attained popularity" (ibid., p. 94). This suggestion of Drucker's is reasonable. M. D. Coe (1960b) and Greengo (1960), in surveys emanating from the same archaeological milieu, assign very early dates to the appearance of Woodland pottery and the beginning of rocker stamping in the United States. Both authors accepted a radiocarbon date of approximately 2400 B.C. (C-794) for the Hunter site, New York, as marking the general time period for Early Woodland ceramics (Libby, 1955, p. 93). The cultural level of this site is better associated with dates of approximately 1500-500 B.C, and although the early Vinette I pottery was found on the site, it

MESOAMERICA & EASTERN UNITED STATES

was not in association with the dated charcoal. This pottery cannot be dated anywhere else earlier than about 1000 B.C. (Griffin, 1961a, pp. 92-93). The developmental sequence in Woodland ceramics in the north has been reasonably clear for some time. Rocker-stamping does not appear until shortly before the beginning of our era in Illinois, in Ohio, or in the Great Lakes area. It appears in Hopewell sites in the north as a technique of decoration which is incorporated into a zoned stamped style already in existence (Griffin, 1952b). This same zoned style also appears in the Lower Mississippi Valley, in the Marksville focus of Louisiana, as an apparently intrusive ceramic complex, and the same interpretation is true for the Santa Rosa-Swift Creek sites in northwest Florida. It is also clear, from radiocarbon dates, that the southern appearance of the Hopewell zoned style of decoration is later than it is in Illinois, and presumably Ohio, and this corresponds with the cultural evidence. In other words, the geographically intermediate area of the Lower Mississippi is later in time than either Mesoamerican or northern Mississippi Valley occurrences and can hardly be employed as a way station for intercontinental diffusion from Asia to Peru, as Greengo (1960) has suggested and implied. If there is diffusion of this zoned dentate style from Mesoamerica to the eastern United States, on our present evidence, it should have been from Veracruz north. If such a diffusion took place, the style should occur in the Lower Mississippi Valley at a time period close to that in Mesoamerica instead of about 800 years later. The approach of Coe and Greengo is similar to that of Willey some years ago (Willey, 1955a, pp. 34-36; 1955b, p. 585; 1956a, p. 89). The same problem is raised with any attempt to explain the origins of any Nuclear American pottery as diffusion from North America. Some years ago I made this suggestion, in jest, as a logical postulate (Grif-

fin, 1945, pp. 235-36), since early simple pottery complexes had not appeared in Middle and South America. In a slightly later paper with Alex Krieger (Griffin and Krieger, 1947), we discussed the appearance of cord- or fabric-marked pottery in central Mexico. This pottery is, I think, on the basis of the data I gathered in Mexico in 1946, best assigned to the Tula-Mazapan and Aztec I-II time span. This ware, now called Texcoco Fabric-Marked (MayerOakes, 1959; Tolstoy, 1958a), has been the object of much recent study. Any connection it may have to Woodland pottery is certainly not clear. The discovery of cord-marked, fabric-impressed, and net- or knot-roughened sherds at La Victoria, Guatemala, by M. D. Coe (1960b), has raised anew the question of connection between Mesoamerica and the United States for these traits. Cord-marked and fabricimpressed pottery is known in the Great Lakes and Ohio Valley area by about 1000 B.C, which is 400 years later than Coe places the Ocos phase pottery. There is no clear ancestor of this Early Woodland pottery in Arctic America (Griffin, 1960a, pp. 810-12) or northeastern Siberia, although the stimulus for its appearance in the United States seems to have come from northeastern Asia. Cord-marked pottery does not appear in the deep south until Hopewell times, and along the Gulf Coast area from Florida to Texas and immediate watershed it never becomes an important technique of surface finish. It is impossible, therefore, for Woodland cord-marked pottery to be seriously considered as an ancestor for similar traits at Ocos. The Ocos zoned style of cord-marking is not North American. Coe recognized some analogies in Japanese Jomon pottery and I will suggest some in Chou China of about 1000 B.C. This raises the suggestion that the most likely stream of ceramic diffusion to account for Ocos cord-marking is the Japanese Current. Another ceramic trait with a presumptive 123

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

source in Mesoamerica is negative painting, which is known on the Middle to Late Hopewell level in Illinois, at the Mandeville site in west Georgia in an Early Swift Creek context, and from the Crystal River site (Willey and Phillips, 1944). The Crystal River examples do not look very similar to any Mesoamerican specimens that I have seen, nor do they appear similar to the Illinois specimens (Griffin, 1952b). In a vague way the Illinois Hopewell negative painted could be likened to the Teotihuacan Late Formative pottery, some few specimens of which are in the collections at the University of Michigan. I have also had the privilege of examining specimens excavated by René Millon from the Pyramid of the Sun in 1959-60. It is certainly reasonable, if not demonstrable, that negative painting was introduced from the Mexican area into the eastern United States, but the way in which this was done is not clear. The general technique may have come in with textile manufacture and then been transferred to pottery. The Hopewellian level negative painted belongs in Late Hopewell and can in no way be regarded as a major cause of Hopewellian development. As far as can be seen now, there is a time gap of about 1000 years between Hopewell and Mississippi negative-painted pottery. In the Sierra de Tamaulipas, MacNeish has described a sequence of ceramic occupations beginning with the Laguna phase which he places about 600-200 B.C. He compares Laguna pottery to the El Prisco period of the Tampico area which is Late Formative to Early Classic. The appearance of these related ceramics in northern Veracruz and Tamaulipas is normally attributed to southern influences from along the Gulf Coast. A number of the Laguna pottery traits such as large solid conical feet on vessels, emphasis on the bowl form, plain surfaces, red filming, and simple incising might suggest Tchefuncte to Marksville pottery. However, the vessel shapes and the location and appearance of the incising and 124

punctating do not suggest any type of direct contact into the Lower Mississippi Valley. The tetrapod supports of Tchefuncte might be viewed as having arisen from the tripod supports of Laguna but this is by no means certain. The vessel forms on which feet appear are markedly different in the two areas. A case could be made for a local Southeastern development beginning in the Early Woodland cultures. Tetrapods do not appear in the north until Late Adena and about Middle to Late Hopewell. Another factor in this is that there is no evidence of ceramics in this time period in northern Tamaulipas or in eastern Texas, either along the coast (Campbell, 1960, p. 170) or inland. The Palmillas Corner-notched and Morhiss Stemmed projectile-point types of Laguna are related in form to types in the eastern United States from about 500 B.C. to about A.D. 400. There are similar notched and stemmed forms from Late Formative sites in Mexico and from the excavations inside the Pyramid of the Sun (Noguera, 1935a, pl. 22). There are, then, a series of cultural traits which are known in the Early and Middle Woodland period sites in the United States which are suggestive of connections with Mesoamerica. Of these, we can say that gourds, maize, and beans are surely brought in as the result of diffusion by the time under discussion, but the exact period for any of the three plants is not known, or the route followed during the introduction. In any event, it was the introduction of agriculture which seems to have caused the considerable cultural growth of the eastern United States and allowed the population to reach a level or stage of evolution which is comparable to that of the Middle Formative in Mexico. Burial mounds may be derived by transference of the idea of mounds as a substructure, to mounds for burials, but this is doubtful. The earspools of Hopewell certainly were derived from Mexico and probably the figurine idea as well. The core-and-blade technique has been sug-

MESOAMERICA & EASTERN UNITED STATES

gested as a possible transfer on this time period but the evidence is not too good. The celt, or ungrooved axe, is a possible connecting link but the history of it in the New World, or any segment of it, has never been worked out. It is significant that this tool is a dominant form in the Late Archaic to Historic period in the eastern United States, from Formative to the Historic period in Mesoamerica, and that it is absent in the Southwest. There are also significant resemblances in projectile-point styles in this time period as was pointed out some years ago (Griffin, 1949, p. 78) and MacNeish's Tamaulipas finds have added strength to the relationships shown by these forms. The varieties of rocker-stamping and the zoned style which is associated with them are strong candidates for direct connections, although the evidence for the transmission is not clear. The same thing may be said about negative painting. On the basis of what we know now, this series of ceramic techniques should have moved from south to north. MISSISSIPPI PERIOD

Although the Mesoamerican influences in earlier periods have been, on the whole, uncertain and difficult to identify, this is not so true of the Mississippi period. It is material of this period that various anthropologists have had in mind when they speak of strong Mexican influences. At present, the Mississippi period may be said to begin with the first introduction of the platform mound, which may be regarded as an initial symbol of some of the important cultural practices of this period. The earthen platform mound was used as a substructure for a building which housed the important social, ceremonial, and perhaps political activities of the village or town group, or of segments of the social group. It is normally associated with a plaza, which was used for important occasions, and with rectangular wattle-and-daub houses. Along with the platform mound and plaza con-

cepts there is an increase in population both at individual habitation sites and in the whole eastern area. It is likely that new types of maize, such as the eastern big butt ear, and perhaps other crops with greater productivity were introduced into the east at this time. It is certain that agricultural practices improved considerably over the preceding period, for efficient flint hoes were extensively used by Middle and Late Mississippi times. Earlier and cruder flint hoes may have been attached at the end of a digging stick and not used with a hoeing motion. These changes appear gradually in a number of areas in the Southeast and are added to the continuing local cultural traditions. No one knows where the significant new traits first appeared, whether singly or in a group, and the manner of spread of the Mesoamerican traits is not certainly known. Some years ago, Radin (1927) employed a boat for a migration of people, and there is a tendency for archaeologists to use such an untraceable device when they think in terms of a migration. Krieger has referred to long migrations of people who left no trace of their path (Newell and Krieger, 1949, pp. 231-32), but such population movements are either traced historically or by means of the recovery of specific artifact complexes which indubitably link two areas together. There is no such evidence for movement between Mesoamerica and eastern North America. The best explanation has been proposed by Kelley (1952), which would allow some Mexican individuals to pass gradually from more advanced Mexican communities across the north Mexican-Texas cultural sink of this period into the more suitable climatic area of the southeast where the native cultural stage was more receptive to influences based on an agricultural economy. While the Davis site, in Cherokee County, Texas, has been proposed as a representative of an Early Mississippi site (Newell and Krieger, 1949, p. 237), there is considerable doubt 125

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

on cultural grounds (Griffin, 1952c) that the Gibson Aspect material at the Davis site can be as early as the carbon black radiocarbon date of approximately A.D. 500. A more reasonable period for the Davis ceramic complex would be approximately A.D. 1000. It has also been proposed that the platform mound is in the Lower Mississippi Valley in the Troyville occupation (Ford and Willey, 1941, pp. 348-49), and some archaeologists still cling to what others regard as a misinterpretation of the data. At the Troyville type site, at Greenhouse (Ford, 1951) and other such sites, in the view adopted in this paper, the platform mounds were built in Coles Creek times and earlier pottery was thrown up into the mound. There are no pure TroyvilleIssaquena sites with platform mounds or plazas, and most of the Coles Creek sites do not have major mound concentrations. Coles Creek probably begins around A.D. 700-900 and the Troyville-Issaquena occupation is about A.D. 300-700 (Griffin, 1958, pp. 2 3 24; this is also close to the current opinion of Philip Phillips and S. B. Williams). Another area in which it has been suggested that platform mounds are early is in southwest Georgia and northwest Florida. This involves the radiocarbon dates on the Kolomoki-Weeden Island cultures, and Willey (1960, p. 86, note 72) believes that platform mounds may date as early as A.D. 350. A series of radiocarbon dates, at this writing, places Early Swift Creek and Early Kolomoki-Weeden Island between A.D. 1 and 600. The Mississippi culture spread into Tennessee and Georgia took place around A.D. 1000. Late Swift Creek and Late Kolomoki-Weeden Island persist up to the Mississippi intrusion. Platform mounds are very rare on Kolomoki-Weeden Island sites (Sears, 1956, p. 67) and not one has been dated. There is no evidence that they occur much before A.D. 1000 in Georgia and Florida and, at present, we cannot say that the platform mound is 126

known in the Mississippi Valley or the southeast before A.D. 700. The point should be made that in the north, between Hopewellian and the Mississippi culture, there is a period of cultural deterioration and stagnation which is represented culturally by the Late Woodland complexes. It has been suggested that this period of decline of about 500 years, between A.D. 200 and 700, corresponds to a slightly colder climate in the northern Mississippi Valley and Great Lakes area (Griffin, 1960b). It precedes the major period of the appearance and growth of Mississippi cultures which occurs between A.D. 700 and 1200 (Griffin, 1960c). In other words, at the same time the Mississippi expansion in the north took place under favorable climatic conditions, the Southwest has its major period of growth and expansion. Both these developments were under relatively strong stimulation from Mexico, and it is tempting to propose that the northern Mesoamerican frontier was undergoing a stimulation at this time. This corresponds with the current interpretations of J. Charles Kelley, who has been studying the connections of northwestern Mexico with the Southwest. The early stimuli into the Southeast for the Mississippi development should arrive during the Tula-Mazapan cultural expansion. In the Southeast there is not much evidence for a cultural decline following the Hopewellian period. There are some continuities between Hopewellian and Mississippi (Griffin, 1946, pp. 72-75, 88), but Krieger's (1945) position, overemphasizing this, has now been substantially modified to the position adopted in this article (Krieger, 1953). Just as it took some time for the Mesoamerican Formative cultures to evolve to the level of development of a widespread religious ceremonialism, so did the Mississippian cultures. Between A.D. 700 and 1000 there must have been an, as yet, undocumented cultural rise which took place from

MESOAMERICA & EASTERN UNITED STATES

Oklahoma to Georgia and from Louisiana to Wisconsin. There are some evidences of common acceptance of ceremonial concepts at this time. It was this relatively long period which set the stage for the dissemination and acceptance between A.D. 1000 and 1400 of the art styles, artifacts, and behavior which characterize the major prehistoric tribal centers of the Southeast, and which has become known as the Southeastern Ceremonial Complex (SCC). Almost all practicing American archaeologists agree that this was primarily a broad regional development. It was not introduced as a complex, and there is no single primary center for its origin and diffusion. It is in the SCC that the greatest number of resemblances to late Mesoamerican art and ceremonial features is to be found. The engraved technique of decoration, often with paint subsequently rubbed into the lines, is one of the best of the ceramic connections between Mesoamerica and eastern North America. The major region for this technique in the United States is the Caddoan area, south of the Arkansas River, north of the Trinity, and west of the Mississippi as far as eastern Oklahoma. There are some general design similarities with the La Salta-Eslabones phases of southern Tamaulipas, and the emphasis in these phases on the incurved-rim bowl and carinated bowl (MacNeish, 1958, p. 172) may have influenced the Caddoan shapes. An extended discussion of this problem has been presented (Du Solier, Krieger, and Griffin, 1947). Further discussion has been given (Phillips, Ford, and Griffin, 1951, pp. 127-29, 178-80), and the interpretation still stands that this technique is one of our best indications of interarea contact about A.D. 700-900. Some specific Mexican artifacts have been found in the eastern United States. Most of these have been in Texas (Krieger, 1953, 1956). We can agree with Campbell's (1960, p. 170) interpretation, "As these ob-

jects were not clearly associated with other archaeological materials, they merely suggest the possibility of trade or travel in prehistoric times." Three-legged metates in the United States have been found in areas along railroad tracks, or are otherwise attributable to recent Mexican importation. There are some ceramic traits which are best viewed as imports from Mexico in addition to engraving. One of the most significant is negative painting which is much more common in the Mississippi period than it was during Hopewell. Its major appearance seems to be about A.D. 1000-1500 and is closely associated with SCC. As a technique of decoration it probably did not carry over from Hopewell into Mississippi but is a new introduction, for the styles, designs, colors, and vessel shapes are quite different. There is a relatively small area gap between San Luis Potosí, where negative painting has been reported (Du Solier, Krieger, and Griffin, 1947, p. 26), and the Mississippi Valley. By and large, the Mississippi negative painting more closely resembles Late Classic and Early Postclassic northeastern Mexican styles than it does that of any other Mesoamerican group. For some time archaeologists have thought of the southwestern United States as much as Mexico, when looking for an outside source for the considerable development of painted vessels in the Memphis area. Because the designs in the latter area are closer to those of the Southwest we should look to that area for the closest connections. While Mississippi tripods seem like an influence from Mexico, the Southeastern potters made a wide variety of shapes at the same time, whereas in Mexico some of these same forms are separated into distinct period styles. The stirrup-neck bottle of Late Mississippi and perhaps the shoe form are to be regarded as Mexican in origin. The bottle, on the other hand, is a form which could easily be copied in clay, 127

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

from gourd containers which had almost certainly been known for a long time. The normal Mississippi bottle form is closer to Mexican Formative bottle shapes than to those of the later more advanced cultures. The Mississippi potters developed an interest in form which produced shapes of many kinds, and even though some of these will resemble Mesoamerican products there need be no direct connection. There are certain resemblances then, for example, between the Southeast effigy vessel complex and that of Chihuahua and the Tarascan area (Griffin, 1949, p. 94), which may be a parallel development. The ceramic complex as a whole does not suggest any kind of population movement of the nature of a migration. In this conclusion we are following the earlier view of Phillips (1940). In earlier periods certain projectile-point similarities have been mentioned; in the Mississippi period there also are projectile shapes associated in the United States with the increasing utilization of the bow and arrow for hunting and warfare. In the Early Classic period at Teotihuacan there seems to be a sudden shift to relatively small stemmed points (Linné, 1934, pp. 147-50; 1942, p. 135; Noguera, 1935a, p. 46; Rubín de la Borbolla, 1947, fig. 9). These points are surely connected to the stemmed tradition of the Caddoan area as represented by such types as Hayes, Alba, Bonham, and others (Suhm and Krieger, 1954, pp. 494502). Such forms do seem to appear in the United States, both in the Southwest and in the Caddoan area around A.D. 700-900. They are older then, according to almost anyone's calendar correlation, at Teotihuacan than in the north. Was the bow and arrow known in Mexico at this time or did the small stemmed form move north while the bow and arrow moved south? Another common arrowpoint shape is the small triangular side-notched form which has its highest popularity from the Southwest to the Mississippi Valley from about 128

A.D. 900 to the Historic period. Arrowpoints of this form are in a collection I made at Culhuacan in 1946 and are also in a collection made by Robert Winters from central Mexico. A variant of this shape with a concave base belongs to Period V of the Panuco area (Ekholm, 1944, p. 491). The plain triangular form which is the dominant form of the Mississippi period is also in Panuco Period V. At Aljojuca in the state of Puebla, Linné found a large side-notched form with a burial in a mound and as a surface find from Tepetitlan (Linné, 1942, pp. 32, 47). One of the best-documented examples of the small side-notched concave base arrowpoint is in the ceremonial burials on the west side of the pyramid at Tenayuca (Noguera, 1935b, pp. 162-65). Noguera also gives a brief account of the arrowpoints from the Cholula excavations from which triangular small stemmed and notched forms were found (Noguera, 1954, pp. 16364; 247-48). It is presumed that future study will have a significant correlation in the changes of projectile-point styles between the two areas, with the small stemmed appearing in the Classic period and an emphasis on triangular forms with various modifications in the Postclassic (Kidder, 1947, pp. 12-14). The long bifacial and bipointed ceremonial knife is a prominent chipped stone artifact in Mesoamerica (Kidder, 1947, pp. 24-25). They are found in Late Formative levels but are more common in Classic and Postclassic sites (Acosta, 1957, pp. 139-48). In the eastern United States they are a feature of the Mississippi period associated with the SCC. They are found in graves and as knives carried in the hand of priest figures (Willoughby, 1932, pp. 54-55) from the Etowah site in the eastern part of the Mississippi culture distribution, to Spiro on the western edge of its distribution (Hamilton, 1952, p. 228). The association of this type of knife with sacrifice in the two areas is probably a better connecting link than the

MESOAMERICA & EASTERN UNITED STATES

presence of eccentric flints in Oklahoma, Tennessee, and in Mesoamerica (W. R. Coe, 1959, p. 25). Any mention of specific similarities or indications of the movement of people between Mesoamerica and the eastern United States should include the mutilated teeth from Cahokia and Fulton County, Illinois (Stewart and Titterington, 1944), which are classified by Javier Romero (1958) as types Al-3, 5, D, and F5. These types of mutilations are post-A.D. 1000 in Mesoamerica; this is also their probable temporal position in the Mississippi Valley. The prehistoric Indians north of the Rio Bravo did not practice dental mutilation. It is assumed in this article that the Illinois occurrences are from individuals who had had dental work done in Mexico. The same observations will be made about a skull from Sikyatki, Arizona, with Al mutilation which should date about A.D. 1300-1400. One of the few traits which can be accepted as moving from the Southeast to Mexico is pipe smoking. This was recognized by 1946 (Du Solier, Griffin, and Krieger, 1947; Μ. Ν. Porter, 1948). Furthermore, it is clear that the initial spread was from the Caddoan areas, on a post-Hopewell to Early Caddo level, into the San Luis Potosi-Tamaulipas area. Pipe smoking caught on much more readily in Mexico than it did in the Southwest, and reflects the more civilized status of Mesoamerica at that time. One of the strongest suggestions of Mesoamerican connections is in the art style and portrayals of priests, priest-warriors, and other concepts which are a part of the Southeastern Ceremonial Complex (SCC). This development, of course, is associated in the United States with the more complex socio-religious organization of the Mississippi culture. With the possible exception of the Natchez, the native tribes of the Southeast did not approach the Mesoamerican elaborate expression of the relation of man

to the supernatural. There are no real "gods" in the Mesoamerican sense. Although the SCC art forms obviously carry a message readily understood by participants in the culture, there is no evidence that such portrayals were combined into a connected historical record. An adequate analysis and comparison between SCC and Mexican representations should also include the mural paintings of the Southwest. It will also require the collaboration of individuals thoroughly versed in late Mexican prehistory and the codices, with those familiar with American archaeology. The position taken in this article is that there is a significant series of resemblances, but not identities, between the beliefs and behavior of the participants in the SCC between A.D. 1000 and 1400, suggesting that some amount of Mesoamerican ceremonial and religious ideology was incorporated and adapted into the Mississippi culture pattern. It is significant that the SCC development and the styles of portrayal on material culture objects most closely resemble concepts and beliefs expressed in Mexico in the Postclassic period and more particularly in the Mixteca-Puebla style. As an example, there may be cited the appearance of the skull, heart, hand, and longbones on pottery from Moundville, Alabama (Moore, 1905, pp. 223-27). These have been compared with portrayals of the Mexican god of death, Mictlantecuhtli, at such places as Tizatlan in Tlaxcala (Caso, 1927). The series of comparisons presented by Krieger (1953, pp. 505-15) are tantalizingly suggestive but still without specific resemblances. They are all from Postclassic sources. Rands (1957) has discussed the hand-eye and related motifs. He has emphasized their appearance in other areas, such as the Northwest Coast, and suggested that some of the designs suggesting diffusion from Mexico may be connected to quite old widespread religious concepts. Their preserved art forms would then have little to 129

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

do with the time of their first appearance and spread. In terms of broad cultural development or evolution, we can recognize basic resemblances between Mesoamerica and the United States during the Paleo-Indian period and also during that of the hunters and collectors of the Archaic period. The general cultural level would correspond roughly to what might be called an Archaic "stage." It is more reasonable, I think, to group the Paleo-Indian hunters with the later pre-agricultural populations than it is to regard them as in an essentially different level of development. In any event, the sharing of common artifact types and ways of life by peoples both north and south of the Mexican-United States border indicates that cultural changes moved back and forth between the two areas with little difficulty, once they had been adopted by groups in the Texas-northern Mexican area. The early development of agriculture in Mesoamerica gradually produced a series of cultural changes which resulted in the regional groups included within the Mexican Preclassic or Formative. This was a significant cultural evolution which had a slowly increasing effect in the eastern United States. In the latter area, however, the long-

resident cultural tradition was stimulated primarily by the adoption of agriculture rather than by major introductions of Mesoamerican practices. In terms of a comparative developmental stage of culture, the Adena-Hopewellian societies were certainly not on a level with those of Mesoamerica during the Late Formative. During the Mississippi period, from A.D. 700 to De Soto, there is a marked shift and reorientation observed in most of the phases of prehistoric life. In developmental terms, compared with Mesoamerica, the Mississippi pattern of life may be equated with that of the Late Formative, with recognition of many striking differences between the two. The Mississippi cultures arose primarily as the result of improved agricultural efficiency. The archaeological pattern we recover indicates a gradual growth over a large area with considerable freedom of exchange of ideas and materials and with some population movement. There was also, apparently, a continuing seepage from Mesoamerica of ideas and practices which, added to the eastern cultures, helped to produce in their increasingly complex social and religious behavior, a definite Mesoamerican or even Nuclear American tinge.

REFERENCES Acosta, 1957 Campbell, 1960 Caso, 1927 Champe, 1946 Coe, Μ. D., 1960a, 1960b Coe, W. R., 1959 Covarmbias, 1954 Crane and Griffin, 1960, 1961 Crook and Harris, 1957 Davis, 1959 Deuel, 1952 Diamond, 1960 Dick and Mountain, 1960 Dixon, 1959 Drucker, 1952

130

Du Solier, Krieger, and Griffin, 1947 Ekholm, 1944 Ford, 1951 and Quimby, 1945 and Webb, 1956 and Willey, 1940 García Payón, 1950 Greengo, 1960 Griffin, J. B., 1944, 1945, 1946, 1949, 1952a, 1952b, 1952c, 1953, 1955, 1956, 1958, 1960a, 1960b, 1960c, 1961a, 1961b, 1962 and Espejo, 1947, 1950 and Krieger, 1947 , McKern, and Titterington, 1945 and Morgan, 1941

MESOAMERICA & EASTERN UNITED STATES

Griffin, J. W., 1949 Hamilton, 1952 Holmes, 1903 Jelks, Davis, and Sturgis, 1960 Jennings, 1956 Jones and Fonner, 1954 Kelley, 1952 Kidder, 1947 Krieger, 1945, 1953, 1956 Libby, 1955 Linné, 1934, 1942 MacNeish, 1958, 1959 Mayer-Oakes, 1959 Moore, 1905 Morley, 1946 Newell and Krieger, 1949 Noguera, 1935a, 1935b, 1954 Phillips, 1940 , Ford, and Griffin, 1951 Porter, Μ. N., 1948, 1953 Radin, 1927

Rands, 1957 Romero, 1958 Rubin and Alexander, 1960 Rubín de la Borbolla, 1947 Sears, 1956 Shetrone, 1926 Snow, 1957

Spaulding, 1952, 1955 Stewart and Titterington, 1944 Suhm and Krieger, 1954 Tax, 1952 Taylor, W. W., 1956 Tolstoy, 1958a Wauchope, 1950 Webb, 1960 and Baby, 1957 and Snow, 1945 Willey, 1955a, 1955b, 1956a, 1956b, 1960 and Phillips, 1944 Willoughby, 1932

131

7. Archaeological Survey of El Salvador1

JOHN M. LONGYEAR,

A

RCHAEOLOGICALLY speaking, El Sal-

vador is really two countries, one lying west, the other east, of the lower Lempa River. The western region, comprising two-thirds of the republic, is dominated by a chain of volcanoes, now largely extinct, which runs in a northwestsoutheast direction across the central part of the country. These volcanoes rise to elevations of over 2000 m. from a plateau some 600-700 m. in altitude. The plateau, in the south, descends abruptly to a low coastal plain, 5-25 km. wide, which borders the Pacific Ocean. To the north, the plateau descends much more gradually, mostly in the form of intermont basins, to the broad valley of the Lempa River, which here flows west to east. North of the Lempa, the land rises again, into the rugged Honduran cordilleras. After flowing eastward through northcentral El Salvador for more than 100 km., the Lempa River suddenly turns south and cuts through the coastal range to the ocean. In this part of its course, the river, although shallow, runs broad and strong, forming an 1 Research for this article was aided by a grant from the Lucius N. Littauer Foundation.

132

III

effective barrier between the western and eastern parts of the country. Much of eastern El Salvador is low-lying. The coastal plain is broad for the most part, and the Rio Grande de San Miguel forms a great basin extending well into the interior. These lowlands, however, are dominated by a trans-Lempan extension of the coastal range which culminates in the great volcano of San Miguel. The country also becomes high and rugged as one approaches the Honduran border in the north. To judge by the distribution of archaeological remains, the aboriginal population of El Salvador tended to cluster in the intermont basins and in the areas surrounding some of the lakes and larger rivers. From west to east, the principal centers of population, as far as our present knowledge permits, may be listed as follows (fig. 1): (1) shores and islands of Lake Guija, in the extreme northwest part of the Republic, (2) valley of Chalchuapa, (3) valley of the Rio Grande de Sonsonate, (4) valley of the Rio Sucio, near Sitio del Niño, (5) intermont basin of San Salvador, and the valley of the Rio Acelhuate to the north, (6) valley of the Lempa River near Suchitoto,

FIG. 1—EL SALVADOR AND PORTIONS OF HONDURAS AND GUATEMALA

FIG. 2—EL SALVADOR, SHOWING LOCATION OF ARCHAEOLOGICAL SITES

133

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

(7) upper valley of the Rio Amate, south of Volcan San Vicente, (8) lowlands around the present city of Usulutan, and (9) valley of the Rio Grande de San Miguel. ANCIENT POPULATIONS

The archaeological evidence of an east-west division of El Salvador is supported by all early accounts of the aboriginal population. Western El Salvador, south of the Lempa, was Maya, probably allied to, if not actually, Pokomam, for much of the preconquest era. By the time of the conquest, however, most of this area had been taken over by the Pipil, a Mexican group, leaving the Pokomam Maya in small enclaves around Chalchuapa, San Salvador, and possibly near Tehuacan, on the eastern slope of Volcan San Vicente, overlooking the Lempa. The area north of the Lempa in this section of the country is almost completely unknown, both archaeologically and historically. From the only information so far published (Lothrop, 1939, p. 46) we must assume that the inhabitants here were Chorti Maya, both before and after the conquest. All sources agree that eastern El Salvador was Lenca, in prehistoric times as well as in the 16th century, except for a small island of Matagalpa-speakers in the extreme northeast corner. This Lenca linguistic group extended north into Honduras at least as far as the Comayagua Valley, and probably right down into the Sula-Ulua plain (Longyear, 1947). ARCHAEOLOGICAL SURVEY

Because eastern and western El Salvador appear to have different archaeological and ethnological backgrounds and connections, they will be discussed separately. El Salvadoran archaeology has not yet progressed to the stage where phase distinctions may be drawn within periods, so the available material has been grouped broadly as Preclassic, Classic, and Postclassic. Although 134

FIG. 3—STRATIGRAPHY IN T H E SAN SALVADOR AREA, a, Cerro Zapote (after Lothrop, 1927a). b, Barranco Tovar (after Μ. Ν. Porter, 1955).

these Salvadoran periods and their connections with other archaeological complexes will be discussed later, I define them here as they will be used in this survey, relying of necessity almost exclusively on pottery. "Classic" refers to a period characterized by an absence of classic Maya or Mayoid or of pottery clearly influenced by and/or roughly contemporaneous with classic Maya polychromes. An example of the latter would be the large class of pottery called by Stone (1957, p. 19) "Ulua Polychrome," found over a wide area in central and southern Honduras and El Salvador. The "Preclassic" period precedes the Classic. Its diagnostics are: (1) the absence of polychrome wares of Classic or Postclassic type; (2) the presence of some or all of the following: handmade figurines, unpainted or monochrome-painted pottery,

ARCHAEOLOGY: EL SALVADOR

and pottery decorated with Usulutan technique. The "Postclassic" period follows the Classic and, like the Preclassic, is marked by an absence of classic Maya or Mayoid polychrome pottery. The polychrome wares which do appear here are late types: "Pipil" or "Nicoya Polychrome." This period is also frequently characterized by Plumbate ware, by large hollow pottery figures, and by incense burners and vases decorated with the head of the Mexican rain god Tlaloc. Central and Western El Salvador The honor of type site for this period should undoubtedly go to the excavation made by Lothrop in 1924 at Cerro del Zapote, just outside San Salvador. In a deep road cut here above the Rio Acelhuate, Lothrop found a dark humus layer some 20 cm. thick under a thick (up to 12 m.) overburden of volcanic ash (fig. 3,a). In the humus layer were a solid, handmade figurine head, fragmentary obsidian blades, and several sherds of orange ware bowls, with thickened and everted lips, supported on small solid conical feet. The bowls were decorated by grooving, particularly around the rim, and by painting in Usulutan technique (Lothrop, 1927a, pp. 173-76, fig. 4). Twenty years after Lothrop's dig, this site was visited by Boggs, who was unable to find any evidence of human occupation in the lower humus layer (Longyear, 1944, p. 14). Lothrop's original find, however, has since been amply confirmed by Porter's excavations at Barranco Tovar, in the southern environs of San Salvador, where stratigraphy similar to that at the Cerro del Zapote exists (fig. 3,b). Porter here recovered over 3000 potsherds, plus parts of three handmade figurines, several fragments of obsidian, and two small pieces of metates from a humus stratum overlaid by a thick cap of volcanic ash (Μ. Ν. Porter, 1955). The principal pottery types are classified by Porter as: (1) dark brown ware low cylinPRECLASSIC PERIOD.

FIG. 4—CHALCHUAPA ARCHAEOLOGICAL ZONE. (After Longyear, 1944.)

der vases, decorated with simple geometric incising, or with a combination of painting in red plus incising; (2) orange ware bowls with lips everted, sometimes grooved or with protuberances, supported on small hollow rattle feet, probably tripod, decoration in Usulutan technique; (3) white ware flat-based bowls, a few decorated with simple incising. Except for the difference in form of support, the Usulutan bowls from Porter's excavation agree generally with those recovered by Lothrop. The only other excavations in Preclassic deposits are reported from the extreme west of the country. Dr. Wolfgang Haberland (1958) recovered pottery and figurine fragments from two sites, one at Atiquizaya, Department of Ahuachapan, and the other on the coast near Acajutla (fig. 2). In both cases, the pottery was in the fill of featureless earth mounds. Brown, orange, and red wares were classified, all decorated with simple geometric grooving and, on the red ware particularly, fine-line incising. Simple rim bands and geometric designs in red paint frequently accompanied the relief decoration. The few specimens of Usulutan technique which were found occurred in sherds from vessels of complex shape, and lacking red rim bands. Figurines were hand135

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

made, with somewhat rectangular flat bodies and prominent buttocks. At Acajutla, two red ware sherds bore thick green and cream paint, similar to a pottery type in the Guatemala highlands assigned by Edwin Shook to a period immediately following Las Charcas (Haberland, 1958, p. 490). Haberland notes that his figurines also resemble Las Charcas types. The last Preclassic deposit in our survey was excavated by W. R. Coe at El Trapiche, one of the sites in the Chalchuapa zone (fig. 4 ) . As above, the artifacts here were recovered from the fill of earth mounds, and consisted of potsherds, with Usulutan decoration common, thousands of broken figurines, obsidian flakes, fragments of metates, cylindrical and flat stamps, pieces of flanged greenstone bowls, and stone bark beaters. At the time of Coe's report (1955) the artifacts had not been studied, so little more can be said about them, except that Coe feels they are late Preclassic in date. In spite of the scant material, we can list features which can be called characteristic of the Preclassic in this region: 1. Brown or red ware cylinder bowls or vases, decorated with simple primitive incising, sometimes supplemented with painting in red, form a prominent part of this Preclassic complex. In seeking comparable material, Porter and Haberland both look toward the Guatemala highlands, and particularly to Mound E-III-3 at Kaminaljuyu where similar forms and wares occur, although the incising on the latter specimens is much finer than that on the El Salvador sherds, and the red painting is lacking (Shook and Kidder, 1952,fig.64). 2. Decoration in Usulutan technique is noted at every site. The principal form is a bowl with slanting wall and everted, grooved lip, the whole supported on small solid, conical feet. Only at Barranco Tovar were the feet hollow. This type of solidfooted Usulutan bowl again is found at Mound E-III-3 (Shook and Kidder, 1952, 136

fig. 72), but it also occurs north of El Salvador, in Archaic deposits at Copan (Longyear, 1952) and in the Comayagua Valley, especially at Yarumela, where it is common in Late Preclassic deposits (Canby, 1951, pp. 81-82). In eastern El Salvador this type appears in collections (fig. 8,a-d), but its exact provenience and associations are unknown. 3. Handmade clay figurines are common, particularly in the western sites. Lack of illustration precludes any detailed comparison of these, but figurines of similar manufacture are frequent in the Preclassic of the Guatemala highlands. In Honduras handmade figurines are much rarer, being characteristically associated only with the Playa de los Muertos complex of the middle Ulua River. On the basis of comparative typology, the bulk of the El Salvador Preclassic material equates with the Miraflores phase of the Guatemala highlands (see Borhegyi, vol. 2, Art. 1; Rands and Smith, vol. 2, Art. 4 ) . The principal diagnostics for this equation are the fine-line incised red ware sherds, and the Usulutan bowls with solid knob feet. Haberland is inclined to place his Acajutla site somewhat earlier, chiefly because of the two polychrome sherds found there (see above), but in any event, this would not push Acajutla back further than Providencia, the phase immediately preceding Miraflores. The hollow feet associated with Usulutan decoration at Barranco Tovar are an indication that this deposit is somewhat later than those mentioned above. In Guatemala, for instance, this combination of support and decoration does not appear until the Arenal phase, whereas it is Early Classic at Copan. CLASSIC PERIOD. This period in central and western El Salvador is well represented by specimens in private collections, but actual sites of this horizon, scientifically excavated and documented, are regrettably few. What excavations we have are due

FIG. 5—CLASSIC POTTERY TYPES FROM CENTRAL EL SALVADOR, α-c, From Club Internacional cache, San Salvador, d-f, Late Classic or Early Postclassic censers from Campana-San Andres. (Boggs, 1945b, figs. 1, 2; 1943a, fig. 3.)

137

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

almost entirely to the efforts of S. H. Boggs, on whose material the following summary is based. The outstanding characteristic of the Classic period in central and western El Salvador is polychrome pottery of Maya or Mayoid type. Spinden (1915) and Lothrop (1927a) both recognized this years ago, and Lothrop actually excavated Maya polychrome sherds from the upper level of Cerro del Zapote (Lothrop, 1927, fig. 10). The best assemblage of Classic material from the central part of the country, however, comes from the city of San Salvador itself, where a cache of pottery vessels was found while the foundations for the Club Internacional were being dug (Boggs, 1945b). The most important types of this collection are: (1) Gray ware jars, high-necked, two handles. Anthropomorphic faces applied on neck, simple geometric designs in red paint on shoulder. (2) Gray ware jars, high-necked, squarish shoulder, no handles. Red panels on neck and shoulder with white rings, in groups and rows, on the red background (fig. 5,b). (3) Red, black, and orange bowls, simple or composite silhouette, designs of glyphlike motifs and monkeys in horizontal bands on exterior only (fig. 5,a). (4) Copador Polychrome simple bowls and vases, with distinctive human figure and glyphic designs characteristic of this ware. (5) Bowls, simple, annular-based, or with tetrapod cylinder support, black and red on orange "rapid" or "shorthand" style decoration of geometric and anthropomorphic figures. (6) Large cylindrical slab-foot tripod vases, Ulua Polychrome (fig. 5,c). (7) Tetrapod bowls, solid animal-head effigy feet, simple geometric decoration in red and purple on cream. A number of distinctive and widespread styles here find a common meeting-ground: Copador (4) and Ulua (6) Polychrome, together with the less well-defined "Rapid" (5) and "Purple" (7) polychrome types. The significance of this association will be 138

discussed later, but in passing, we can mention that Purple Polychrome was found, together with Maya polychrome bearing glyphic designs, at Hacienda Tula, a small site excavated by Boggs 20 km. south of Santa Tecla (Boggs, 1944c). As we move westward, our next Classic site is Campana-San Andres, in the broad valley of the Rio Sucio. This whole valley appears to have been thickly populated in prehistoric times, as mounds and artifacts are everywhere, but Campana-San Andres is the most imposing assemblage of structures and the only one systematically excavated (Dimick, 1941). In plan, the site is comparable to Copan in its plaza-acropolis arrangement, but the construction of most of the mounds—adobe block overlaid with lime plaster—is like that of Tazumal and Kaminaljuyu. This combination of resemblances—some to Copan, others to Tazumal —extends to other features of the site, particularly to the artifacts. Boggs' preliminary report (1943a), while attempting no complete listing of the pottery types here, notes vessels and sherds of "scraped slip" ware as being so common as to represent products of local manufacture. Copador Polychrome also occurs at Campana-San Andres, as does a little Ulua Polychrome. An extension of occupation here into Postclassic times is indicated by finds of Plumbate and Nicoya Polychrome sherds in the latest debris. Several incense burners, some vase-shaped, the rest of tripod ladle type (fig. 5,d-f), are associated with late construction and probably date to Late Classic or Early Postclassic. Although several periods of construction are evident at Campana-San Andres, Boggs feels that this site was formally occupied for only a short time, not more than three or four centuries at most. The last Classic site to be considered in this résumé of central and western El Salvador is Tazumal, one of a number of large mound groups located in the Chalchuapa Valley, near the Guatemalan border (fig.

ARCHAEOLOGY: EL SALVADOR

4). Here again, we are indebted to Boggs for excavation and for a number of reports, all of them preliminary. Although Tazumal is the only site in the Chalchuapa area which has received any kind of systematic excavation, pottery and stone artifacts from the entire zone are plentifully represented in collections. Of particular interest are a number of large stone sculptures, some in the round, others in low relief. Most of these date to Postclassic times, and will be discussed later, but one seems to belong to the Classic period. This is a large lava boulder in the Las Victorias group, carved with four human figures in low relief. Boggs (1950, pp. 85, 90) observes that these carvings are so similar to certain Olmec sculptures from La Venta, Tabasco (Stirling, 1943, especially pl. 41; see also Stirling, vol. 3, Art. 28, fig. 20,a), as to be almost certainly inspired by the same artistic tradition. If so, the Las Victorias carvings should be early to middle Classic in age, according to Drucker's chronology of La Venta (Drucker, 1952). Other chronological interpretations of La Venta (see Stirling, vol. 3, Art. 28) suggest a much earlier dating. These carvings stand alone, however, as nothing else yet described from western El Salvador can be attributed to Olmec influence. Although Coe's excavations show that there was Preclassic occupation of this region, pottery vessels from the Chalchuapa zone which have found their way into private collections are overwhelmingly of Classic and Postclassic types. Of the former, Copador and other Maya polychromes predominate (fig. 6,e,f,g). Boggs' excavations in the large mound (no. 1) at Tazumal are not extensive enough to give any complete picture of this site, but they provide much additional information on the Classic occupation of this area (Boggs, 1943b, 1944c, 1945a; Notes and News, 1946). Most of the construction here is of adobe brick or stone set in adobe, with a facing of lime plaster. In this respect,

Tazumal resembles Campana-San Andres and Kaminaljuyu, and it also resembles these sites in complexity and number of successive building periods. During at least the latter part of the Classic stage, Mound 1 was surmounted by a temple with square columns. The earliest pottery yet reported from the excavations comes from a rubbish deposit underlying the platform on which Mound 1 stands, and resting directly on sterile soil. Even so, the types represented here are no earlier than early Classic, equating with that period at Copan. Particularly diagnostic are bowls with hollow rattle feet, decorated with Usulutan technique plus red paint (fig. 6,d; cf. Longyear, 1952, figs. 55,d, l l l , g ) , and small brown ware incised bowls with knob feet (cf. Longyear, 1952, fig. 56). Later Classic pottery is represented by sherds from the mound debris and by specimens from tombs. Sherd types include vases carved with Maya glyphs, Maya polychromes, Copador Polychrome, local types such as Purple and Rapid Style Polychrome, and fragments of vase-type incense burners, similar to the ones reported from CampanaSan Andres (fig. 5,d). Only one tomb (no. 1) is described in detail. This contained nine polychrome vessels, including one Copador bowl, one Maya polychrome vase (fig. 6,0) similar to a specimen from Tomb 1, Copan (Longyear, 1952, fig. 107,e,e'), one Rapid Style Polychrome tetrapod (fig. 6,b), and two Ulua Polychrome vessels, one a vase (Boggs, 1943b, frontispiece) and the other a large tripod dish (fig. 6,c). Other objects in this tomb were a ladle censer with an effigy handle, two pyrite mirrors, and a jadeite mosaic mask. From other tombs (eight in all, seven with furniture, were excavated) come additional Copador and Ulua Polychrome vessels, more pyrite mirrors, carved jadeite figurines and a plaque, two thin carved stone "hachas," and three fragments of copper-gold figurines (Boggs, 1945a, pp. 35-36). These last are 139

FIG. 6—CLASSIC POTTERY TYPES FROM W E S T E R N EL SALVADOR. All from Tazumal, except g, from Lake Guija zone, α-c were found together in Tomb 1, Tazumal. (Boggs, 1943b, fig. 5; Longyear, 1944, pls. 11, 13, 14.)

140

ARCHAEOLOGY: EL SALVADOR

identified by Boggs as imports from southern Central America. All these tombs are Classic in date, but may have a considerable range within this period. The Tazumal finds give support to the association of pottery types noted above in our description of the Club Internacional cache in San Salvador. Copador, Ulua, and Rapid Style Polychromes are again found together, in tombs and in debris, and Purple Polychrome is also represented in the latter deposits. Ceramically, Classic Tazumal appears to have strong connections with Copan and the Ulua region to the north. Tazumal architecture, on the other hand, lies entirely within the adobe-core-plasterfaced tradition exemplified by southern highland sites such as Kaminaljuyu and Campana-San Andres. In our survey of the Preclassic sites of western and central El Salvador, it was noted that the closest ceramic comparisons were with Kaminaljuyu. In Classic times, however, a new source of influence appears. The Salvadoran sites in the Chalchuapa and Rio Sucio zones are still allied with Kaminaljuyu in architecture and in the presence of a few other objects such as pyrite mirrors. But for imported ceremonial pottery, they now turn north, to Honduras. This change in alliance may have been occasioned by two events which took place in this region at the beginning of the Classic period. At Kaminaljuyu, perhaps because of the conquest of that site by invaders from the Valley of Mexico, ceremonial pottery came to be strongly influenced by the style of Teotihuacan (Kidder, Jennings, and Shook, 1946, pp. 255-56). Vessels in this style, however, never reached El Salvador, so it would appear that trade connections between Kaminaljuyu and the southeast, at least for this class of goods, were disrupted by events at the Guatemalan site. At the same time that Kaminaljuyu was swinging over to Teotihuacan styles, the valley of Copan was apparently invaded by a group of Maya (Longyear, 1952, p. 68).

By the middle of Cycle 9—a century and a half later—this group had established here a large and powerful ceremonial center, whose influence must have been felt over a wide area. It certainly reached as far south as western and central El Salvador, to judge by the frequency of Copador Polychrome in this region. We may picture Classic Tazumal, then, following Kaminaljuyu architecturally, partly because of tradition and partly because of the type of building materials available, but turning northward to newly risen Copan and the adjacent Ulua region for ceremonial pottery, because trade with Kaminaljuyu had been cut off. The new masters of the Guatemalan center looked to Mexico, not El Salvador, for trade connections. Copan, on the other hand, was eager to extend its sphere of influence toward the south. POSTCLASSIC PERIOD. This period in central and western El Salvador is known only from scattered finds. Scanty Postclassic deposits are reported from Campana-San Andres (Boggs, 1943a) and Tazumal (Notes and News, 1946), but no extensive remains of this period have ever been excavated. This is the more regrettable in view of the rather extensive and detailed documentary material which we have relating to the population of this region at the time of the Spanish conquest (Alvarado, 1924; Fuentes y Guzmán, 1882-83; García de Palacio, 1860; Herrera, 1726-30; Ponce, 1873). The only Postclassic tomb thus far excavated and described (Boggs, 1944b) was near the Cerro del Zapote, in the southeastern part of the city of San Salvador. It was constructed of irregular stone slabs and held a single extended burial, an incised Plumbate "lamp chimney" vase, and a monochrome incised tripod bowl. Just outside the tomb, and probably associated with it, was a polychrome bowl with three animal-head effigy legs. Boggs describes this bowl as related to Nicoya Polychrome. Plumbate and Nicoya Polychrome pot141

FIG. 7—POSTCLASSIC POTTERY AND STONE FROM CENTRAL AND WESTERN EL SALVADOR. a,b,d,h,i, From central El Salvador. c,e-g,j,k, From Tazumal zone. (Longyear, 1944, pls. 10, 11, 13, 14; Boggs, 1944a, fig. 2; 1949, fig. 1; Lothrop, 1939, fig. 2.)

ARCHAEOLOGY: EL SALVADOR

tery, on the basis of dates assigned them in other areas, can be considered Postclassic in El Salvador. If, as Boggs believes, they are associated in the tomb described above, they are contemporary, or at least overlap, in their time range, an association supported by similar occurrences in Tombs 10 and 1-40, both Postclassic, at Copan (Longyear, 1952, pp. 43, 44). These two pottery types appear rather frequently in private collections from central and western El Salvador (fig. 7,a-g; Lothrop, 1927a, figs. 21, 22; Spinden, 1915, figs. 75, 76), but record of association with each other or with other types of pottery is completely lacking here. About 50 km. north of San Salvador and not far from the confluence of the Rios Acelhuate and Lempa, lie the ruins of Cihuatan. A brief excavation in 1929 under the auspices of the Salvadoran Ministry of Public Instruction gives some information about the plan and construction of this extensive site but very little about artifacts. The plan is regular, with plazas, some large pyramids, and a ball court (Longyear, 1944, fig. 4). Construction is mainly of squared slabs set in mortar, with facings often covered with lime plaster. Although no surely datable architectural features are reported, the pottery still remaining from the excavations appears to be Postclassic. Unfortunately, although pottery was reported as plentiful at the site, only a handful of specimens still exists in the Museo Nacional, so our sample is far from adequate. Boggs made a study of these and records the presence of sherds from small, blue-painted "Tlaloc jars," which he considers a late type (Longyear, 1944, p. 16). The only other identifiable specimen was a cast of one of several large pottery zoomorphic figures found at the ruins (fig. 7,h). This also must be considered Postclassic. On the basis of this very scanty material, therefore, we can assume that Cihuatan was occupied during Postclassic times, probably by Mexicans (Pipil). If there was also an earlier, Classic, occupation we have no evi-

dence of it, although the possibility of such occupation, in view of the site's size, seems extremely likely. Postclassic remains also exist in the Rio Sucio valley. At Campana-San Andres, Boggs found a few sherds of Plumbate and Nicoya Polychrome in late rubbish deposits (Boggs, 1943a, pp. 118-19). A number of large Tlaloc-effigy censers, found near Campana-San Andres and now in a private collection, give further evidence of Postclassic occupation of this region (fig. 7,i; Boggs, 1949). In the extreme west of the republic, Postclassic remains are numerous. Most of the specimens, unfortunately, are undocumented, but Boggs (in Notes and News, 1946, pp. 207-08) has reported briefly on evidence for a late occupation at Tazumal: In addition, we found this year the remains of two new-type structures which vary greatly from the general Mound 1 norm, as well as from that of Campana-San Andres. These structures differ in plan, construction, and profile from the rest and are stratigraphically later than they are. I suspect they may have been "Pipil" (whatever that means) constructions, partially from the fact that quite a lot of Plumbate sherds were found at their base. . . . Tazumal must have started in Esperanza times and been inhabited more or less continuously until almost, if not, the Conquest. About 1 km. east of the Tazumal group, pothunters excavated a large, hollow, human-effigy pottery figure (fig. 7,j). This subsequently was acquired by the Museo Nacional and was described by Boggs (1944a). The figure stands about 140 cm. high and is an almost exact duplicate of an effigy from Coatlinchan in the Valley of Mexico. The latter specimen was identified by Saville as a representation of the Aztec god Xipe Totec, and was dated by Vaillant (1937, p. 319) to the Mazapan period. Because of its size, the Salvadoran effigy is almost certainly of local manufacture. Boggs (1944a, p. 4) believes it was made by Mexi143

FIG. 8—PRECLASSIC POTTERY TYPES FROM EASTERN E L SALVADOR. a-j,l,m, Vessels with Usulutan decoration. k,n, Orange ware, o, Red ware. a-c,h,i,k-n, From Quelepa zone. (Longyear, 1944, pl. 7.)

ARCHAEOLOGY: EL SALVADOR

can migrants sometime between A.D. 1300 and 1450. Our final evidence of Postclassic Mexican influence in this region consists of a number of large stone sculptures: several colossal jaguar heads carved on boulders, two chac mool figures, and a stela bearing a human figure on the front and glyphs on the sides (the "Virgin of Tazumal"). The jaguar heads have been described by Richardson (1940, pp. 402-03), who relates them to similar carvings from the Pacific slope of Guatemala, and ascribes them to Postclassic Mexican penetrations into these areas. The "Virgin of Tazumal" (fig. 7,k) has often been described as Maya because of the glyphs on its sides, but the glyphs are not of Maya type, and the carved figure bears no resemblance to those found on Maya monuments (Boggs, 1944c, p. 71). On the other hand, the possibility that it may be of Mexican origin has been raised by the headdress of the figure, which appears to represent a Tlaloc face (Boggs, 1944c, p. 72). As a matter of fact, the figure itself has been identified by Henning (1918) as Xipe Totec, although the reasons for this are not clear to me. As with most of the Postclassic pottery specimens discussed earlier, about all we know of these stone carvings is their original location, and not even this for some. Detailed descriptions of their findings, or records of their associations with construction features or pottery are completely lacking. Thus, the Postclassic period in central and western El Salvador is represented for the most part by a fairly large number of isolated, undocumented pottery and stone objects. As we lack excavation of Postclassic remains, our conclusions must necessarily be few and tentative: 1. Among the pottery types characteristic of this period are Plumbate and Nicoya Polychrome (contemporaneous at least in part); large vase-shaped censers; ladle censers with serpent-head handles; Tlaloc-

effigy censers (large) and jars (small), the latter usually painted blue. 2. The period appears to be one in which influences from post-Teotihuacan Mexico are dominant, and Maya or Mayoid influences are, as far as our evidence goes, nonexistent. This is in direct contrast to the preceding Classic period, and doubtless reflects the decline or abandonment of southern Maya centers, particularly Copan, at the close of Classic times. It also supports the theory, gained mainly from historical sources, of an invasion of western and central El Salvador by Mexicans (Pipil) shortly after A.D. 1000. Eastern El Salvador PRECLASSIC PERIOD. SO much has been written about the Preclassic remains of eastern El Salvador that one is likely to think of this period here as one well-documented and known in detail. In reality, however, we know almost nothing about it. Eastern El Salvador has provided a wealth of Preclassic pottery for private collections; Usulutan decoration, a noted Preclassic technique, derives its name from a city and Department in this region, and is thought by many to have originated here. Still, with the one possible exception to be noted below, not a single Preclassic site or deposit has ever been excavated in this section of the republic. The possible exception is a small excavation conducted by Wolfgang Haberland on the Rio Gualacho, just north of El Triunfo in the Department of Usulutan (fig. 2; Haberland and Grebe, 1957). Here, in a coarse soil stratum 2.5 m. thick and beginning 1.5 m. below the surface, Haberland recovered a bowl with four small conical feet, plus an undisclosed number of sherds. This material was not available to Haberland when he wrote his report, so description is scanty and general. The sherds are all monochrome—orange, brown and red— except for a very few with Usulutan technique. The complete absence of polychrome

145

FIG. 9—LOS LLANITOS POTTERY. a,e,f,m-dd,ii, Los Llanitos Polychrome. b,d,h,k,l, Unslipped ware, c, Red-on-orange ware. g,ff-hh, Fine Line Polychrome, i,j, Ulua Polychrome ware. (Longyear, 1944, figs. 24, 25.)

ARCHAEOLOGY: EL SALVADOR

wares seems to indicate a Preclassic date for this site, but the necessarily brief description helps us very little in determining the composition of the deposit, and, of course, we cannot tell if it is typical of the Preclassic in this area. A study of pottery specimens from eastern El Salvador in private collections indicates that rather close ties exist between this region and the Comayagua Valley to the north, in Honduras. The latter area has been investigated thoroughly by Stone (1957) and Canby (1951); by analogy with their finds, we can characterize the Preclassic ceramic complex of eastern El Salvador as being composed exclusively of monochrome and bichrome wares, with Usulutan technique prominent among the latter. Vessel supports are tetrapod, small solid knobs, and handles and spouts are frequent. There is a considerable use of modeling and grooving, especially on rims and shoulders. The vessels in figure 8 would fall into the Preclassic, according to the above definition. Obviously, only excavation can tell us whether our characterization of the pottery of this period is anywhere near accurate or complete. Excavation is also needed to give us some idea of the nonceramic artifacts and types of construction which may accompany this pottery. CLASSIC PERIOD. This period in eastern El Salvador is represented by one site, Los Llanitos, which I excavated and described some years ago (Longyear, 1944), and, as usual, by a number of pottery vessels in private collections. Los Llanitos is a small site lying in the valley of the Rio Grande de San Miguel, 22 km. south of the city of San Miguel (fig. 2). The site consists of 10 mounds grouped around two small plazas, and a ball court, which forms the northern boundary of one plaza. The mounds are outlined by crude walls of pumice blocks set in adobe, with the space between the walls occupied by a fill of claylike soil containing numerous sherds and other artifacts. In addition,

crudely squared pumice slabs face the sloping benches of the ball court and the ends of the mounds bordering the court. Nowhere is there evidence of more than one period of construction, and the uniformity of the pottery throughout the excavations supports the conclusion that this is a oneperiod site. About half the pottery recovered is from unslipped utility vessels, large and thickwalled, the majority in the shape of twohandled storage bowls (fig. 9,b,h). Other important local types, to judge by frequency of occurrence, are a white-slipped ware, stained orange, and a ware with similar slip, plus polychrome decoration in red and black. In both these wares, by far the commonest shape is a shallow bowl, set on four hollow conical or cylindrical legs. The polychrome ware, called "Los Llanitos Polychrome" in the report of the site, is a highly distinctive pottery, with decoration characterized equally by boldness of motif and carelessness of execution (fig. 9,m-dd,ii). In 1944 there was nothing with which to compare this pottery, but it now appears certain that it can be equated, at least in style of decoration, with a ware described by Stone from the Comayagua Valley in Honduras and called by her "Las Vegas Polychrome" (Stone, 1957, pp. 3 3 34, frontispiece, fig. 20). The tall-legged tetrapod bowl form, however, so typical of Los Llanitos Polychrome (fig. 10,α), seems to be completely lacking in the Comayagua region, where the principal shapes for Las Vegas Polychrome are simple bowls and jars and squat tripods. A number of pottery vessels, almost certainly imported, were found in cache deposits at Los Llanitos. One cache yielded six bowls, all with short solid conical legs, two tetrapod and four tripod, and a small effigy jar or censer cover (fig. 10,b-h). The bowls all bear painted designs, five of them in polychrome, but neither the designs nor the shapes of these vessels give us a clue to their site of origin. For the present, they 147

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

FIG. 10—LOS LLANITOS POTTERY. (Longyear, 1944, pl. 6.)

must be considered unique. Another cache contained a number of ladle incense burners, all decorated with knobs. One handle terminates in a crude hand (fig. 10,i), another is surmounted by a well-modeled jaguar effigy (fig. 10,k). The base of a knobbed, vase-type censer also was found in this cache (fig. 10,j). These censers bear a general resemblance to those found in Late Classic or Early Postclassic debris at Campana-San Andres (fig. 5,e,f), Tazumal (Boggs, 1943b, fig. 54; 1944c, pp. 67-68), and Copan, Honduras (Longyear, 1952, figs. 102,b; 108,m,n). A few other imported pottery types are represented in the sherd material. Among these are flat-based, footed bowls of Fine Line Polychrome (fig. 9,g,ff-hh), of the 148

same type as vessels previously noted in this report from the Club Internacional cache in San Salvador and from Tomb 1 at Tazumal (fig. 6,b). These vessels are also known from Classic deposits at Copan (Longyear, 1952, pp. 100-01, fig. 79), and occur in association with Ulua Polychromes at La Ola, on the lower Choluteca River (Stone, 1957, figs. 47,A,b; 77,A). A few sherds of Ulua Polychrome slab-foot tripod vases were also found at Los Llanitos (fig. 9,i,j). Nonpottery artifacts of clay are rare at Los Llanitos, consisting only of a few fragments of solid figurines, mostly of the "cleft foot" variety, and several spindle whorls. Stone implements include manos and metates (the latter invariably undecorated),

FIG. 11—CLASSIC POTTERY TYPES FROM EASTERN EL SALVADOR. b,c,f-j, From Quelepa. (Longyear, 1944, pl. 9.)

149

FIG. 12—POSTCLASSIC POTTERY AND STONE FROM EASTERN EL SALVADOR. b-d,l,m,o-q, From Quelepa zone. (Longyear, 1944, pls. 7, 8, 9, 12.)

ARCHAEOLOGY: EL SALVADOR

polishing stones, numerous blades, and a few points, both the latter of obsidian. To summarize the ceramic situation at Los Llanitos, the dominant polychrome type there is Los Llanitos Polychrome, apparently a local variant (particularly with respect to vessel shape) of a ware found rather widely to the north and east, where Stone reports it, under the name Las Vegas Polychrome, from several sites in the Comayagua and Tegucigalpa regions. Ulua Polychrome, on the other hand, is rare at Los Llanitos. Quite the reverse of this situation obtains in private collections from eastern El Salvador. Here, Ulua Polychromes of various types are common, while Los Llanitos Polychrome is apparently entirely lacking. Fine Line Polychrome, however, is represented both in collections (fig. 11,f) and at Los Llanitos, but is never common. A sampling of Classic polychrome types in private collections is shown in figure 11. Eastern El Salvador, then, like the western two-thirds of the country, had its closest ties during the Classic period with Honduras to the north. The influence of Copan, however, did not extend this far east, as no Copador or other Maya polychromes are in evidence here. Neither do we find in the east any particular evidence of communication with central and western El Salvador. Los Llanitos bears little resemblance to Campana-San Andres or Tazumal, in either architecture or ceramics. The principal lines of influence and trade appear to extend from eastern El Salvador to the north and east, particularly to the upper reaches of the Humuya River, between Comayagua and Tegucigalpa. Not only is there considerable evidence of trade in pottery between these two regions, as noted in several places above, but there are similarities in architecture as well. The ball court at Tenampua, for instance, to judge from Stone's description of it (1957, pp. 52, 55), employs stone slabs on facing and borders just as does the court at Los Llanitos. Trade in pottery appears to have been

overwhelmingly one-way, from north to south, but this impression may result in large part from a general ignorance about pottery types indigenous to eastern El Salvador. It is distinctly possible, for instance, that the Los Llanitos Polychrome style inspired Honduran potters to produce Las Vegas Polychrome. Although Los Llanitos is the only excavated Classic site in eastern El Salvador, Classic occupation of this region was widespread. Most of the polychrome pottery in collections comes from the Quelepa zone, just west of San Miguel, and I have collected polychrome sherds of Classic type on the surface of mounds at Santa Rosa, near the summit of Cerro Cacaquatique in the Department of Morazan, and from Isla Conchaqua in the Gulf of Fonseca. In addition to Los Llanitos, furthermore, the entire valley of the Rio Grande de San Miguel yields evidence of extensive Classic deposits. POSTCLASSIC PERIOD. Not even one Postclassic site has been excavated in eastern El Salvador, and all our knowledge of this period, therefore, comes from reconnaissance and from examination of material in private collections. Plumbate and Nicoya Polychrome pottery can be considered indicative of the Postclassic period here, just as they are in western El Salvador. Both types, however, are rare. Only five examples of Plumbate are known from eastern El Salvador, three from sites in the Department of Usulutan (Shepard, 1948, p. 112, fig. 20,k,k') and two from Quelepa (fig. 12,b; Shepard, 1948, p. 112, fig. 15,u). The extreme drop in frequency of this ware from west to east in the republic is another indication that the Lempa River was an effective barrier to travel and trade in this direction. The rarity of Nicoya Polychrome is more surprising, however, considering the proximity of eastern El Salvador to the homeland of this ware. Haberland and Grebe (1957, p. 285) report two vessels of Nicoya Polychrome from a mound at La Rama, south of Usulutan, and I have 151

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

picked up a sherd or two of Nicoya Polychrome at the site of Quelepa (Longyear, 1944, p. 14). These finds, however, plus one vase "from eastern El Salvador" in a private collection (fig. 12,a), constitute the entire corpus of the ware in this region. Aside from the above two pottery types, we have no ceramics which can with certainty be assigned to the Postclassic period. If we go through the numerous collections of pottery from this area, however, and eliminate all known Preclassic and Classic types, we are left with a considerable body of material which cannot be placed as to period on the basis of excavation in this or neighboring regions. Some of these vessels, at least, must date to Postclassic times. One type, rather frequently represented, which we believe merits such classification is a two-handled bowl with globular body and vertical or slightly flaring neck. The exterior is customarily slipped in orange, and simple linear designs are painted on this in red (fig. 12,c). The case for a Postclassic dating of this type is strengthened by the specimen shown in figure 12,d, which bears a Tlaloc effigy. Another frequent type of pottery which may have Postclassic affiliations is a small, unpainted, two-handled jar, with a conventionalized animal or human face applied on the neck (fig. 12,e,f). Effigy jars appear in Classic deposits at Los Llanitos (fig. 9,d), but they are larger, bear red paint, and have a different placement of the handles. There are some other effigies from eastern El Salvador, most of them quite grotesque, which have a Postclassic appearance or "feeling," at least to me. One is probably a representation of Tlaloc (fig. 12,k), another is a human head or skull (fig. 12,n), and there are other curious forms (fig. 12,g-j,l,m,p-q). Almost all the specimens mentioned above are reported to have come from Quelepa, but without excavation at this or at other sites with Postclassic occupation, it would be fruitless to push this 152

"classification" of Postclassic pottery any farther. We are on slightly firmer ground in assigning a few collected stone artifacts to the Postclassic period. One is a thin, tenoned stone head from Quelepa, undoubtedly a representation of the Mexican god Xipe (fig. 12,o). This almost certainly should be dated to a period following the Pipil invasion of western El Salvador in the 11th or 12th century A.D. Postclassic influence from the south is shown by a number of greenstone effigy pendants, all with "duckbill" appendages over the lower face (Longyear, 1944, pl. 12, 18). These are related to, if not identical with, Chorotegan pendants from the Nicoya Peninsula, Costa Rica (Lothrop, 1926, fig. 11), shown by Richardson (1940) to date to a period considerably later than the Maya Classic. Note that all the finds discussed above have been small, easily portable objects. It is also worthy of comment that objects displaying external Postclassic influences, either from across the Lempa to the west, or from Nicaragua and Costa Rica to the east and south, are extremely rare. There is no evidence, therefore, to indicate that the transition from Classic to Postclassic in eastern El Salvador was occasioned or even accompanied by any movement of populations, either into or out of the region, a conclusion I reached on somewhat the same grounds several years ago (Longyear, 1947). What does seem to have brought about the shift was the cessation of Classic influence from the north. This influence owed its origin, ultimately, to the Maya, and its most obvious material expression was in polychrome pottery of Mayoid type, made by non-Maya peoples—presumably Lenca—of central Honduras and eastern El Salvador. When Copan was abandoned by the Maya ceremonial cult, shortly after A.D. 800 (Longyear, 1952, pp. 74-79), Maya influence disappeared from this entire area. Ulua Polychrome and other Maya-inspired

ARCHAEOLOGY: EL SALVADOR

polychrome pottery types may have continued in vogue for a short time after this, but not for long (Epstein, 1959), and with their demise, the Classic period in the El Salvador-Honduras area gave way to the Postclassic. In contrast to western El Salvador, where the Postclassic was accompanied by an invasion of Mexicans, the Lenca of eastern El Salvador and south-central Honduras were left pretty much alone. Here, the void—at least in pottery—left by the departed Maya was filled by expansion of a number of local styles. Of these styles, however, in spite of the "classification" attempted above, we know almost nothing at the present time. Our survey of the Postclassic period in this region should conclude with mention of a number of sites which appear to have been occupied shortly before, and during, the Spanish conquest. Most of these are fortress sites, concentrated in the mountainous country to the north. Three such sites are on the peaks of Volcan Sociedad in the Department of Morazan. One of them, Coroban, I visited and found traces of low stone walls, a few irregular stone mounds, and numerous obsidian chips and blades, but no pottery (Longyear, 1944, p. 12). The other sites on the volcano are similar, if we may judge by native descriptions. These sites do not appear to have been inhabited by many people or over a long period of time; in truth, they are not suitable for habitation, from lack of permanent water supplies and their extreme elevation (about 1000 m.) over the surrounding countryside. Since local migration legends dating back to about the time of the conquest make frequent mention of these fortress sites, it is assumed that they are of fairly recent occupation. Another site, on Conchagua Island in the Gulf of Fonseca, was certainly occupied at the time of the conquest, and for some years following. Conchagua Vieja consists of a

number of stone house mounds and the ruins of a Spanish church. Potsherds of coarse undecorated brown ware, fragments of obsidian, and metates litter the surface in abundance. Fray Alonso Ponce preached at this village in 1586 (Ponce, 1873, p. 384) and it may also have been visited by Vásquez de Espinoza in 1613 (Vásquez de Espinoza, 1942, p. 232). Evidence of an earlier, Classic occupation of this island has been noted above. CONCLUSIONS

By now, it should be abundantly evident that El Salvador does not comprise a distinct archaeological or ethnological region. It is, rather, an extension of other regions, particularly of those to the north and west. Further than this, the country itself does not constitute a unit. The eastern and western sections act independently of each other for the most part, each pursuing its own course of history and making its own cultural alliances. In the Preclassic period, western El Salvador was closely tied to the cultures of the Guatemala highlands. At the same time, influences from the north or east, indicated by pottery decorated in Usulutan technique, were making themselves felt. The exact source of this Usulutan influence is unknown, but quite probably it stems from the Comayagua region, or from somewhere between there and the Pacific coast, through eastern El Salvador. In Classic times, western El Salvador's connections with Guatemala were broken off, and external relations were dominated by Copan and south-central Honduras. The Postclassic period witnessed a return of relations with the west, this time through an actual invasion and conquest by Mexicans, while influences from northern areas died out. This state of affairs persisted until the Spanish conquest. The course of events in eastern El Salvador was less erratic. All through its history, this region appears to have formed a unit 153

FIG. 13—CULTURAL RELATIONS BETWEEN EL SALVADOR AND NEIGHBORING GIONS, a, Preclassic period, b, Classic period, c, Postclassic period.

RE-

ARCHAEOLOGY: EL SALVADOR

with the central highlands of Honduras, immediately to the north. During Copan's florescence, in Classic times, a good deal of Maya influence was felt in this area, but this seems to have been the only external cultural force ever to make much of an impression here. A good many years ago, George C. Vaillant used a series of diagrams to chart the distribution of pottery types found at Holmul, Guatemala (Merwin and Vaillant, 1932, fig. 26). Since most of El Salvador's relations with neighboring areas are indicated by portable artifacts of pottery and stone, diagrams similar to those made up by Vaillant have been used here to illustrate the changing cultural influences which affected El Salvador, west to east, during prehistoric times. Figure 13,a depicts the Preclassic period. Influences from the Guatemala highlands reach western El Salvador in the form of various fine incised pottery vessels and rare stucco polychromes. From somewhere to the northeast, perhaps from Comayagua Valley, this same region receives Usulutan tetrapod bowls, a dominant ware from Comayagua south to the Pacific coast of eastern El Salvador. This type of Usulutan pottery reached Copan at this time, and also appears in the Guatemala highlands. The latter region could have obtained this ware either from Copan or from western El Salvador. In the Classic period, diagrammed in figure 13,b, the Guatemala highlands are invaded by Mexicans, as shown by a host of Teotihuacan traits, mostly in pottery, at Kaminaljuyu. Some of this pottery also reached Copan by trade. Copan, at this time, blossoms forth as a Maya ceremonial

center and spreads its influence east, resulting in Ulua and other Mayoid polychromes which dominate the ceramic scene throughout central Honduras and eastern El Salvador. The Maya culture of Copan also extends to the south, where western El Salvador, cut off from Guatemala by the Mexican invasion there, now imports Copador Polychrome direct from Copan, and Ulua Polychrome from central Honduras. The final diagram, figure 13,c, shows Postclassic distributions of influence. The Mexican invasion of western El Salvador, by way of the Guatemala highlands, is evidenced by a multitude of portable and nonportable artifacts of stone and pottery, Plumbate ware being conspicuous among the latter. Copan now goes into almost total eclipse, by virtue of its abandonment by the Maya hierarchy, but does receive a little Plumbate, probably from Guatemala. Central Honduras also nearly drops out of the picture, but eastern El Salvador, although no power in itself, becomes an avenue for the trade of Nicoya Polychrome vessels into western El Salvador. In return, eastern El Salvador receives a very few Mexican artifacts from west of the Lempa River. From western El Salvador, too, may have come the few Nicoya Polychrome specimens found at Copan. This reconstruction of El Salvadoran prehistory, filled as it is with speculations or with gaps in knowledge that no guess could span, is nevertheless, we hope, essentially correct in its broad outline. I offer it merely as a framework which future archaeological work in this region may fill in, or twist into a new shape, or even completely take down and reassemble, as the evidence directs.

155

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

REFERENCES Alvarado, 1924 Boggs, 1943a, 1943b, 1944a, 1944b, 1944c, 1945a, 1945b, 1949, 1950 Canby, 1951 Coe, W. R., 1955 Dimick, 1941 Drucker, 1952 Epstein, 1959 Fuentes y Guzmán, 1882-83 García de Palacio, 1860 Haberland, 1958 and Grebe, 1957 Henning, 1918 Herrera, 1726-30 Kidder, Jennings, and Shook, 1946

156

Longyear, 1944, 1947, 1952 Lothrop, 1926, 1927a, 1939 Merwin and Vaillant, 1932 Notes and News, 1946 Porter, Μ. Ν., 1955 Ponce, 1873 Richardson, 1940 Shepard, 1948 Shook and Kidder, 1952 Spinden, 1915 Stirling, 1943 Stone, 1957 Tax, 1951 Vaillant, 1937 Vásquez de Espinoza, 1942

8. Archaeological Survey of Western Honduras

JOHN B. GLASS

I

boundaries of modern Honduras include archaeological manifestations that relate to both Mesoamerica and lower Central America. The western part of Honduras (fig. 1) is decidedly Mesoamerican throughout the known course of its prehistory, whereas the cultural remains of eastern Honduras relate to styles and traditions of lower Central America. Most attempts to define a southern boundary for Mesoamerica recognize this division (Kirchhoff, 1943, for example). The archaeology of Honduras as a whole was summarized in the Handbook of South American Indians (Strong, 1948) but the inclusion of much of western Honduras therein was essentially an attempt to relate the archaeology of Honduras to the betterknown culture sequences of Mesoamerica rather than to define culture-area boundaries. The major outlines of the archaeology of western Honduras are relatively well known, despite certain important gaps, and a working chronology has been established, owing largely to the Smithsonian Institution-Harvard University expedition of 1936 (Strong, Kidder, and Paul, 1938, hereafter HE PRESENT

abbreviated as SKP, 1938). Since the summary article by Strong in 1948 a number of contributions to the archaeology of the area have been made. The first of these was Longyear's (1947) discussion of ethnic and archaeological correlations along the southern frontier of Mesoamerica, which summarized problems which still persist. In 1951 Canby published a summary of the long ceramic sequence which he obtained at the site of Yarumela in central Honduras. The third contribution in point of time was the full publication of the ceramics and the phases of Copan, the important Maya ceremonial center in far-western Honduras (Longyear, 1952). The archaeological survey of central and southern Honduras by Stone (1957) provided new data on areas of Honduras that were formerly unknown. Finally, an article by Epstein (1959) discussed the dating of the Ulua Polychrome complex and briefly summarized his unpublished sequence established for eastern Honduras and the Bay Islands. The publications of both Canby and Epstein are based on unpublished doctoral dissertations (Canby, 1949; Epstein, 1957). Recently I have been engaged in a 157

FIG. 1—WESTERN HONDURAS. (After Stone, 1957, fig. 1.)

ARCHAEOLOGY: WESTERN HONDURAS

re-analysis of the material excavated by Gordon for Peabody Museum, Harvard University, in 1896-97; by Strong, Kidder, and Paul in 1936; and of certain other collections. This analysis is still incomplete and insufficiently advanced to warrant publication here. It promises, however, to define an Early Postclassic phase for the lower Ulua region and to provide a temporal segregation of certain aspects of the Late Classic cultures of northwestern Honduras. For present purposes western Honduras may be divided into five regions: Far-western, Northwestern, Central, Southwestern, and Southern. The first region, the Farwestern, might well be considered a subregion of the southern Maya lowlands. The Maya site of Copan is located there, as are other sites where Classic Maya sculpture has been found. Archaeological manifestations attributable to the Maya in this region will not be discussed at length. It may be noted here that the Copan River, which drains the Valley of Copan, flows into the Motagua River, thus removing Copan from the river-drainage systems of Honduras. The Northwestern region centers on the valley of the lower Ulua River and the Lake Yojoa subregion. This region of Honduras has been the most thoroughly investigated. In Late Classic times it was distinguished by a group of painted pottery types and a style of polychrome painting known as Ulua Polychrome. The presence of this ceramic group is the primary criterion for defining Western Honduras as used here. The Central, Southwestern, and Southern regions are based on data included in Stone's archaeological survey of central and southern Honduras. Her survey is practically the only source for information on the Southwestern and Southern regions. The eastern boundary of western Honduras, which, of course, is thus part of the southern limit of Mesoamerica, may be described as a line paralleling the lower course of the Ulua River on the north, then turning southeast to follow the Comayagua

or Humuya River and its tributary, the Rio Grande or Sulaco, thence south to the upper reaches of the Choluteca River. The latter river forms an indefinite boundary on the south to its flood plain on the Gulf of Fonseca.1 The geographical basis for this demarcation is dependent on the riverdrainage systems of Honduras. Thus, westtern Honduras includes the areas drained by the rivers that flow into the Gulf of Fonseca as well as the areas drained by the Ulua River and its tributaries. The areas drained by the Aguan, Paulaya, Patuca, and Segovia rivers, which flow northeasterly into the Caribbean, together with the Bay Islands, define eastern Honduras. The southern limit as here proposed accords generally, though not specifically, with the proposal made by Stone (1959), which is based, in part, on ethnohistorical rather than on purely archaeological data. FAR-WESTERN HONDURAS

As might be expected, Mayan remains are found in the western parts of the present Departments of Copan and Santa Barbara that lie adjacent to the Guatemalan border. The sites of Copan, Santa Rita de Copan, and Paraíso, all of which have monumental sculpture in the Classic Maya style of Copan, are located on tributaries of the Motagua River in Guatemala. The sites of El Puente and Los Higos, on the upper Chamelecon River, for the same reasons, are Mayan and unaffiliated with other cultures of Honduras. These sites have been reviewed by Yde (1938, pp. 40-58). Copan, with its monumental architecture, its acropolis, corbeled vaults, and hieroglyphic stelae, shares few distinctive traits 1 Ed. note: The Pacific coast and slopes of Nicaragua and northwestern Costa Rica may also be considered within the Mesoamerican orbit (see Kirchhoff, 1943; Willey, Ekholm, and Millon, vol. 1, Art. 14), although they are definitely marginal in their Mesoamerican cultural characteristics. Eastern Honduras, as defined and explained by Glass, is generally considered to be outside Mesoamerica and so is dealt with as part of lower Central America (see Lothrop, Art. 9, this volume).

159

FIG. 2—NACO AND ULUA BICHROME CERAMICS, a-d, f-g, i-j, Naco red or black-on-white. e, Miniature tripod vessel, h, Net sinker (?). k, Sherd from "grater-bottom" bowl, from Naco, Honduras, l-n, Incised, everted lip, Usulutan, and rocker-stamped sherd from Ulua Bichrome complex.

160

ARCHAEOLOGY: WESTERN HONDURAS

with the rest of Honduras during the Classic period.2 Ceramics of the Archaic (Formative, Preclassic) phase of Copan, however, may well prove to have been part of a ceramic complex that extended at least into the lower Ulua subregion of Northwestern Honduras. In the Early Classic period the ceramics of Copan show their greatest degree of relationship to other regions of the Maya lowlands at a time for which no complexes are known for the rest of Honduras. The most distinctive Late Classic (Copan "Full Classic") pottery type at Copan, Copador Polychrome, has a distribution that encompasses eastern and central Salvador as well as Copan but it is almost absent from other regions of Honduras. Interrelationships between Copador Polychrome and Ulua Polychrome are nevertheless evident in pottery forms and in the placement of designs and decorations on bowls. Occasionally a design element characteristic of one of the types is found on a vessel of the other. Ulua Polychrome was traded to Copan, where it has been found in Late Classic sherd collections as well as in tombs. Somewhat removed from the Maya sites mentioned above, but still in the Far-western region, is the site of Naco, located on an affluent of the Chamelecon River. Naco is a contact site and is mentioned in all records of the conquest of Honduras and in particular by Bernal Díaz del Castillo, Montejo, and Las Casas. To judge from these records, Naco was the center of a relatively populous area, the language of which may have been Nahuatl or Nahuat. The site was visited by Strong, Kidder, and Paul in 1936, when they made test excavations which are reported in their preliminary publication (SKP, 1938, pp. 2734). The site is extensive and consists of platform mounds, some of which were original2 Ed. note: Copan is considered as part of the Maya lowlands (see vol. 2, Art. 13).

ly faced with cut and squared stones and which supported heavy plaster floors. Some of these were stained a deep dark red; other fragments, apparently from walls, show successive layers of different-colored washes. One ball court, with a fragment of a stone ring, was partially excavated. The ceramic assemblage of Naco includes a very distinctive red and black-on-white type with red-on-white and black-on-white variations (fig. 2,a-d,f,g,i,j). The typical form of this class of pottery is a tripod bowl which frequently has three lateral projections near the bases of the feet (fig. 2,f,i). The interior of the bowl is sometimes carved in a piano-relief technique to form relatively elaborate designs (fig. 2,k). The painted designs are geometric and curvilinear with straight lines, dashes, simple scrolls, and dots. The bowl form and the "grater" bottom also occur in one of the plain wares. Ladle or frying-pan censers (fig. 2,j), small tripodal single-cell candeleros or miniature vessels (fig. 2,e), textile-impressed pottery, carved spindle whorls, net sinkers (?) (fig. 2,h), and both hollow and solid modeled clay figurines are among other components of the assemblage. One fragment of 16thcentury majolica was found at the site. The Naco phase is the only complex in Honduras that has been assigned to the Late Postclassic period. Its relationships are to the north and it does not appear to be an indigenous development. Indeed, its most distinctive trait, Naco Polychrome, may be part of an as yet vaguely recognized regional style of red-on-white ceramic decoration. Chinautla Polychrome, which was in use in highland Guatemala at the time of the conquest, is another member of this horizon (Wauchope, 1948, pp. 156-57; Woodbury and Trik, 1953, pp. 195-96). A generalized relationship to certain examples of Managua ware from Nicaragua, which also have greater bottoms, may be mentioned (Lothrop, 1926, vol. 1, fig. 107). Other manifestations deserving a note in 161

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

the Far-western region are the bottle-shaped pottery vessels associated with cremations found in caves in the hills of the Copan Valley (Gordon, 1898b). These have not been related to other complexes of Honduras though M. D. Coe (1957b, p. 17) remarked on their similarity to bottle forms from the Formative or Preclassic period culture of Tlatilco in the Valley of Mexico. Near Naco, Blackiston (1910) reported on a cache of over 800 copper bells, together with a mask that had once been inlaid with turquoise. This discovery, from a cave, is an isolated find. NORTHWESTERN HONDURAS

The valley of the lower course of the Ulua River, also known as the Plain of Sula, and the Lake Yojoa subregion have been the scene of relatively extensive archaeological exploration. Together they constitute the known portions of Northwestern Honduras. In 1896-97 Gordon excavated at the sites of Playa de los Muertos and Santa Ana, both on the Ulua River, and indicated a difference in depth of deposit between the monochrome and polychrome wares (Gordon, 1898a). Later, Dorothy Popenoe (1934) isolated at Playa de los Muertos a series of burials which were quickly recognized as a phase of the Mesoamerican Formative or Preclassic (Vaillant, 1934).3 Further excavations at the site by Strong, Kidder, and Paul confirmed the stratigraphic position of the Playa de los Muertos phase (SKP, 1938). The latter party also excavated at other sites in the Plain of Sula but particularly at the sites of Santa Rita and Las Flores. The sites just reviewed are cemeteries and deep refuse deposits situated along the high banks of the Ulua or Comayagua Rivers. Although Steinmayer (1932) had provided sketch-maps of mound groups in the Plain of Sula, it was not until 3 Ed. note: At that time the terminology used was "Archaic" rather than Formative or Preclassic. In this article Glass uses the term Formative for this period.

162

1941 that specific data on architectural remains of the subregion became available with the report on Travesía (Stone, 1941). Archaeological data on such sites as La Ceiba and Los Naranjos and Jaral at the northern end of Lake Yojoa come mainly from the report of Strong and his associates (SKP, 1938) and from the survey by Yde (1938). These explorations have provided a large body of descriptive material on the ceramics and artifacts of the region and four assemblages have been defined. Three of these, Yojoa Monochrome, Ulua Bichrome, and Playa de los Muertos, are Formative in date (Strong, 1948, fig. 15; Wauchope, 1950, pp. 236, 239-40, 250). No known assemblage or complex in the region is assignable to either the Lithic (Paleo-Indian) or Early Classic periods, though scattered finds of individual pottery vessels suggest the presence of an Early Classic phase. Early Classic Maya basal-flange bowls, probably left by itinerant Maya, have been reported from the lower portion of the Plain of Sula (Epstein, 1959). No name as such has yet been given to the complex represented by the Ulua Polychrome ceramic group and associated wares and artifacts, which can now be dated with certainty as co-eval with the Full (Late) Classic phase at Copan. With the courteous assistance of Robert E. Smith, I have identified a fifth, Early Postclassic, isolate. These ceramic assemblages and their related complexes are described below. Yojoa Monochrome In 1936 Strong, Kidder, and Paul (SKP, 1938, pp. 111-15) discovered a stratum of refuse stratigraphically separated from and underlying a Late Classic polychrome zone at the site of Los Naranjos at the northern end of Lake Yojoa. The collection that was recovered contained approximately 700 sherds, most of which were of a crude unpainted and unslipped ware. Only 12 of the sherds showed traces of black, buff, red, or white paint. Vessel forms represented

ARCHAEOLOGY: WESTERN HONDURAS

include small, flat-bottomed vessels and low slightly flaring lips. There were no spouts, handles, lugs, or feet in the collection. Two solid hand-modeled figurines suggest a relationship to Playa de los Muertos types. Some of the specimens have been illustrated (SKP, 1938, pl. 18,c-w; Strong, 1948, pl. ll,n-f). The Yojoa Monochrome assemblage can be dated as Formative on the basis of its stratigraphic position, rim shapes, and figurine forms (Strong, 1948, fig. 15; Wauchope, 1950, pp. 236, 250). Its relationships to other Formative remains in Honduras, however, are unclear. Ulua Bichrome The assemblage known as Ulua Bichrome was isolated at the site of Santa Rita, which is located on the Comayagua River near its confluence with the Ulua. Like the Yojoa Monochrome deposit, it underlay a Late Classic polychrome horizon. The collection is small as the 1936 expedition was able to sample only a portion of the deeply buried midden. Nevertheless its temporal position and relationships can be better defined than is the case with Yojoa Monochrome. One sherd, decorated with an indented fillet, which separates a red-painted zone from an unpainted zone, is the same as a Copan Archaic type, "coarse ware filleted bowls" (Longyear, 1952, fig. 32), and may be a trade piece or may indicate some closer cultural relationship between the two regions at this date. The same cluster of attributes occurs on a type from Playa de los Muertos (fig. 3,c). A coarse red-on-white type is also present in the Playa de los Muertos collection retrieved by Strong, Kidder, and Paul. A polished plain ware exhibits the characteristic Formative rim shape, which is sharply everted and which is sometimes incised on the flat upper surface (fig. 2,1). Usulutan ware, with various types of resist decoration, is also present (fig. 2,m) as are such traits as rocker-stamping (fig. 2,n), fine and groove incision, pottery stamps, and small nubbin feet.

The small size of the Ulua Bichrome collection presents various difficulties in its interpretation, but its Formative status is clear (Strong, 1948, fig. 15; Wauchope, 1950, p. 250). Whether or not it is contemporary with Playa de los Muertos or Copan Archaic or whether it may bridge a temporal gap between them is a problem that must await further knowledge of the Ulua Bichrome phase. Playa de los Muertos The site of Playa de los Muertos is a cemetery and refuse zone that has been exposed by the Ulua River and that is, and has been, continually eroded away by the annual flooding of the river. Gordon excavated here in 1896 and in his report illustrated a number of whole monochrome vessels, but, as he did not publish data on their associations, no complex could be defined and some of the vessels in question are considerably later in date (Gordon, 1898a, pl. 7). In 1929 Dorothy Popenoe excavated a series of burials, most of which were accompanied by grave furniture; these served to define the Playa de los Muertos complex (Popenoe, 1934; Vaillant, 1934). By 1936, when Strong, Kidder, and Paul excavated, the river had washed away that portion of the site opened by Popenoe. Their excavations revealed a typical Late Classic polychrome deposit overlying the more deeply buried and earlier material. Their sherd collection duplicates some but not all of the types found by Popenoe. Vessels excavated by Gordon, and now in the collections of Peabody Museum, Harvard University, also exhibit types and traits not represented in the Popenoe and Strong, Kidder, and Paul material, but whether these differences are due to an amplification of data or to temporal differentiation at the site has not been analyzed in print. The following description of the complex considers the SKP and Popenoe material as representing the same phase but excludes the material excavated by Gordon. 163

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

The various ceramic wares and their most common forms have been discussed and illustrated by Strong (SKP, 1938, pp. 6275). The ceramic types include a ware with a thick, white slip (fig. 3,f), a dark gray to black, highly polished ware (fig. 3,h,k), a sooty gray ware (fig. 3,j), a polished cream to light gray ware, and a polished red type (figs. 3,g; 4,b). An unpolished red-on-white and a type with a white underslip, covered with broad areas of red paint, leaving the white showing in reserve-space designs (fig. 3,á), are well represented. A redpainted type with the red applied within zones outlined by incision (fig. 3,a,b) or on one side of an indented fillet (fig. 3,c) also occur. Various types of incised designs are found, including gouge and groove incision in lineal patterns (fig. 3,e,h,i); more elaborate designs (fig. 3,i) are less common. Gadrooning, fluting, punctation are among other decorative techniques. Everted and incised and thickened lips are illustrated in figure 3,g,h. Vessels with vertical spouts, both bridged (fig. 4,b) and unbridged, are common. Several effigy vessels with stirrup spouts have been attributed to the complex (M. N. Porter, 1953, pl. 14, for example) and one was found by Popenoe (fig. 4,a). Hand-modeled, solid figurines are an important part of the Playa de los Muertos complex (fig. 4,c-g). Human female figurines are the most common. A pregnant woman and one carrying a child on her back (fig. 4,d,e) are illustrated here as well as one modeled in the full round (fig. 4,c). Several unusual objects of an unidentified form (fig. 4,h) were found by Popenoe. Numerous isolated finds attributable to this complex have been found elsewhere in the Plain of Sula as well as in the Comayagua Valley of central Honduras and indicate that the Playa de los Muertos culture was widespread in the Northwestern and Central regions of western Honduras. It is usually dated as Middle Formative (Wauchope, 1950, pp. 239-40; MacNeish, 1954; Sorenson, 1955). 164

Ulua Polychrome

Complex

The most distinctive and publicized remains from Northwestern and Central Honduras are pottery vessels that belong to a large and varied class of polychrome ceramics collectively known as Ulua or Ulua-Yojoa Polychrome. Few sites in Northwestern Honduras or in the Comayagua Valley of Central Honduras fail to exhibit the presence of Ulua Polychrome and, aside from the Formative assemblages reviewed above, it is present in most collections from the Northwestern region. Although the group as a whole has been well published (Gordon, 1898a; Yde, 1938; SKP, 1938; Strong, 1948; Stone, 1941, 1957), no completely satisfactory typological or chronological analysis has been published. The interlocking categories of Mayoid, Bold Animalistic, Ulua and Yojoa with lower (Santa Rita) and upper (Las Flores) divisions proposed by Strong (SKP, 1938; Strong, 1948) are imprecisely defined. The additional categories used by Stone (1957) present a more static arrangement of styles and designs, but as descriptive classes of individual vessels they do not pretend to provide any chronological ordering of the group. Until 1952, with the publication of Longyear's report on the ceramics of Copan, the date of the complexes that included the Ulua Polychromes was in doubt. Seven tombs and graves at Copan, dated as Full Classic, contained as many whole Ulua Polychrome vessels, six of which are illustrated by Longyear (1952, figs. 104,g; 105,a; 109,h; 110,c; 117,c,g); Ulua Polychrome sherds were also found in refuse deposits of the same date. This finding served to establish the contemporaneity of the Ulua Polychromes with the Full Classic phase at Copan (probably equivalent to Tepeu 1 and 2 of the Uaxactun sequence) and thus provide a date for the associated remains in Northwestern and Central Honduras. The question was reviewed by Epstein (1959), who noted the presence of Ulua Polychrome

FIG. 3—PLAYA DE LOS MUERTOS. Sherds illustrating various ceramic types of the Playa de los Muertos phase.

FIG. 4—PLAYA DE LOS MUERTOS, a, Effigy vessel with stirrup spout, b, Polished red vessel with bridged spout and incised decoration, c-g, Solid hand-modeled figurines, h, Problematical object.

ARCHAEOLOGY: WESTERN HONDURAS

trade pieces in the Classic Selin phase of eastern Honduras and suggested that its absence in the subsequent Cocal phase, which contained plumbate pottery, provided a terminal date (i.e., pre-plumbate) for the group. No full publication has yet been made of the complex to which the Ulua Polychromes belong. All excavations have been of a testing or sampling nature. The largest of these were at Santa Rita and Las Flores (SKP, 1938), Santa Ana (Gordon, 1898a), and Travesía (Stone, 1941), all in the Plain of Sula; and at the sites of La Ceiba and Los Naranjos (SKP, 1938), at the northern end of Lake Yojoa. Strong (SKP, 1938) noted a marked difference in the collections from Santa Rita and Las Flores in both the Ulua and other polychrome types as well as in the plain wares. He utilized this difference in suggesting an upper and lower division of the Ulua Polychrome complex but, in the context of a preliminary report, did not amplify or demonstrate the basis of the division. Inasmuch as both his upper and lower polychrome types are present in collections of Full Classic date at Copan (Peabody Museum collections), it would appear that the temporal difference between his divisions are encompassed within the Late Classic period. I have recently sorted the polychrome sherds from Santa Rita into individual vessels with several interesting results. The first of these is the demonstration that Levels 3-12 of Excavation 1 (SKP, 1938, fig. 6) are a contemporary and functionally selected deposit of ceramic refuse, with its great depth explained by its having been thrown down a talus slope of the river. The second is the identification of an Early Postclassic horizon marker as the dominant component of Levels 1 and 2. The sherd collection from Levels 3-12 is evidently neither a random nor a typical collection of the Late Classic complex, as polychrome and decorated types far exceed the number of plain and utility types. The Gordon collection from Santa Ana,

in the Peabody Museum, although selected by Gordon for decorated pieces, contains several types of Ulua Polychrome which are absent to rare in the Santa Rita and Las Flores collections. As it also contains most of the Santa Rita types as well as Early Postclassic material, the complex situation represented by these collections is evident; we can only lament that Gordon's excavations were not conducted according to modern stratigraphic techniques. Preliminary seriation of the collections from the Plain of Sula and from the Lake Yojoa subregion have indicated the possibility of defining regional and temporal variations in the Ulua and other polychrome types as well as in the plain wares. Some of these will be discussed below. This rather tedious recital of the nature of our sources of information on the provenience of the Ulua Polychrome complex has been designed to show the complexity of the present status of our knowledge. The picture is that of a cultural assemblage of considerable variety in typological, chronological, and regional aspects; yet, we can safely presume that the entire manifestation is restricted in time to the Late Classic period. The following review of the archaeological content of the Late Classic complex of Northwestern Honduras must be considered as selective and as generalizing about the region as a whole, not as applying to any specific site except as noted. Further research and publication will undoubtedly contribute refinements in the three aspects that we have mentioned. As we have frequently noted, the polychrome ware known as Ulua Polychrome, and here treated as a group of types, is the single most outstanding feature of the complex. The variation in the group is extensive in slip or ground color, in forms, and in the range of subject matters represented in the designs that occur. Following the precedent set by Strong (SKP, 1938; Strong, 1948) and followed, with alteration, by Stone (1957), the group may be said to polarize 167

FIG. 5—LATE CLASSIC CERAMICS, NORTHWESTERN HONDURAS, α-c, Mayoid. d, Fine-line Mayoid. e, Santa Ana classes of Ulua Polychrome. f, Large hollow figurine.

around three classes, based not on typology but on style and the representative content of the designs. The first of these corresponds generally to what Strong termed "Mayoid." It is best represented in the collection from Santa Rita though it occurs at all the sites to which we have referred. The designs of this class suggest that they are part of a definite and fixed repertory of ritual iconography. If any vessel can be said to be typical of the class, it is the cylindrical flatbottomed vase with a central band of "dancing" figures drawn as isolated figures in a processional sequence (fig. 5,a), in facing pairs (fig. 5,b), seated (fig. 5,c), or within 168

circular medallions. A row of glyphlike elements, usually heads, almost invariably encircles the vessel in a band near the top and bottom of the design. These elements are also found together on simple bowls with indented bases and tripod bowls with flaring sides. A repetitive design, termed counters, is frequently found on the interior of vessels of this class (SKP, 1938, fig. 14). Space requirements preclude a more extensive treatment of the designs of this class, which are by no means limited to those illustrated here. Individual vessels frequently have more than five different deliberate colors, all of which are hues of red, black, and

ARCHAEOLOGY: WESTERN HONDURAS

orange. A white color appears less frequently and the slip color is usually a light orange. The semitransparent quality of the pigments indicates that they were mixed in a fine suspension, a trait that distinguishes most of the Ulua Polychrome from other polychrome and painted types in the complex. A particularly fine example of an important variation of this class, known as "Fine-line Mayoid" is illustrated in figure 5,d. The second class of Ulua Polychrome has been designated as "Naturalistic" by Stone (1957, p. 24) after the manner of portrayal of animal and other forms usually depicted, though highly conventionalized and abstract forms are also common in the class. Vessels of this class usually tend to be bowls, rather than cylinders, and small handled jars with rounded shoulders. The most common design is that of an aquatic bird (fig. 6,a,b) or of a monkey (fig. 6,c,d). Alligators, a piglike form, serpents, and less readily identifiable and monstrous forms are also found. The third class of Ulua Polychrome is one that was not well represented from Santa Rita and Las Flores and was thus not singled out for special treatment by Strong nor given special status by Stone. Two nearly complete examples of this class, both from Santa Ana, were illustrated by Gordon (1898a, pl. 4, 5). The most distinguishing characteristic of the Santa Ana class is the thickness of vessel walls, which are consistently heavier than those of the Mayoid class, and in the rim profile which is thickened and bulges slightly before tapering or curving inward at the top (fig. 5,e). Longyear (1947, p. 2) was the first to recognize the sorting value of this mode, which may well prove to be of chronological significance. Cylindrical vases of this class are usually supported by low round or slab-footed tripod feet or have a false bottom consisting of an annular base fully as wide as the vessel itself. The main design is usually a series of human figures, each

different, engaged in some activity. When the cylinder vase has a protruding knob representing the head of a monkey or bat, above the drawing of its body, the human figures appear isolated in panels but otherwise are depicted in a processional sequence. They are more boldly drawn than figures in the Mayoid class and their costumes are distinct. Large serpents, sometimes depicted as emerging from a shell (SKP, 1938, pl. 12,b), felines, and a wide variety of other forms also appear as elements of the main design. As in the Mayoid class of Ulua Polychrome, there is usually a narrow band of repeating designs below the rim. Only a few of these are shared with the Mayoid class. Chevrons, step-frets and scrolls, and interlaced textile designs are among the wide variety of decorative devices found in the sub-rim area. The "counter" design which occurs on the interior walls of Mayoid vessels does not appear. In the Mayoid class the existence of virtual duplicates in vessels and their design combinations is quite common; this has not been observed in several hundred specimens of the Santa Ana type that I have examined. The preceding review of the Ulua Polychrome group has only touched on the complexity and variation of these ceramics and not at all on other polychrome types that may be given independent status. It should be noted here that even the broad tripartite segregation of the group here presented cannot always be applied to specific examples. Although Mayoid and Santa Ana cylinder vases can be sorted with ease, both classes merge imperceptibly in the area of the naturalistic bowls. Here the beveled rim profile can be found in vessels that, on the basis of other criteria, might be classified as Mayoid or as naturalistic (fig. 6,e,f). The distribution of Ulua Polychromes in group is fairly wide along the southern periphery of Middle America. In western Honduras it is present in all the regions defined in this article. In Far-western Hon169

FIG. 6—LATE CLASSIC VESSELS, N O R T H W E S T E R N HONDURAS, a-f, Naturalistic bowls of Ulua Polychrome, g, Jar from Lake Yojoa decorated in Bold Geometric style, h, Marble vase.

duras, however, particularly at the site of Copan, its presence is undoubtedly due to trade rather than to local manufacture or use. In eastern Honduras it is found as trade pieces in the Selin phase (Epstein, 1959, p. 126). An occurrence in Guatemala is an isolated find (Kidder, 1949, p. 16, fig. 6,d). At El Cauce, near Managua, Nicaragua, Ulua Polychrome sherds from all three of the classes that we have described have been found (Richardson and Ruppert, 1942, p. 271; Peabody Museum collections). 170

The distribution of Ulua Polychromes in El Salvador is wide. Whole vessels have been excavated at the site of Tazumal in western El Salvador (Boggs, 1943b, frontispiece; 1945a, fig. 5,c) and at the Club Internacional site in the central part of the country (Boggs, 1945b, figs. l,g; 2,g) in contexts that are contemporary with the Full Classic period at Copan. Numerous examples from El Salvador are known and large private collections of vessels usually include representatives of the Ulua Poly-

ARCHAEOLOGY: WESTERN HONDURAS

chrome group. Some of these, particularly from San Salvador and Quelepa, are illustrated by Longyear (1944, pl. 9, figs. 5-8, 13, 14, 21; pl. 10, figs. 29, 30; pl. 11, figs. 1,2,4); Lothrop also illustrates some occurrences (1927, figs. 12; 13,b; 25,e). The few sherds of Ulua Polychrome found at Los Llanitos represent the southeasternmost known occurrence in El Salvador (Longyear, 1944, pp. 34-37). A counterpart to the latter find is the presence of a few sherds of Los Llanitos Polychrome at the site of Santa Ana. At Santa Ana, Santa Rita, and Las Flores the second most common painted pottery is a red and black-on-orange type known as Bold Geometric. This very distinctive pottery has a light orange paste and a very fine temper; it is harder, more brittle, and its pigments are more opaque than is the case with any of the Ulua Polychromes. Its distribution is markedly different from Ulua Polychrome. Although both categories are present in the Plain of Sula and in the Comayagua Valley, Bold Geometric ware is absent from the Copan collections, has not been reliably reported from El Salvador, and is rare at Lake Yojoa. It is found in the Bay Islands (Strong, 1948, p. 80, fig. 4) and in eastern Honduras at San Marcos (Strong, 1934, p. 47, fig. 54). The vessel which Strong illustrates from San Marcos is indistinguishable, in the photograph, from specimens from Santa Rita. Thus Ulua Polychrome has a distribution which embraces all of western Honduras and was traded to Copan and El Salvador, whereas Bold Geometric has a more easterly and northern distribution which extends north to the Bay Islands and into eastern Honduras, and it was traded less widely. Their distribution overlaps in the lower Ulua and Comayagua valleys. Enriching these factors is the presence of a type of ceramics that we interpret as a copy of Bold Geometric and that occurs in the Ulua Polychrome ware in the Plain of Sula. At Lake Yojoa, where Bold Geometric is almost absent, there is a type of pottery

which, in its forms and designs but not in other technical aspects, appears to be the cultural equivalent of Bold Geometric (fig. 6,g; Yde, 1938, fig. 45). Ulua Polychrome and Bold Geometric share few attributes and their forms and designs are completely different. Among more than 800 sherds of Bold Geometric ware from Santa Rita, less than two or three are from cylindrical vases. Variations on three basic forms account for the vast majority of Bold Geometric shapes. These are a wide-mouthed storage jar with two handles, each having a modeled effigy lug, a tripod bowl with slightly out-curving walls, and a simple bowl form. Examples of all three of these are illustrated by Strong (SKP, 1938, figs. 8, 9, 10, pl. 7; Strong, 1948,

pl. 7).

The designs of Bold Geometric include a number of recognizable animal forms such as alligators (SKP, 1938, pl. 7,c), bats (fig. 7,b), serpents (fig. 7,c,f), as well as grotesque forms. Geometric designs include frets (fig. 7,e), interlocked and cross-hachured textile designs (fig. 7,d,g), checkerboards (SKP, 1938, pl. 7,6,c), triangular designs, and, on the lower portions of large jars, radial and concentric lines. At Santa Rita the jar forms all have low collars with just one design zone (fig. 7,c,f,g) whereas the collection from Santa Ana and elsewhere includes collars that have three and more design zones or bands (fig. 7,d; Stone, 1957, fig. 49), one of which is frequently an incised red band. The same tall-collar jar form also occurs at Las Flores, where the Bold Geometric component is more poorly made and the incision often quite slovenly in execution (SKP, 1938, pl. 5,a-e). Preliminary analysis of these collections indicates that the tall- or low-collar jar may be indicative of a chronological difference within the Late Classic. The third most common decorated pottery at Santa Rita is a large class of utilitarian pottery decorated with red paint directly on the natural surface of the clay. It occurs at Santa Ana and Las Flores as 171

FIG. 7—BOLD GEOMETRIC CERAMICS, NORTHWESTERN HONDURAS, α-c, e-g, Santa Rita. d. Santa Ana.

ARCHAEOLOGY: WESTERN HONDURAS

well as in other sites in the Plain of Sula and in low frequency in collections from the sites at Lake Yojoa. It is probably related to the red-on-natural ware reported by Stone from the Comayagua Valley (Stone, 1957, pp. 36-37, fig. 22), though the examples she illustrates do not coincide with pieces in the collections from the Plain of Sula. Two types may be distinguished. The first is unpolished and its usual form is a large round jar with a low everted rim with round handles on the body around which is a sunburst-like design. In the second type the red paint is polished; simple designs, usually geometric, but occasionally including animal forms, occur. The shapes embrace a wide variety of utilitarian forms. A common and very unusual shape is a squat storage jar with an extremely tall vertical collar with a sublabial flange. Below the flange there is usually an incised decoration consisting of parallel wavy lines, comparable to specimens illustrated by Stone from the Comayagua Valley (Stone, 1957, fig. 53, C,α,α',α''). Occasionally the incision is done through a light white wash. One of the better-known remains from the lower Ulua Valley are the marble vases which are carved with elaborate scroll designs and which usually have two effigy lugs or knobs protruding from the walls of cylinder-shaped vessels (fig. 6,h). These marble vases have been the subject of a topical monograph (Stone, 1938), and Kidder (1947, pp. 36-37) has discussed the problems inherent in their dating in connection with a superficial find of one example at Uaxactun and in a late context at San Jose in British Honduras. Most of the known examples of this type of stone carving have no formal archaeological associations, but the presence of several pieces in the Santa Ana collection and the occurrence of a carved pottery type in the region, which appears to be in imitation of the marble vases, indicate that the date of these vessels is surely contemporary with the Late Classic complex.

Included in the wares associated with the Ulua Polychromes is a group of carved, modeled, and moldmade pottery vessels, usually of relatively small size. These include the type to which we have referred in connection with the marble vases. These are not, however, the same as the Ulua Marble Vaselike type that occurs in the Bay Islands (Strong, 1935, pl. 24; 1948, pp. 7879, pl. 5,d,e) and which is a much closer imitation. Several sherds of the latter type have been found at Las Flores but are not represented in other collections from the Plain of Sula. Examples of the carved and moldmade types from Northwestern Honduras have been illustrated by Strong (SKP, 1938, pls. 5,n; 64-f; H e ) . Occasionally these types are slipped in white or orange. Two types of censers are common in the collections under review. The first of these is the frying-pan or ladle censer with a perforated bottom. At Santa Rita there are examples of this type of censer in Ulua Polychrome, red-on-orange, and in red-onnatural wares. The handles are sometimes shaped into serpent, human-hand, or claw effigies. Three-pronged censers, with both plain and modeled effigy prongs, are of somewhat less frequent occurrence. Accompanying these are shallow pans with deep crisscross incisions on the exterior surfaces and with a large loop handle rising from the interior surface of the pan, which was designed to rest on the prongs. These threeprong censers are generally comparable to specimens from Guatemala illustrated by Borhegyi (1951a,b) that have been dated much earlier than our specimens. Often termed "candeleros" and considered to be censers are small cylindrical pottery containers that are common in the Santa Rita collection. These are sometimes divided, by partitions, into two or three cells. One has been illustrated (SKP, 1938, fig. 7,j). Hollow figurines, frequently moldmade, are an ever-present component of the complex. Many are whistles and take the forms of various animals and humans (SKP, 1938, 173

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

fig. 7; Gordon, 1898a; pl. 9). A large example, not a whistle, depicts a woman in an elaborate headdress holding a small human figure to her bosom (fig. 5,f). To round out this highly selective review of the Late Classic remains of Northwestern Honduras, brief mention should be made of some of the meager nonceramic artifacts from the sites that have been mentioned. From Las Flores and Los Naranjos there are barbless, copper fishhooks (SKP, 1938, p. 41; Stone, 1957, fig. 64,c), but there is little evidence to indicate that metalworking or the use of metal was characteristic of the Late Classic in the region (Strong, 1948, p. 100). The usual form of metates in the region is stated by Strong (1948, p. 101) to have been flat, undecorated, and supported by three legs. Manos were usually rectangular or cylindrical. Cylindrical, rectangular, or ovoid bark beaters and polished small and medium-sized celts made of fine greenstone help round out the assemblage. Early Postclassic As previously noted, in the discussion of the Santa Rita site, Levels 1 and 2 of Excavation 1 at Santa Rita constitute a distinct chronological deposit. Of more than 842 Ulua Polychrome sherds distributed throughout Levels 2-12, only five were found in Level 2; Level 1 contained none. A comparable situation was observed for Bold Geometric where 12 of over 813 sherds were found in Levels 1 and 2. Seven eroded and unclassified polychrome sherds out of a total of 19 such sherds fell in the first two levels. On the other hand, 57 of 58 thin buff to cream to orange-colored sherds were confined to the first two levels. The situation is less clear with respect to unslipped, coarse utility wares but the entire collection of eight polished coarse gray sherds, having the same shapes as those of the thin buff ware, were confined to the upper two levels. This thin bluff or orange ware has been identified as a type of the very widespread 174

group of X Fine Orange pottery. X Fine Orange is an Early Postclassic horizon marker in the area south of central Veracruz, northern Yucatan, and Guatemala. A summary view of the status of the various types of Fine Orange has been published by R. E. Smith (1958). Future work on the present type of Fine Orange may prove that the identification of it here made with X Fine Orange is oversimplified. The Fine Orange pottery found at Santa Rita also occurs at Santa Ana, where an additional vessel shape occurs. The type is absent to rare in all other collections I have examined. One sherd was found at Naco. Sula Fine Orange, as we may designate the type, is a thin, hard, brittle ware that is untempered, or only accidentally tempered. Minute specks of mica are visible on the surface of this pottery; mica, in fact, is usually present in wares from the Northwestern region. The type is thinly slipped, carved decoration is rare, and incision is confined to the interior upper surface of flat-bottom bowls, which suggest "grater" bottoms. One restored piece has been illustrated by Gordon (1898a, pl. 7,j). A surface find at Copan, attributed to the Postclassic, may well be a Sula Fine Orange example (Longyear, 1952, fig. 102,j). Another Early Postclassic horizon marker, Plumbate pottery, has been found at Copan but not elsewhere in western Honduras in an archaeological setting, though various scattered finds are reported from the central region (Shepard, 1948, p. 112). Whether or not it may ultimately be found associated with Sula Fine Orange cannot now be said. CENTRAL HONDURAS

The Central region of western Honduras is here considered to include the drainage of the Comayagua or Humuya River and that of the Rio Grande or Sulaco. Stone, in her archaeological survey of central and southern Honduras, has discussed the aboriginal remains of the region (Stone, 1957, pp. 47-73), devoting special attention to the

ARCHAEOLOGY: WESTERN HONDURAS

archaeology of the Comayagua Valley, particularly to sites near the present town of Comayagua (ibid., pp. 6-46). Within the compass of the present review little attention can be given to indications of archaeological complexes that have not as yet been dated or fully defined. Prominent among these is a ceramic type from Agalteca, designed as Naco Style ware by Stone (ibid., pp. 67-68), which suggests that a Postclassic phase will be defined in the region. In the Comayagua subregion two groups of sites deserve mention. The first of these are the valley sites, among which Las Vegas and Yarumela are the best reported. Las Vegas is a mound site which covers an area at least 3 miles in circumference, with the central part consisting of approximately 60 rectangular house mounds with a smaller number of rectangular ceremonial mounds at the southern end. The average house mound is described as about 16 by 30 m. at the base and varying in height between 1 and 1.5 m. Retaining walls are made of stones and the floors of adobe (ibid., pp. 15-17). In addition to ceramics typical of the Ulua Polychrome complex, there is a black, red, and orange-on-white type which is not a member of the Ulua Polychrome group (ibid., figs. 4, 44, and frontispiece). Stone metates with carved effigy heads and three legs, usually associated with Central America or eastern Honduras, are also reported at the site (ibid., fig. 43, A, B). Yarumela is distinguished by a large rubble mound, apparently Formative in date, that is some 20 m. high, 165 m. long, and 100 m. wide. Canby's excavations at the site, of which only a summary has appeared in print (Canby, 1951), yielded a long sequence with three Formative phases and the Ulua Polychrome complex at the periphery of the site. The first, Eo-Archaic, phase exhibited a general absence of such modifications as handles, feet, and spouts. Red paint was occasionally used but no vessel had any formal design. Rim profiles of near-

ly flat, sideless plates showed considerable elaboration of the treatment of swollen lips. Pattern-burnishing, bichrome sherds, handles, and a greater number of shapes appeared in the Proto-Archaic phase. The Archaic phase is described as identical with the Ulua Bichrome phase, defined at the site of Santa Rita in the Plain of Sula. In the mountains surrounding the Comayagua Valley are several important ruins which are not only situated in naturally fortified locations but show evidences of manmade fortification. Three of these sites, Calamuya, Quelepa (which has a ball court), and Tenampua, are described by Stone (1957, pp. 47-56). Even though the artifact assemblages from these sites are incompletely reported, their variation in both site location and ceramics suggests that the hill sites may be later than the valley sites, where the ceramics fall into the Late Classic complex described for the Northwestern region. The largest of the hill sites is Tenampua, which was described by Squier in the last century and by Dorothy Popenoe in 1936. The site is on a mesa some 400 m. above the Comayagua plain. Numerous stonefaced mounds, some in rectangular and formal relationships, abound; included among them is a ball court. Stone (1957, p. 53) has discussed the known artifacts from the site, which include a carved stone metate of the elaborate Costa Rican type (Popenoe, 1936, fig. 4). Excavations in 1955 (Stone, 1957, p. 50) have resulted in the isolation of a new polychrome type which is in the Ulua Polychrome group (fig. 8; Stone, 1957, figs. 55, A, B, C; 57, A). The globular, handled bowl, without shoulders and with an incised decoration on the collar (fig. 8,f), appears to be restricted to Tenampua Polychrome; other attributes, such as a loop design (fig. 8,e) and a characteristic wavy white line on the interior wall of open bowls and cylinder vases (Popenoe, 1936, fig. 2), enable it to be segregated positively from other Ulua Poly175

FIG. 8—POLYCHROME CERAMICS FROM TENAMPUA, CENTRAL HONDURAS

176

ARCHAEOLOGY: WESTERN HONDURAS

chrome types. The processional human figures (fig. 8,a-c) are in the tradition of the Santa Ana class of Ulua Polychrome, though both Mayoid (fig. 8,f) and naturalistic (fig. 8,e) designs occur. A common foot form, consisting of a coiled loop (fig. 8,d), was first reported by Squier. One sherd of Tenampua polychrome, without provenience as to level, is in the Santa Rita collection; I know of several examples found in Nicaragua. A large sherd collection from the site, excavated by the Willauer party, and examined by me, contained no pieces typical of the Ulua Polychrome complex; but the demonstration that the Tenampua assemblage is later than the Late Classic complex of the Northwestern region and the Comayagua Valley must await more detailed examination of the question. SOUTHERN HONDURAS

The Southern region of Honduras, as described by Stone (1957, pp. 82-102), lies between the northern reaches of the Choluteca River and the Gulf of Fonseca. It includes the valleys of the Goascaran and Nacaome rivers, about which we have no archaeological information, the islands in the Gulf of Fonseca, the Tegucigalpa area, and the Choluteca Valley and flood plain. Ulua Polychrome pottery is reported at various sites in the region, including the islands in the Gulf of Fonseca. At the site of La Ola, near Choluteca, it is common (ibid., p. 98) as it is at some sites in the Tegucigalpa area, but it is rare to unreported at others. A red-on-cream or white pottery, Tegucigalpa ware (ibid., fig. 31), is the most consistently reported ceramic type in the region, it occurs also near Talanga, in central Honduras, where it appears in almost equal quantity with Ulua Polychrome (ibid., p. 66). At such sites as Palo Blanco and Los Calpules, both near Tegucigalpa (ibid., pp. 92-97), Ulua Polychrome is either rare or absent and Tegucigalpa ware and its subtypes are of common occurrence. Stone does not, however, draw

any chronological conclusions from these data, and the assemblages which include Tegucigalpa ware are undated but may well prove to be somewhat later in time than the Ulua Polychrome complex. SOUTHWESTERN HONDURAS

The Southwestern region of western Honduras corresponds to the modern Departments of La Paz, Intibuca, Gracias, and Ocotepeque that border on El Salvador to the south. The scanty archaeological material from this mountainous region has been summarized by Stone (1957, pp. 10319). The site of Cerquin is a fortified location on the summit of a mountainous peak and, in some respects, is reminiscent of the site of Tenampua. Nearby is a small site with low stone and earth mounds. The site of Sensenti is located near a river that ultimately flows into the Ulua. Here, terraced rectangular mounds are placed formally around courts. Stone reports that these are built like those in parts of the Comayagua Valley. Ceramics from the site include Ulua Polychrome. CHRONOLOGICAL REVIEW

The relative chronology of the phases and ceramic complexes of western Honduras are shown in Table 1. Two manifestations are omitted. The first is a nonceramic stone assemblage found at Copan (Longyear, 1948). The second is represented mainly by the Tenampua type of Ulua Polychrome found at the site of that name in central Honduras. The Tenampua material is in the tradition of the Late Classic complex of the area but may well be more than a regional variant. A later date is suggested for the complex, not only on account of its variation, but because of the generally late date, in Mesoamerica, for the appearance of fortified sites. At the present time, however, these are insufficient reasons for classifying the complex in a different period from the related Ulua Polychromes. The relative chronology of the Formative 177

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

TABLE 1—RELATIVE CHRONOLOGY OF WESTERN HONDURAS

PERIODS Postclassic

WESTERN HONDURAS Far-western

Northwestern

Formative*

HONDURAS

Naco Plumbate (pottery)

Fine Orange (pottery)

Full Classic Classic

EASTERN Central

Cocal

Ulua Polychrome complex

(Copan)

Selin

Early Classic (Copan) Archaic (Copan) Copan Caves

Ulua Bichrome Playa de los Muertos Yojoa Monochrome

Archaic ProtoArchaic Eo-Archaic

°The sequence of phases within the Formative period is tentative; those from the Central region of western Honduras are based on the excavations at Yarumela.

phases of western Honduras is still a matter for further research. The elaborate material from Playa de los Muertos has relationships to other Mesoamerican Formative cultures, especially to the material from Tlatilco in the Valley of Mexico (Sorenson, 1955, p. 46), and more particularly to the middle phase of the Formative sequence established for central Mexico (Pina Chan, 1955, p. 113). The complexity of correlating the Formative phases of western Honduras is evident in the discussion of the Mesoamerican Formative presented by MacNeish (1954, pp. 617-25). Despite the positive knowledge represented in Table 1, and reviewed in the present article, the chronological picture of western Honduras has important gaps. Remains attributable to the Lithic or PaleoIndian period have not been found, except for the find at Copan; part of this may be due to the preoccupation of past research in the area with the polychrome ceramics and the very deeply buried nature of all deposits in the Ulua Valley. The absence of any complex belonging to the Early Classic period is strange, but I feel that this is due 178

more to the lack of identification of undated materials than to the absence of occupation of western Honduras (with the exception of the Far-western region) during the Early Classic period. Future research may well indicate a survival of Late Classic elements into the Postclassic, perhaps with the disintegration of the polychrome styles. At this time the presence of Early Postclassic horizon markers, typical of southern Mesoamerica, such as Tohil Plumbate and the Fine Orange that we have described, begin to appear in western Honduras. The development of fortified sites might well be related to these events of the Early Postclassic. The lack of the identification of a Late Postclassic or contact complex in any region of western Honduras, other than the Farwestern, is a particularly important gap in the archaeological knowledge of the area. The correlations of archaeological material with Indian groups known to have been in the area at the time of contact all suffer from this deficiency. Such a basic question as who made the Ulua Polychromes is still an open one on account of the difficulty of

ARCHAEOLOGY: WESTERN

HONDURAS

equating tribes in the area in the 16th century with complexes made as much as 500 or more years before. The stimulating essays in this direction made by Stone in numerous articles (see her 1957, bibliography) probably indicate the trend that such identifications will take. The more con-

servative appraisal of these problems by Longyear (1947) indicates, however, that the Lenca may well have occupied the area of the distribution of Ulua Polychrome in Classic times, with the Jicaque replacing them in the more northern portions of Northwestern Honduras.

REFERENCES Blackiston, 1910 Boggs, 1943b, 1945a, 1945b Borhegyi, 1951a, 1951b Canby, 1949, 1951 Coe, M. D., 1957b Epstein, 1957, 1959 Gordon, 1898a, 1898b Kidder, 1947, 1949 Kirchhoff, 1943 Longyear, 1944, 1947, 1948, 1952 Lothrop, 1926, 1927a MacNeish, 1954 Piña Chan, 1955 Popenoe, 1934, 1936

Porter, Μ. Ν., 1953 Richardson and Ruppert, 1942 Shepard, 1948 Smith, R. E., 1958 Sorenson, 1955 Steinmayer, 1932 Stone, 1938, 1941, 1957, 1959 Strong, 1934, 1935, 1948 , Kidder, and Paul, 1938 Tax, 1951 Vaillant, 1934 Wauchope, 1948, 1950 Woodbury and Trik, 1953 Yde, 1938

179

9. Archaeology of Lower Central America

S. K.

W

HEN THE Spaniards entered the lands between Honduras and Colombia, they encountered many local chiefdoms, none of which exercised political power over an extensive area. This condition is reflected by archaeological finds today. There are no indications of large and compact communities or of any permanent architectural remains except foundation platforms of earth, sometimes faced with walls of uncut stone. On the other hand, there were notable developments in aboriginal stone carving, ranging from large statues to jewelry of semiprecious stones. Several important metallurgical centers existed. Literally tens of thousands of pottery vessels have survived and can be seen in both public museums and private collections, the most important of which have been listed by Henri Lehmann (1959). This great mass of material obviously represents many centuries of stylistic growth, but comparatively little has been ascertained about archaeological sequences, and even less has been published. Lower Central America has been regarded as a country cousin of the more publicized Mayan and Mexican areas to the north. 180

LOTHROP

Intrusive ideas and influences from both north and south indeed are recognized, but the basic art forms and styles of lower Central America are deemed to be of local and independent origin. More specifically, the polychrome pottery exhibits regional aesthetic characteristics both in the quality of line and in the display of fired pigments. Colors such as blue, purple, chocolate brown, gray, green, and orange must be the result of local experiments and discoveries. On the other hand, Panama and Costa Rica probably made a major contribution to Mexico—almost the entire metallurgical complex, including alloys and technical processes. EARLY MIGRATIONS

Passage of Palaeo-Indians through Central America long ago has been postulated to explain the association of man and extinct animals in South America. Human footprints in solidified volcanic mud flows, now deeply buried by subsequent deposits, have been found at two localities in El Salvador (Haberland and Grebe, 1957) and at least five in Nicaragua (Lothrop, 1926, ch. 2). Discovered originally over 75 years ago, the

ARCHAEOLOGY: LOWER CENTRAL AMERICA

Nicaraguan footprints became the subject of controversy during the past century but were forgotten until F. B. Richardson (1941) uncovered a new set of tracks. The geological implications have been studied by H. Williams (1952), who recognized an associated deer jaw as a type now found in Mexico rather than in Nicaragua. Footprints of bison have been identified, but there is no clue to the date of existence or extinction of this animal in Nicaragua. Palaeo-Indian fluted points have been identified in Costa Rica (Swauger and Mayer-Oakes, 1952) and Panama (Sander, 1959). Radiocarbon dates with comparable associations in South America suggest that men had passed through the Isthmus to reach that continent more than 12,000 years ago. There is no indication, however, of a permanent Isthmian settlement earlier than 4840 ± 110 B.C. NICARAGUA

Most recorded archaeological finds come from the region adjacent to Lakes Managua and Nicaragua, i.e. from the Pacific slopes, which, both in ancient times and today, have contained the bulk of the population. Recent field studies by Haberland, Norweb and Willey, Richardson and Ruppert, and Termer, all still unpublished, may be expected to throw new light on cultural and stylistic sequence, which now can only be elucidated through associations of native and European artifacts (Bransford, 1881, 1882; Flint, MS.) or by identification of trade pieces and stylistic extensions from other areas. SCULPTURE. Numerous stone statues, ranging up to 4 m. in height, have been discovered on the shores and islands of Lake Nicaragua. As no photographic record is available, our knowledge is based on drawings, chiefly those published by Squier (1852), Bovallius (1886), and Richardson (1940). The distribution is given by Lothrop (1926, p. 93, app. I ) . Several distinct types can be recognized, and stylistic affilia-

FIG. 1—STONE STATUE. Chontales, Nicaragua. (After Richardson, 1940.)

tions with areas to both north and south have been suggested. As yet, however, the chronological position and sequence of these carvings remain vague. East of the lake, statues are basically cylindrical (fig. 1) and represent men or deities, sometimes with elaborate details rendered in low relief. They suggest a tree trunk which has been only slightly modified in shape. Neither in concept nor in symbolism is there any suggestion of Mexican or Maya influence, but they may be regarded as vaguely South American. Statues from the islands and the so-called Isthmus of Rivas to the west of Lake Nicaragua are better known. Typically they consist of a round or square column surmounted by a seated or standing human 181

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

figure, often with an animal covering the human head and shoulders. This concept, known as the alter ego motif, occurs in both Mesoamerica and South America. The carving, however, is in the three-quarters or full round, with limbs sometimes freed from each other and from the body, along with an attempt at anatomical accuracy. These traits imply northern affiliations and are definitely not characteristic of South American carvings. METALWORK. The first Spaniards to overrun southwestern Nicaragua secured gold ornaments valued at over 70,000 pesos, but archaeological finds of metal have been negligible. Aboriginal gold must have reached Nicaragua by trade from Costa Rica or Panama. It evidently was not buried with the dead but was cherished for generations above ground. No local styles have been detected and there is no indication of local manufacture. POTTERY. Probably no area discussed in this volume has yielded so much pottery about which so little is known as Nicaragua. Although many examples of types now in museums have been published by Lothrop (1926), his classification of styles obviously needs amplification. Present knowledge of associations and sequence still depends largely on Bransford's excavations (1881), buttressed by comparisons with recently discovered stratigraphy in Costa Rica and by identification of a limited number of trade vessels. Bransford, who had no archaeological training, nevertheless was the first observer to define and to date a ceramic type in Central America. This he named Luna ware, and he demonstrated that it existed until the 16th century because examples were found with European artifacts such as glass beads. Today it is known that Luna ware was manufactured in northwestern Costa Rica as well as in Nicaragua. It stands apart from other ceramic types in some symbolic concepts. It is recognized that Luna ware in182

cludes several qualities of line which may represent chronological phases (fig. 9 ) . Bransford also defined a ceramic type which he called Santa Helena ware. This is a brilliant aspect of the Polychrome wares found both in Nicaragua and Costa Rica (Lothrop, 1926, pls. 30, 36, 57, 72). The decorative motifs in part are linked with Late Classic Maya and they can be placed in the Costa Rican stratigraphic sequence as Middle Polychrome (Table 1). Another pottery type isolated by Bransford in Nicaragua is known as Palmar ware, distinguished by broad incised lines (Lothrop, 1926, pl. 114). Examples of this occur with the Zoned Bichrome in Costa Rica (fig. 3,b). Other contemporary styles include certain aspects of a miscellaneous pottery group found near Nandaime (ibid., pl. 97,b-d; figs. 111, 113). Among these are a White-line on Red ware (fig. 3,h), a group with simple patterns outlined by ridges (fig. 4,g) and grater bowls with incised floors set on globular tripod legs. Several pottery styles of northern origin reached Nicaragua by trade. Usulutan ware has been found near Managua by Richardson and on Ometepe Island by Haberland (verbal information). Examples of Maya Copador and Tohil Plumbate have been published by Lothrop (ibid., figs. 280, 281). A small flask of Ulua Valley type with molded Maya glyphs was discovered by Flint. Nicaraguan figurines are discussed with those from northwestern Costa Rica. COSTA RICA

In spite of its small size, the Republic of Costa Rica has produced an enormous amount of archaeological material, much of which has been illustrated by Hartman (1901, 1907), Walter Lehmann (1908, 1913), Lothrop (1926), Mason (1945), Stendahl (1952), and Stone (1958). For many years, the chronology was limited to discoveries of European artifacts with aboriginal remains by Bransford (1882) and

ARCHAEOLOGY: LOWER CENTRAL AMERICA

FIG. 2—JADE PENDANTS. Nicoya peninsula, Costa Rica. (After Lothrop, 1926).

Hartman (1901) and to stylistic speculation, but recently stratigraphic sites have been opened by Baudez, M. D. Coe, Haberland, Lothrop, and Stone. Although preliminary announcements have been made, this material is not yet available in detailed published form. Costa Rica is marked by strongly developed local archaeological styles, which, however, tend to blend or to pass from one region to another in trade. The use of postconquest linguistic terms to define archaeological finds has complicated the picture. We divide the country into three major geographical areas: (1) Northwest, also known as the Nicoya region, which is chiefly allied with southwestern Nicaragua; (2) Northeast, which includes the Meseta Central and the Atlantic plains to the east, often spoken of as the Guapiles area or the Linea Vieja (on account of the railroad line); and (3) South, which is best considered in relationship with Panama. Each area calls for subdivisions. Northwestern Costa Rica This area includes the lands fronting on the Gulf of Nicoya, the peninsula of the same name, the so-called Isthmus of Rivas

and adjacent islands in Nicaragua. Archaeological sites are listed in Lothrop, 1926, app. I. Aboveground archaeological remains consist of stone statues of Nicaraguan styles (Stone, 1958, Nicoya, fig. 7), grave markers in the form of stone columns, stone circles or small stone mounds, and refuse mounds which may outline the shape of village plazas. STONE ARTIFACTS. Apart from a few statues of Nicaraguan type, the area is marked by elaborate metates, usually with three legs which may be round or triangular in section (Hartman, 1907). Many have complex designs on the base, which suggests that the metates were placed on end when not in use (Stone, 1958, p. 34, fig. 2). Elaborately carved circular stone seats are hourglass shape in outline (Baudez, 1959). Smaller artifacts include club heads (usually carved to represent a bird or animal head), stirrup pestles, bark beaters, and mirror backs. Nicoya is famous for its "jades," several hundred examples of which have been published by Hartman (1907). The material has been identified as diopside-jadeite, chalcedony, and other softer stones. Among the most elaborate forms are the so-called axe 183

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

TABLE 1—CHRONOLOGICAL SEQUENCE IN NORTHWESTERN COSTA RICA

(After Baudez and Coe)

Middle Tempisque Area

Dates (estimated from radiocarbon tests)

Period

Chahuite Escondido

Late Polychrome Β Late Polychrome A

La Cruz Β La Cruz A

Middle Polychrome

Doscientos '

Tamarindo

Palo Blanco

Early Polychrome Β

Santa Elena

Matapalo

San Bosco Β

800-450

San Bosco A

450-100

Tamarindo

1550-?

Early Polychrome A Zoned Bichrome

Chombo

Monte Fresco

gods, carved on what appears to have been a celt cut in half (fig. 2 ) , and winged pendants representing bats. Foreign jades found in Nicoya include two Olmec and one Classic Maya example. METALWORK. In 1522 a Spanish expedi-

tion traveled overland along the shores of the Gulf of Nicoya and then northward to Nicaragua. An official report (Cereceda, 1926) shows that gold ornaments valued at over 30,000 pesos were secured from the natives. Modern excavations have produced only a handful of gold artifacts. There is no indication that a local metalworking industry ever existed, but aboriginal trade evidently introduced a considerable amount of jewelry just before the conquest. TEXTILES. NO fragments of cloth or basketry have survived, but what appear to be textile motifs may be seen on pottery vessels and figurines. After the conquest, this area paid tribute in cotton dyed with Pupura patula. POTTERY. In a short article it is not possible even to summarize the infinite variety of ceramic types found in northwestern Costa Rica. Published sources have been listed above. Here we discuss chiefly those aspects which have been placed chronologically, mostly as result of 1960 stratigraphic excavations in the Tempisque Valley by Claude Baudez (1962) and on the 184

Catalina

A.D. 100-300 B.C.

Santa Elena peninsula and Tamarindo Bay by M. D. Coe (1962). Both presented stillunpublished papers at the 34th International Congress of Americanists and also a joint paper correlating their results (Baudez and Coe, 1962), which are shown in Table 1. Zoned Bichrome. This period is marked by Red and Black on Buff or Black on Red styles with the colors separated by fine-line incising. Although the type is widely distributed, it is known chiefly through sherds and a few complete vessels (fig. 5,a,c). A second kind of incising in Nicoya has broad parallel lines (fig. 3,b) recalling part of the Palmar ware of Nicaragua (Lothrop, 1926, pl. 114), about which little is known. Incised and white-filled rim patterns, often in the form of hatched triangles, make their appearance at this time (fig. 3,c). Also there are appliqué and punctate motifs which have yet to be classified. Painted decoration consists of parallel zigzag lines, apparently applied with a multiple-brush technique (fig. 3,e) which persisted into the Middle Polychrome period. The early painted styles are Black on Red, Black on Buff, or Red on Buff. A radiocarbon date (Y-810) of A.D. 90 ± 200 has been announced, and the estimated span of the period is 300 B.C.-A.D. 100. Early Polychrome A. The Black- or Red-

FIG. 3—STRATIGRAPHIC SHERDS. Nicoya peninsula, Costa Rica, α-e, Zone Bichrome period (Catalina and Monte Fresco), f-i, Early Polychrome (Santa Elena and Matapalo), j-m, Middle Polychrome (Doscientos and Tamarindo). Width 5-18 cm. (Photos, Baudez and Coe.)

FIG. 4—MISCELLANEOUS NICOYA POTTERY TYPES. a,c, Late Polychrome, e, Late incised. b,f,i, Middle Polychrome. d,h, Early Polychrome, g, Zone Ridged. Height of i, 26 cm. (Photos, Baudez and Coe; Museum of the American Indian, Heye Foundation; Peabody Museum, Harvard University.)

FIG. 5—MISCELLANEOUS POTTERY. a,c, Zone Bichrome. b, Black- and White-line, d, Black-line, c, Guacatepe, Nicaragua; others, Costa Rica. Diam. 11-25 cm.

line multiple-brush technique persists (fig. 3,e), also the White-filled incised patterns, which assume more elaborate forms. The first sherds with color-filled black outlines make their appearance. Early Polychrome B. Although the older painted forms continued, many innovations appear. Among these are White-line on Red vessels (figs. 3,h; 4,g), previously recorded in the Nandaime Region in Nicaragua (Lothrop, 1926, pl. 97,c,d). Also there are bowls adorned with vertical black lines or with black-line patterns flanked by white (fig. 5,d,b). A new type consists of shallow grater bowls with incised floors, set on

bulbous tripod legs (Lothrop, ibid., figs. 111, 113), also found in the Nicaraguan Nandaime region. In addition, there are jars with decoration outlined by small ridges (fig. 4,g). A radiocarbon date (Y-811) of A.D. 565 ± 90 has been obtained and the estimated time span is A.D. 450-800. The major stylistic change is the development of what has been regarded as typical Nicoya Polychrome. The largest group which can be isolated consists of boldly executed designs in black, partially filled with red (figs. 3,f,g; 4,d,h). Although many patterns are purely geometric, in some instances they may share a common an187

FIG. 6—EARLY POLYCHROME MOTIFS. Nicoya peninsula, Costa Rica. Diam. 7.5-12.5 cm. (After Lothrop, 1926.)

ARCHAEOLOGY: LOWER CENTRAL AMERICA

cestry with the well-known Chiriqui Alligator motifs and other dragon-like designs (Lothrop, ibid., pl. 75, figs. 69-75). The full range of motifs which first appeared in this period has yet to be ascertained, but we should mention the design seen in figures 3,i, and 6, which may be of Classic Maya inspiration, both in the quality of line and the depiction of a jaguar with large black spots. The black background showing the animal in silhouette is characteristic of the Nicoya area (figs. 4,i; 7). In addition to pottery vessels, cylindrical stamps and "napkin-ring" ear spools occur in this period, also metates with ornamented bases. The well-known stone material excavated by Hartman (1907) at Las Huacas near the town of Nicoya, including many jade artifacts, is assigned in part to an early phase of this epoch, but most of the pottery has never been published and cannot be classified yet. Middle Polychrome. The aboriginal population, to judge by the number of archaeological sites, probably was at a maximum during this period and a correspondingly great proportion of the pottery will ultimately be assigned to it. Among the types which can now be placed here, at least in part, are elaborately incised vessels with designs filled with white pigment and bridge-spout effigy jars. Painted forms include tall effigy jars (fig. 4,b,i), the group known as Culebra ware (Lothrop, 1926, pp. 107-08), many zoomorphic motifs, and also designs apparently derived from textiles (ibid., pls. 27; 57; 60,c; 8l,d,e; figs. 38, 60, 79). Baudez and Coe emphasize the diagnostic importance of a bar-and-dot motif, often associated with crosses suggesting the kan cross of the Maya area (fig. 8). The former occurs in the Costa Rican highlands and to the south in the Diquis delta, thus providing evidence for cross-dating. The design shown in figure 3,m is primarily associated with the highlands but can be placed in this period in Nicoya.

FIG. 7—EARLY POLYCHROME JAGUAR MOTIFS. Nicoya peninsula, Costa Rica. (After Spinden, 1917a; Lothrop, 1926.)

Examples of Middle Polychrome were traded to the northwest as far as El Salvador, Guatemala, and Honduras and have been found in association with Tohil Plumbate and X Fine Orange pottery. Late Polychrome. The last ceramic period in northwestern Costa Rica may be divided into two phases, both characterized by monochrome pottery, often of complex profile, adorned with broad grooves (fig. 4,e). The black pottery admired by Oviedo (bk. 42, ch. 12) first appears in this group. The contemporary painted ware (fig. 4,a,c) is marked by the use of blue-gray pigment (Lothrop, ibid., pls. 24, 25, 64), sometimes associated with incised designs covered by a light slip (ibid., ch. 8). The last phase is characterized by Luna ware (fig. 9), first analyzed by Bransford (1881), who demonstrated contemporaneity with European artifacts. Several subtypes now can be recognized, in both Costa Rica and Nicaragua, but their chronological significance has not been determined. Figurines. Many varieties of figurines have been found in Nicoya and Nicaragua, some of which may be placed chronologi189

FIG. 8—MIDDLE POLYCHROME BOWLS. Nicoya peninsula, Costa Rica. (After Lothrop, 1926.)

ARCHAEOLOGY: LOWER CENTRAL AMERICA

FIG. 9—STYLISTIC VARIANTS O F LUNA WARE. Nicaragua. Length of b, 10 cm.

cally. Oldest are a type with zoned bichrome incising. Second is a group of hollow, handmade figures with details of face and breasts in high relief enhanced by painted detail (fig. 10). Third is a type made in molds with minimum modeling but details indicated in paint. Illustrations of these styles are found in Stone, 1958, Nicoya, figs. ll,d; 9,0; 11,e. Figurines probably imported from Chiriqui (Lothrop, 1926, pl. 129) and Luna ware figurines (ibid., pl. 128) are of later date. Types primarily associated with the Ulua Valley in Honduras also are found (ibid., pl. 132,f,g; fig. 167), in both Nicaragua and northern Costa Rica. Northeastern Costa Rica The archaeology of the central highlands and Atlantic plains of northeastern Costa Rica is known through the field investigations of Hartman (1901) and Skinner (1926), as well as illustrated ceramic material selected from the approximately 15,000 specimens collected by Minor C. Keith and now in the Museum of the American Indian, Heye Foundation (Lothrop, 1926). In addition, many artifacts have been published by Stone (1958); and the large collection of sculpture made by Keith,

FIG. 10—HEAD OF EARLY POLYCHROME FIGURINE. Liberia, Costa Rica. Height 7.5 cm.

191

FIG. 11—POTTERY FROM NORTHEASTERN COSTA RICA. Three top rows: Curridabat ware. Bottom rows: Stone Cist ware. Diam. 8.75-17.5 cm. (After Lothrop, 1926.)

ARCHAEOLOGY: LOWER CENTRAL AMERICA

which is now in the American Museum of Natural History and the Brooklyn Museum, has been published by Mason (1945). Products of a jadeworking center near Guapiles have been described by Balser (1953) and Lothrop (1955). Local metalwork has been identified by Stone (1951) and by Stone and Balser (1958). Additional sources are given in those here listed. CONSTRUCTIONS. No architectural remains have been recorded, but earth mounds faced with river boulders are known in the lowlands. At times these are elongated and form enclosures. Other mounds are circular and ascended by ramps which may connect with paved roads (Skinner, 1926, fig. 289). Similar roads have been found both on the coast and on central plateau (Stone, 1958, Atlantic Watershed, fig. 25). Graves usually are stone lined with walls either of river boulders or stone slabs. SCULPTURE. Stone statues have been found in quantity, ranging up to 1.5 m. in height. Both men and women are represented, carved in the full round but with little attempt at anatomical accuracy. Metates are common. There is a small, fourlegged effigy type with a raised rim and also a very large three-legged form with a flying panel attached to one leg (fig. 15,bd). Carved jades from the Guapiles region often exhibit the string-sawing technique. A remarkable find at Retes in the highlands was a cache of wooden objects— carved tablets, circular stools, drums, and staffs (Aguilar, 1953). A radiocarbon date of A.D. 990 has been obtained (Stone, 1958, p. 19). METALWORK. Many objects of gold or gold-copper alloy have been found (Stone, 1951). Although some are the products of trade with Panama and beyond, local centers of manufacture are attributed to the Reventazon valley and the Guapiles region. These are the northernmost-known areas of manufacture representing the Isthmian metallurgical traditions. POTTERY. Polychrome decoration is rela-

tively unimportant in this area, although definitely local motifs are recognized. Simple patterns painted in a single color— white, black, red, or yellow—are found. There also are incised vessels related to types known in Nicoya and the southeast. The major part of the material, however, has appliqué and punctate decoration, sometimes combined with the use of color (fig. 11). Most vessels are small. We omit discussion of various classificatory schemes, as the groupings are still without chronological position—with the exception of types associated with European artifacts. Isolated finds lacking exact provenience are of Nicoya styles which indicate a time span fully equal to that now known in Nicoya, i.e. at least 1500 years (Stone, 1958, Atlantic Watershed, fig. 20,c). A stratigraphic sequence in the Guapiles area has been mentioned but not revealed in detail (ibid., pp. 22, 23). Figurines are not so common as in Nicoya. Several local forms are found and also types apparently imported from Honduras and Panama (Lothrop, 1926, figs. 263-72). We should add that, apart from northeastern Costa Rica, no archaeological remains of importance have been reported on the Atlantic slopes of lower Central America or Panama. SOUTHERN COSTA RICA AND WESTERN PANAMA

From an archaeological point of view it is best to disregard present political boundaries and to consider the remains in this region as a whole, marked by widely distributed pottery types as well as by specialized local developments in pottery and sculpture. Although the total area is small, the complexity of archaeological finds is great and must represent population shifts over a period of many centuries. Our present picture of the past is changing rapidly as the result of field research. A century ago, aboriginal gold ornaments 193

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

FIG. 12—DIQUIS DELTA, COSTA RICA, α-c, Early variants of Chiriqui Alligator ware, d, Stone statue. e, Early Brown Incised ware. (Photos, Peabody Museum, Harvard University.)

from graves in the Province of Chiriqui, Panama, began to reach Europe and the United States (Haberland, 1960a; Lothrop, 1919). Since that time, ancient burials have been relentlessly looted, and, although in the past most gold artifacts were reduced to bullion because they had no commercial value, pottery vessels and stonework found their way to museums in huge quantity. The magnificently illustrated monographs of Holmes (1888) and MacCurdy (1911), based on such material, revealed a surprising variety of types but gave no clue to their development. This picture of "Chiriqui culture" remained static until Osgood (1935) published a new analysis which demonstrated that most of the pottery wares defined by Holmes and MacCurdy were not of local manufacture but had been imported from unknown sources. Haberland, 194

the most recent investigator, now recognizes five archaeological subareas in the border region: Boruca, highland and lowland Chiriqui, the Osa peninsula, and Diquis delta. Each may be broken down into chronological periods. DIQUIS DELTA. Actual stratigraphy has been found only in the Diquis delta (Lothrop, 1963). Here well-known "classical" Chiriqui types of pottery and figurines, such as Alligator ware, have been established as terminal phases in a developmental series (fig. 12,a-c). Of even greater age were large crude incised bowls (fig. 12,e) and jars measuring up to a meter or more in diameter. The oldest pottery, at a depth of about 2 m., was undecorated except for a Fugitive Red ware, apparently painted after firing. Trade pottery from northern Costa Rica and the Asuero-Cocle region in

FIG. 13—GOLD AND POTTERY. a,b, Chiriqui gold (Museum of Fine Arts, Boston). c, Veraguas gold (Museum of Fine Arts, Cleveland). d, Cocle-style gold, Sona, Veraguas. Height 5 cm. e, Sitio Conte gold (Museum of Primitive Art, New York). Diam. 17.5 cm. f , g , Veraguas pottery, Bubi, Sona. Diam. 25, 18.5 cm. (Peabody Museum, Harvard University.)

FIG. 14—POTTERY AND STATUE, α-c, Asuero pottery types from Veraguas graves. Diams. 27, 28, 32 cm. d,e, Statue from Barriles, Chiriqui, and Barriles-type mask from Veraguas. Height 220 and 21 cm. Museo Nacional, Panama. (Photos, H. A. Dunn and R. H. Mitchell.)

Panama was found in the upper refuse but could not be accurately placed in the sequence. The Diquis delta has produced a large number of gold ornaments, many of which may be of local manufacture (Stone, 1958; Stone and Balser, 1958). Stone statues (fig. 12,d) of many types have been described by Mason (1945). A unique feature is the great stone spheres which measure up to 2.5 m. in diameter and often occur in large groups (Stone, 1943). Several hundred examples are known. BORUCA. In the Boruca area, Haberland found two distinct ceramic complexes (1959a,b). His Buenos Aires group prob196

ably is contemporaneous with the polychrome development and associated incised and appliqué types from the Diquis delta about 50 km. to the south. The Aguas Buenas complex includes zone-incised pottery as well as Scarified ware, and evidently is of considerable age. Scarified ware (fig. 16,a) has recently assumed new importance owing to its association with a radiocarbon date and to added knowledge concerning its distribution. It was originally named by Holmes (1888), who remarked that it was unlike any other Chiriqui ceramic style. The first recorded specimens came from the vicinity of David and, as early as 1890, the fact that

ARCHAEOLOGY: L O W E R CENTRAL AMERICA

they were not found with painted pottery had been noted. In the 1930's and 40's, many specimens were unearthed in the lowlands behind Puerto Armuelles. Photographs were circulated but not published. About five years ago, the known range of Scarified ware was greatly extended owing to its discovery at Aguas Buenas in Costa Rica (Haberland, 1959a), on Guacamayo volcano in Cocle (Ε. Μ. Harte, 1958, and verbal communication) and at Pueblo Nuevo on the Chiriqui-Veraguas boundary (Feriz, 1959, and data from K. P. Curtis). In each instance, there have been associated pottery types not yet fully appraised. At Pueblo Nuevo, a radiocarbon date of 230 B.C. ± 60 was obtained (DeVries, 1958). Haberland (1960a; in press a) has shown, however, that there is stylistic variation within Scarified ware and a considerable time span may be involved. BARRILES. A truly astonishing discovery in the middle of "classical" Chiriqui territory was the Barriles site where a previously unknown type of sculpture was found (fig. 14,d,e) (Haberland, 1960c). Examples now in the Museo Nacional of Panama depict a slave carrying his master on his shoulders. Also there are ceremonial metates over 2 m. in length. Pottery now in the Panama Museum is bichrome with fine-line zone incising. Haberland (1960a, p. 13) also reports pottery types which are "an exact duplicate of the Aguas Buenas material." Other important recent finds have been made on the Osa peninsula in Costa Rica, at La Concepcion and Villalba in Chiriqui, and near the Virali river where a steam shovel produced intact several large polychrome urns. Discussion must await the publication of pertinent data. Finally we should mention the Chiriqui Red-line ware of Holmes and MacCurdy, adorned with simple elements which often were purposely smudged while the paint was wet. Vessels with comparable decoration were manufactured in the Asuero peninsula, Cocle, and eastward as far as the

Panama Canal. Although found with Blackline and early Polychrome pottery, Red-line ware apparently persisted for centuries and is not indicative of place or period. Petroglyphs have been discovered in quantity in Chiriqui and other provinces to the east (N. A. Harte, 1960). VERAGUAS

In the 1930's when the Pan American Highway was constructed across the Pacific slopes of the Province of Veraguas, goldbearing graves were encountered and a distinctive archaeological region with many features not found in the adjacent provinces was revealed. Although thousands of graves have been ransacked, almost no recorded digging has been done. Hence, we know what has been found but not how it was found and in what associations. Only two habitation sites are known to me, and no attempt has yet been made to find stratigraphy. The Veraguas area, however, was one of unusual commercial activity. The chief export was cast-gold jewelry, which reached Chiriqui and the Diquis delta in quantity or was copied locally. Examples have been found as far north as the Cenote of Sacrifice at Chichen Itza in Yucatan and even in central Mexico. The styles of Veraguas influenced the goldwork of Chiapas and Oaxaca. Fragments of Veraguas gold found under a Copan stela place this trade in the Maya area as early as the 8th century A.D. A few of the gold artifacts from Venado Beach in the Canal Zone are Veraguas types. Radiocarbon dates at this site are A.D. 227 ± 60 and A.D. 900 ± 70. No gold has been found associated with Scarified ware and other pottery types covered by the 230 B.C. date at Pueblo Nuevo. Veraguas imports range from an Olmeclike jade pendant to iron tools (Lothrop, 1950). They also include a little gold from Cocle (fig. 13,d) and a large amount of Polychrome pottery from the Asuero-Cocle region (fig. 14,a,c). Cross-dating is there197

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

FIG. 15-VERAGUAS METATES. (After Lothrop, 1950, fig. 30.)

fore sometimes possible, and it may be stated that the Veraguas graves reflect the entire span of painted pottery found to the east as far as the Canal Zone. GRAVE TYPES. Most Veraguas graves consist of a shaft and lateral chamber or are bottle shape. Depths range from 2.0 to over 10.0 m. Owing to the narrow orifice, excavation is difficult. Although a few multiple burials with rich furnishings have been found, notably the "cacique's grave" at La Peña, in general there are few grave goods. Gold occurs with about one out of ten 198

burials. Cemeteries often are located on mountain ridges. Individual grave shaft3 may be marked by stones. STONEWORK. NO large stone statues are known in the Province of Veraguas, although they occur both to the east and west. Unusually large metates are found, however, sometimes several in a single grave. An early type, recorded at Pueblo Nuevo, is oval in outline, has three cylindrical legs, and birds or animals carved in high relief on the base (fig. 15; Feriz, 1959). A second group also is three-legged with a flying panel attached to one leg, a form found in the Costa Rican highlands. Jaguar-effigy metates are common, also a four-legged rectangular type which may weigh as much as 100 kg. Manos are round or oval in section. A few stirrup pestles of fine workmanship occur. Stone celts may be pear shape, as in Cocle, with a polished blade and chipped poll. An elongated type, diamond shape in section, polished all over, seems most common in Veraguas but also is known in Costa Rica and Cocle. A few stone chisels like those from Cocle are known. Chipped points may be reworked flakes, either with or without a tang (Mitchell, 1959a). Trifacial points, of reddish jasper, usually crudely pressure-chipped, are characteristic. Stone jewelry, relatively common in Cocle, apparently was not manufactured in Veraguas. Imported pieces, usually agate beads or bar pendants, have been found. METALWORK. Except imported pieces in Cocle style, all metalwork is presumed to be a local manufacture. Beaten gold disks, sometimes with repoussé designs, are found, but characteristically everything was cast, even minute gold beads, rivaling those from Ecuador in delicacy of workmanship. Unlike Cocle, castings have an open back and, except for bells, rarely enclosed a clay-andcharcoal core. The metal usually is tumbaga, a conscious alloy of gold and copper with silver present as an impurity in the former.

FIG. 16—SCARIFIED WARE EARLY COCLE POLYCHROME, SITIO CONTE, PANAMA, α, Scarified Ware bowl. Volcano Guacamayo, Panama. Diam. 6 5/8 inches, b-e, Polychrome vessels from Sitio Conte. b-d, Peabody Museum, Harvard University; e, Claflin Collection. Height of b,c,e, 18, 14, 20 cm.; diameter of d, 33.5 cm.

199

FIG. 17—COCLE POLYCHROME VESSELS AND DESIGNS, SITIO CONTE, PANAMA, α-c, Early types. e,f, Late types. Peabody Museum, Harvard University.

Surface enrichment by the mise-en-couleur technique was practiced. Most gold artifacts are pendants with a small loop for suspension at the back. Commonly represented are birds ("eagles"), frogs (fig. 13,c), jaguar, dogs, shark, turtles, monkeys, and men. Many of the largest and most elaborate pieces portray anthropomorphic deities with human bodies and bird or animal heads. There is a constant tendency, however, to link elements of two or more animals. Thus in figure 13,b the bird head is flanked by profile crocodile heads; a fourth head, perhaps a dog, rep200

resents the body to which the wings and tail of a bird are attached. No traditions to explain such symbolism have survived. POTTERY. Veraguas pottery usually is unpainted. The buff-brown clay originally had a burnished self-slip which often has flaked, revealing rather coarse sand tempering. Decoration is simple filleting and incising. To a limited extent, shapes such as hollow bowls on tall pedestal bases are shared with the Asuero peninsula. In general, however, shapes must be regarded as local developments which distinguish Veraguas from adjacent areas. This includes a number of

ARCHAEOLOGY: L O W E R CENTRAL AMERICA

effigy forms, jars which appear to be cut in half, pointed-base jars with huge handles, bottles with tall narrow necks (Mitchell, 1959b, fig. 6) and barrel-shape jars. Among other characteristics are jars with twin necks (Dade, 1959, fig. 5, top right), slab tripods and a variety of pedestal bases, also looped tripod supports and double rims, illustrated in figure 13,f,g. In regard to the several hundreds of Polychrome vessels from Veraguas graves, in many cases they represent imported styles characteristic of Cocle or the Asuero peninsula (fig. 14,a), but there are also types of unrecorded provenience, now known only through scattered finds in Veraguas (fig. 14,b). A few crudely painted vessels appear to be local products. Vessels exhibiting crocodile gods or birds against a black background (fig. 14,c) have been reported in some numbers near Rio Jesus and may have been manufactured in the vicinity. Attempts to cross-date local and imported types have just begun. CENTRAL PANAMA

Under this heading we discuss the area between Veraguas Province and the Canal Zone, including the Asuero peninsula (Herrera and Los Santos Provinces), the Province of Code and Panama, also the Pearl Islands. We have drawn on published material and also museum and private collections, chiefly those of Mrs. Thelma H. Bull, Messrs. K. P. Curtis, P. L. Dade, H. A. Dunn, N. A. Harte, R. H. Mitchell, and K. W. Vinton. From an archaeological point of view, the entire region probably should be regarded as a unit from the development of painted pottery designs in the early centuries A.D. to the conquest. Certain pottery types such as Black-line on Red ware (Santa Maria Polychrome, see Willey and Stoddard, 1954) and early aspects of Polychrome ware from Venado Beach or Sitio Conte may be expected to turn up anywhere.

Local styles abound but there is a common symbolism suggesting a population not unlike that of the 16th century. On the other hand, aboriginal remains of earlier date such as Monagrillo appear to stand by themselves or, as in the case of Guacamayo ware of Code, appear to have affiliations with Chiriqui and Costa Rica. The distribution of certain nonpolychrome pottery types from Venado Beach, the vicinity of Chame, and the oldest Sitio Conte refuse is far from clear. Some pottery styles, for instance from the Madden Lake caves (N. A. Harte, 1958), are completely unplaced in time. Wooden objects have been discussed by Mitchell (in press). Preceramic Finds The oldest indication of the presence of man in Panama are fishtail fluted points found in Madden Lake (Sander, 1959, fig. 9). These undoubtedly are a Palaeo-Indian type which has been reported sporadically as far south as Patagonia, where a radiocarbon date of about 6680 B.C. was obtained. There is no evidence for dating in the Isthmus. The earliest settlement in Panama is Cerro Mangote near the mouth of the Parita river, to which a radiocarbon date of 4850 B.C. ± 100 has been assigned (McGimsey, 1956, 1957). Here rather crude percussion-shaped implements were discovered, also secondary and flexed burials of types which persisted in later times. First Ceramic Horizon The oldest pottery yet found in Panama comes from the Monagrillo shell heap at the mouth of the Parita river, to which a radiocarbon date of 2130 B.C. has been assigned. First explored by the Stirlings and Willey in 1948, the latter conducted additional excavations in 1952 (Willey and McGimsey, 1954). Monagrillo artifacts include percussionchipped scrapers, choppers, and grinding 201

FIG. 18—EARLY PAINTED POTTERY FROM PANAMA. a,c,e, Sitio Conte, Cocle. b,d, Venado Beach, Canal Zone. Peabody Museum, Harvard University.

tools comparable to Cerro Mangote finds. Crude mortars are of a type persisting until the introduction of the metate. Monagrillo pottery, as Willey has pointed out, is the most simple now known in Mesoamerica or South America. It consists of open or globular bowls and beakers in a Red ware or Natural Clay ware. Simple incising occurs but there are no spouts, handles, or basal supports (fig. 17). Except for recently announced radiocarbon dates from Ecuador (Evans and Meggers, 1958), the Monagrillo pottery is the oldest yet found in Latin America. Second Ceramic Horizon If our radiocarbon dates are accurate, there is a hiatus of 19 centuries between 202

Monagrillo and the next dated pottery in Panama, which was found in a deep shaftand-chamber tomb under a large mound of earth near Pueblo Nuevo on the ChiriquiVeraguas boundary (Feriz, 1959). The Pueblo Nuevo date is 230 B.C. ± 60 (DeVries, 1958). Archaeology has not yet probed this long period of obscurity. Pueblo Nuevo pottery consisted of (1) Scarified ware, (2) Guacamayo ware, and (3) large flat rims with broad incising, sometimes separating color zones. Scarified ware from Chiriqui and southern Costa Rica has been assigned an early date by Haberland. It has been found repeatedly by E. M. and N. A. Harte and by others with Guacamayo ware on the thus-named volcano in Cocle.

ARCHAEOLOGY: LOWER CENTRAL AMERICA

Guacamayo ware, as now known, consists of tall cylindrical vessels with partly rounded base and flaring rims. Examples without field data have come from various Cocle, Asuero peninsula, and Veraguas sites. Decoration consists of parallel incised lines or simple filleting with or without incising. Sometimes there is a broad band of red paint. Large flat rims with broad incising like the Pueblo Nuevo specimens were found in deep refuse at the Sitio Conte, together with sherds exhibiting fine-line parallel incising, shell-edge incising, and punctate decoration and filleting. As data on shapes were lacking, the group was described and illustrated under the heading "Miscellaneous." The Sarigua Complex described by Willey and McGimsey, marked by bowls with appliqué ridges, plain and shell incising, and punctate decoration, probably belongs in this ceramic horizon. Undescribed pottery from Venado Beach as well as part of the material excavated by Mrs. Bull near Chame (1959, figs. 21-23) should be included.

FIG. 19—OLDEST-KNOWN POTTERY TYPES FROM PANAMA. Monagrillo, Province of Herrera, a-d, Monagrillo Incised, e-j, Monagrillo Red. k, Stone ware fragment with incised design similar to that of pottery. (After Willey and McGimsey, 1954, fig. 12.)

First Painted Ceramics The oldest record of pottery with painted designs comes from the bottom levels in stratigraphic trenches at Sitio Conte in Cocle and at the Giron site on the Santa Maria River (Willey and Stoddard, 1954) in the Asuero peninsula. In each case, two principal types occur. One is a Black-line on Red ware with both zoomorphic and geometric patterns, usually painted on the interior of shallow bowls (fig. 18,a). The other, called Black-line Geometric, consists of large globular jars with flaring necks. Base and neck are painted red, but a light slip on the upper half of the body is decorated by areas crosshatched in black (fig. 18,c). This type comprised 25 per cent of the sherds in the two bottom cuts at Sitio Conte (Ladd, 1957). Both these Black-line wares occur in Veraguas graves but with no recorded as-

sociations. Black-line on Red has been found on the Pearl Islands (Linné, 1929, fig. 24) and at Venado Beach in the Canal Zone, from which radiocarbon dates of about A.D. 230 and 900 have been obtained, in each case from charcoal samples. Venado Beach exhibits such unity of style, both in pottery and in metal or shell ornaments, that these dates may indicate too long an occupation. Painted pottery is not characteristic, but the principal type is a Black- and Red-line ware with both pigments used as delineating colors on a light slip. In part, the designs are those of the local Black-line on Red ware; also there are scroll patterns and a few zoomorphic motifs (fig. 18,b,d) as seen on the oldest Cocle Polychrome from Sitio Conte. Venado Beach thus exhibits the shift from single203

FIG. 2 0 - E X A M P L E S O F ASUERO STYLES FROM T H E E L HATILLO SITE, HERRERA PROVINCE, PANAMA. Diameters: a,b, 12.5 cm.; c,d, 13 cm.; e, 28 cm. (estimated); f, 17.5 cm. (bowl, excluding wings). Heights: a,b, 7 cm.; c,d, 9.5 cm.; f, 21.5 cm. (total), 8:5 cm. ( b o w l ) . (Courtesy, John Ladd, Peabody Museum, Harvard University.)

ARCHAEOLOGY: LOWER CENTRAL AMERICA

line patterns to a full polychrome technique with black outlines completely color-filled. Black- and Red-line wares include early painted pottery from Costa Rica (fig. 6) where, as in the Canal Zone, conventionalized crocodile motifs are common. This decorative technique had a long history and, in Chiriqui Alligator ware, persisted until the 16th century. At Venado Beach rare pottery types must represent contemporary imports with no specific provenience. Among these are red and brown vessels with fine-line zone incising. Also there is decoration consisting of a black line flanked by white lines. Scattered examples are known from early Sitio Conte graves, a few Asuero and Veraguas sites and from Diquis delta refuse. The technique but not the style occurs at an early date in northwestern Costa Rica (fig. 5,b). Nothing which presumably originated in the Asuero peninsula is known from the Canal Zone. On the other hand, some of the incised pottery such as the shell incised must share a common origin. Polychrome Ceramics The Polychrome pottery from the Asuero peninsula and the adjacent Province of Cocle forms one of the most complex and decorative groups yet discovered in the New World. Present knowledge of the two areas is uneven because most published Cocle material comes from recorded excavations whereas Asuero types still are chiefly known from purchased collections (Lothrop, 1937-42, pt. 2). Ladd's nearly completed monograph on the latter region will greatly amplify the picture. 1 Each area produced generations of great artists who were masters of line and color. Each developed a great number of local styles, which usually can be separated at a glance. On the other hand, Sitio Conte 1 Ed. note: Some Asuero types have been designated as "El Hatillo" (see Willey and Stoddard, 1954).

refuse reveals contemporaneity of the two (Ladd, 1957). Asuero types appear in refuse with the oldest Code styles and in time became preponderant in daily use, although rarely found as funeral furniture in Code. Symbolic concepts were shared. Trade in both directions evidently was on a very large scale in spite of linguistic and tribal barriers. There are certain distinctions which are characteristic. In Code, large flat plates offered a field decoration which called for massive designs, both geometric and zoomorphic. Tall necks on jars and carafes were adorned by boldly executed scroll patterns, of which there are several dozen varieties. Areas filled with color often are large. In contrast, most Asuero vessels are smaller. Shallow bowls on tall pedestal bases replace flat plates. These usually are adorned in a more delicate quality of line with smaller areas filled in color (figs. 14,a; 20; 21). Globular jars with short necks often are decorated with animal motifs. At the Sitio Conte it has been possible to recover many complete designs from funeral pottery and, owing to the superposition of graves, to arrange them in chronological sequence (see figs. 16,b-e; 19). Two major groups emerge with little overlap in styles. Future excavations should expand this series. In addition to Polychrome ware there are many others from the Sitio Conte. Among these are a Red ware, a Panelled Red ware (with small light areas decorated by Blackline patterns), Smoked ware (a Red ware darkened to brown or black), Red-line ware, and many more. During the Polychrome era, the Sitio Conte must have been the summer home of powerful chiefs—it is flooded during most rainy seasons. A feature of this settlement and several others in Code was a large alignment of stone columns, now partly destroyed by the adjacent river. In the largest group of this kind yet reported 205

FIG. 21—EXAMPLES OF ASUERO STYLES FROM THE EL HATILLO SITE, HERRERA PROVINCE, PANAMA. Diameters: a,b, 20.5 cm. (bowl, excluding wings); c,d, 17 cm.; e, 20.5 cm.; f, 16 cm. Heights: a,b, 10 cm.; c,d, 21 cm.; e, 13 cm.; f, 12-13 cm. (Courtesy, John Ladd, Peabody Museum, Harvard University.)

ARCHAEOLOGY: LOWER CENTRAL AMERICA

(Verrill, 1927), many of the columns were carved. Additional investigation is needed. Sitio Conte graves were rectangular pits which contained from one to over a dozen bodies. In general, the richer the individual, the deeper the grave, which might cut through earlier and shallower burials. A family relationship is suggested by the fact that furnishings were transferred from one grave to another. Large graves contained the slaves or retainers of the chief, who could be readily identified by his jewelry. Up to 200 pottery vessels were found, all newly made but usually broken. Large numbers of celts, chipped blades, stone and bone arrow or spear points, traces of bark cloth and woven textiles were present. The jewelry was outstanding and no other site in Panama has yet been discovered comparable to the Sitio Conte (Lothrop, 1937-42, pt. 1; Mason, 1942). Most gold was hammered and embossed with characteristic designs. It included circlets, helmets, beads by the thousands, nose and ear ornaments, arm bands, greaves and great circular breast plates (fig. 13,e). Casting was highly developed (fig. 13,d) and usually was in the round over a clay-

and-charcoal core. Perishable objects were encased in sheet gold, and gold settings were made for emeralds and carved whale'stooth ivory. Beads and pendants were also carved of serpentine and agate. In short, the inventory fully confirmed the picture of the natives left by the Spanish invaders. Cocle manufactures were exported to the west and north. The pottery definitely reached southern Costa Rica and probably Nicoya where typical Cocle forms such as tripart jars appear in local clay. Bar pendants of agate reached Oaxaca in Mexico. Gold artifacts are found in Costa Rica and even in Yucatan. It is strange that today we do not know who these Coclesanos were. Some were as bearded as the Spaniards. West of Chame, they could not talk to each other and needed interpreters from village to village. The Province of Darien to the east of the Panama Canal evidently was thickly settled and prosperous at the time of the conquest. Archaeological types described by Linné (1929), however, are exceedingly simple and cannot be placed in the pattern revealed to the west.

REFERENCES Aguilar, 1953 Balser, 1953 Baudez, 1959, 1962 and Coe, 1962 Bovallius, 1886 Bransford, 1881, 1882 Bull, 1959 Cereceda, 1926 Coe, M. D., 1962a Dade, 1959 DeVries, 1958 Evans and Meggers, 1958 Feriz, 1959 Flint, MS

Haberland, 1955, 1957a, 1957b, 1959a, 1959b, 1959c, 1960a, 1960b, 1960c, in press a, in press b and Grebe, 1957 Harte, Ε. Μ., 1958 Harte, N. Α., 1958, 1960 Hartman, 1901, 1907 Holmes, 1888 Ladd, 1957 Lehmann, H., 1959 Lehmann, W., 1908, 1913 Linné, 1929 Lothrop, 1919, 1926, 1937-42, 1950, 1955, 1963 MacCurdy, 1911

207

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

McGimsey, 1956, 1957 Mason, J. Α., 1942, 1945 Mitchell, 1959a, 1959b, in press Osgood, 1935 Richardson, 1940, 1941 Sander, 1959 Skinner, 1926 Spinden, 1917a Squier, 1852

208

Stendahl, 1952 Stone, 1943, 1951, 1958 and Balser, 1958 Swauger and Mayer-Oakes, 1952 Verrill, 1927 Willey and McGimsey, 1954 and Stoddard, 1954 Williams, 1952

10. Synthesis of Lower Central American Ethnohistory Γ

DORIS

L

ower Central America covers the region from the Aguan River and the Gulf of Fonseca in Honduras through the ancient province of Cueva or Darien in Panama. The early chroniclers noted a marked similarity in the customs of most of this territory, especially on the Caribbean side, to those of the Antillean islands of Cuba and Haiti (Hispañola) (cf. Fernández de Oviedo, 1851-55, 1:133; López de Gómara, 1941, 2:202; Herrera, 1726, dec. 4, lib. 1, cap. 10, p. 18; Las Casas, n.d., cap. 244, pp. 577-81). Three cultural spheres can be determined: that of rain-forest people, of lower Andean groups, and of people of northern or Mexican extraction. SOURCES

This survey is based on reports dating from 1502 through the 17th century. Because material covering much of the Caribbean littoral is scarce, I have supplemented it in footnotes with information from later accounts. PEOPLES

Physical types differed. In Darien, Panama, copper-colored people with straight

STONE

black hair, round faces, short noses, large gray eyes, high foreheads, white even teeth, thin lips, and large mouths predominated. Men were tall and big boned; women, full breasted and well shaped. Albinos apparently belonged to the same ethnic group (Wafer, 1934, pp. 78-79, 80-81). In addition, extremely dark men were seen (López de Gomara, 1941, 1:162). Westward on the Pacific, very tall people were found in Escoria (Fernández de Navarrete, 1945, 3:404) and in Chiriqui (Cockburn, 1779, p. 147). The Chorotega of Costa Rica and Nicaragua are described as "presentable and handsome" (Fernández de Oviedo, 1851-55, 3:109). On the Caribbean side, at Puerto de Bastimentos, Panama (Nombre de Dios), the people were lean and good-looking (Colón, 1947, pp. 288-99). In Costa Rica, attention is called to well-built inhabitants (ibid., p. 282; Mártir de Anglería, 1944, p. 231), but westward, by Gracias a Dios, people were "dark and homely" (Colón, 1947, pp. 277-78). Still farther to the west, near Cape Caxinas, were tall, well-formed men (ibid., p. 277; Mártir de Anglería, 1944, p. 229). 209

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

LANGUAGES

Lower Central American languages have been little studied, either descriptively or with respect to their external affiliations. Nevertheless, it has been suggested that this area is a meeting-ground between representatives of stocks whose centers of gravity lie either farther south or farther north. A majority of the languages here mentioned are probably Chibchan outliers or members of a group ( M i s u m a l p a n ) which may b e part of a larger Macro-Chibchan entity. A minority constitute either southern outliers of well-known northern groups ( O t o m a n guean, Hokaltecan, Mayan, Aztecoidan) or are isolated languages for which no external affiliations have been m a d e plausible. Of the southern outliers, one group, those speaking varieties of Nahuatl, are clearly relatively recent invaders. It is not impossible that a fair portion of the Chibchan affiliates in lower Central America constitute back-migrations from a dispersal center in Colombia. On the m a p we have placed the tongues belonging to the Pacific southeast of Costa Rica under the term Boruca. This is a modern n a m e and covers the anand cient Coto, Turucaca, Burucaca, Abubaes who inhabited this region. Until careful descriptive and historical studies have been carried through on these languages, however, it will not be possible to bring linguistic data to bear on these problems. T h e following affiliations have been suggested for the languages whose locations are discussed in the following text († represents languages not spoken t o d a y ) . A. Arawak B. Macro-Chibchan Misumalpan Matagalpa † Misquito Taguaca (Sumu) Ulva Chibchan (Caribbean) Bribri 210

Cabécar Cueva-Cuna (both Caribbean and Pacific) Guatuso (a modern contraction of Corobici t and Voto t ) Guaymi (both Caribbean and Pacific) Guetar † Paya Rama (from Voto) Suerre † Térraba (Teribí) Urinama † Chibchan (Pacific) Boruca (a modern contraction of Coto †, Turucaca †, Burucaca †, Abubaes †) Changuena † Chiru † Cueva-Cuna (both Caribbean and Pacific) Dorasque f Guaymi (both Caribbean and Pacific) Nata † Quepo † C. Macro-Otomanguean Otomanguean Chorotegan Chorotega-Mangue † D. Hokan-Siouan Hokaltecan Supanecan Maribio (-Subtiaba) † Jicaque Azteco-Tanoan Uto-Aztecan Aztecoidan Nahua (Nahuat, including Nicarao †, and Nahuatl [Chuchures? †]) E. Unclassified Cuba (see note 1) Escoria † Lenca (Poton †: see note 21 and Stone, 1957, pp. 84-85) Tacacho † Cueva was predominant in most of eastern P a n a m a extending through the province of Darien to C h a m e on the Pacific coast ( F e r n á n d e z de Oviedo, 1851-55, 4:117), although Chocó ( P a p a r o ) was spoken at the extreme eastern periphery (Serrano y

ETHNOHISTORY: LOWER CENTRAL AMERICA

Sanz), 1908, p. 123). At Chame, Coyba (Coiba), and Purulata, the "language of Cuba" was spoken1 (Herrera, 1726, dec. 4, lib. 1, cap. 10, p. 18; Serrano y Sanz, 1908, p. 5). Between Chame and Chiru where a different tongue was used, lay 8 leagues of uninhabited territory (Fernández de Navarrete, 1945, 3:402-03) and then four more to Nata. Nata and Escoria each had a distinct language, 2 whereas westward to Burica, the actual boundary with Costa Rica, were varied dialects. On the Caribbean side of western Panama, documents cite the Dorasque, Térraba, Changuena, and Guaymi. The last extended both west and south (Costa Rica-Panama Arbitration, 1913, 1:488, doc. 172). 3 Columbus found that the interpreter he took on board at Cariari (Limon), Costa Rica, was useful as far as Cubiga in the vicinity of the Cocle River (Colón, 1947, p. 286). There was one totally different speech beyond this area, however, that of the Chuchures, spoken at Puerto de Bastimentos. According to tradition, the Chuchures came from Honduras in canoes and were killed or dispersed at the time of Diego de Nicuesa. It is possible that they were from the Nahua colony Cortés found near Trujillo (MacNutt, 1908, pp. 317-18). Torquemada also recounts that Mexicans went from Nicaragua to Nombre de Dios (Torquemada, 1943, tomo 1, lib. 3, cap. 40, p. 333). 4 Bribri (Viceyta) and Cabécar, both known as Talamanca, were spoken in northeastern Costa Rica. The Urinama may have had a language of their own, but in 1709 spoke Cabécar (Fernández, 1881-1907, 5:458). In the northwest was Suerre,5 the language used as far as the Jori or Sarapiqui River where Voto took over (ibid., 2:22730; Peralta, 1883, p. 733). Corobici, which seems to have been the mother tongue of the Nicoya peninsula and perhaps of the Chara and Pocosi islands in the Gulf of Nicoya (Fernández de Oviedo, 1851-55, 3:110) as well as southwestern Nicaragua (cf. ibid., 3:541; López de Gomara, 1941,

2:219; Herrera, 1726, dec. 3, lib. 4, cap. 7, p. 121 ), 6 was important at the time of the conquest from the Tempisque River eastward to the mountains of Tilaran and probably to Rio Frio (Fernández de Oviedo, 1851-55, 4:108; Peralta, 1883, p. 54). Guetar was spoken from Tayutique, the boundary of Suerre and Pocosi (Matina) 7 (Fernández, 1881-1907, 4:315, 362) to the Gulf of Nicoya. Guetar-speakers served as interpreters in the south among the Quepo and Coto-Turucaca, which includes Boruca (ibid., 4:294; 1:286), and in the north among the Talamancas. On the Caribbean coast, however, in 1540 when the first Spaniard reached there, 8 a Mexican dialect was used in the most eastern flood plain of present Costa Rica. This plain was then referred to as Coaza or Coaça, and is today called Talamanca (Fernández, 1881-1907, 6:308; Peralta, 1883, pp. 334, 350). 9 One historical reference indicates the Utaztecan stock which was spoken by the Chichimecs (Fernán1 "The dances, rites and religion" of Cuba and Española and the customs of Española were said to exist in Darien, Panama, and Veraguas (Colón, 1947, p. 297; López de Gomara, 1941, 2:202). It is possible that the language referred to was a branch of Arawak, the most widespread tongue on these islands at the time of the Spanish arrival. There is no other linguistic evidence, however, to suggest that Arawak was ever spoken in Central America. 2 Lehmann (1920, 1:144) suggests the speech of Escoria might have been Chorotega. 3 I do not list the Mexicans on the island of Tojar because documents mentioning this are relatively late, 1697 and 1766. These Mexicans probably arrived with the first Spaniards in 1540 or later. See Stone, 1956b. 4 This statement of Torquemada is acceptable as the Nicarao were Nahua-speaking. It is not farfetched to imagine that they might have joined their countrymen from the Honduran colony. 5 See Lehmann, 1920, 1:213-14. 6 See Lehmann, 1920, 1:375-78. 7 Pocosi should also be included as Talamanca. The Chirripo who inhabited this region were Cabécar-speaking. See Stone, 1962. 8 It must be remembered that Columbus only touched at Cariari. 9 For a discussion of this see Lothrop, 1942; Stone, 1956b.

211

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

dez, 1881-1907, 6:297). Oto-Manguean speech was also found in a small enclave by the Naranjo River on the Pacific watershed (ibid., 4:369) and was used by prisoners kept for sacrifice in the central plateau at Pacara (today Villa Colon) (Fernández Guardia, 1908, p. 38). Two Mexican tongues were dominant on the Nicoya Peninsula when the Spanish arrived: Chorotega-Mangue in Nicoya, Orotiña, and Orosi (Fernández de Oviedo, 1851-55, 3:111, 121); and Nicarao in the Santa Elena Peninsula, Bagaces (Fernández, 1881-1907, 1:270-71), and as a wedge between the Corobicis and ChorotegaMangue to the Gulf of Nicoya (Fernández de Oviedo, 1851-55, 4:108). ChorotegaMangue speech seems to have been in this region for so long that López de Gomara called it "natural and old" (López de Gomara, 1941, 2:219). Beginning with Nicaragua and extending into and beyond Honduras are general terms which include non-northern or nonMexican tongues, the most common being Carib and Chontal (Squier, 1860; Herrera, 1726, dec. 4, lib. 8, cap. 3, p. 156; Vázquez, 1937-44, 4:81, 91-92, 121, 199). Another similar name belonging more to Honduras is Jicaque (Xicaque) (ibid., 4:77, 123). 10 There are also three over-all designations in the early documents: Ulva, Taguaca, and Lenca. 11 Each of these terms is applicable to diverse groups which were dialectically and ethnologically bound together. One of the first reports concerning Caribbean Nicaragua is that of Alonso Calero in 1539 (Peralta, 1883, pp. 737-38). "Three or four" tongues are cited along the San Juan River without direct reference to which bank (ibid., p. 731). We have seen that in Costa Rica, Suerre, Voto, and Corobici were spoken. The Voto language appears to have extended across the San Juan basin as far as Bluefields (Peralta, 1898, p. 125; Strangeways, 1822, p. 31) but was called Rama, Arrama, Mechora, and Carib 12 (García Paláez, 1852, 3:154). 212

The neighboring language to Rama, and perhaps the most important non-northern tongue of Nicaragua, was Ulva, often termed Chontal. The name is spelled in many ways including Ulua, Oldwaw, Woolwa (García Peláez, 1852, 3:156; Μ. W., 1752, p. 305), 13 but I prefer Ulva as offering less confusion with the geographical designation Ulua.14 The Ulva language extended from the southeastern portion of the department of Chontales throughout the tributaries of Rios Escondido and Bluefields.15 The exact northern boundary in Nicaragua is not known, but at least one town, Zomoto, on the Pacific coast is cited as Ulva-speaking in the latter half of the 16th century (Ponce, 1873, 1:342). It is highly probable that the Guanexicos who were located east of Olocoton also spoke Ulva.16 The Matagalpa bordered on the Ulva at the northwest in the actual department of Chontales. A close relationship existed between the two tongues. Regardless of dialectal connections, if we base our investigation on historical sources, we do not find the name Matagalpa outside Nicaragua (de Vilaplana, 1763, p. 116). 17 From the Prinzapolka river system northward into eastern Honduras, which seems to have been their homeland, were the Taguacas (Vázquez, 1937-44, 4:110-14, 119, 196, 202). 18 Vásquez notes (ibid., 10 For a more lengthy discussion see Stone, 1941, pp. 10-12. 11 See Stone, 1957, pp. 3-5, 77-78, 84-87. 12 See Lehmann, 1920, 1:420. 13 See Lehmann, 1920, 1:468. 14 Lehmann, 1910, p. 718; 1920, 1:468, 477, 479; Conzemius, 1932, p. 14-28; Stone, 1957, pp. 77-78. 15 Lehmann, 1920, 1:463. 16 Lehmann, 1920, 2:1012; cf. Conzemius, 1932, pp. 20, 28. 17 Linguists, however, associate the language with south-central Honduras and northeastern El Salvador (Habel, 1878, p. 2 1 ; Lehmann, 1920, 1:471, 997), where it is known as CacaoperaMatagalpa. Lehmann (loc. cit.) identifies this with the Ulva language mentioned by Palacio. 18 Lehmann, 1920, 1:471-79; Conzemius, 1932, pp. 14-16. There are many ways of spelling this name, e.g., Twocka, Twaka, Tauaxka.

ETHNOHISTORY: LOWER CENTRAL AMERICA

4:110) that their language resembled Mexican. Later the Sumu became identified with this group. 19 The Lenca were allied with the Taguacas in the provinces of Taguzgalpa and Tologalpa, which included from Castilla Point to Lake Nicaragua (ibid., 4:78-79, 189; Serrano y Sanz, 1908). Allusion to a Lenca tongue in Nicaragua was made in 1576 by Palacio who as " O y d o r " of the Royal Audience of Guatemala arranged a contract with one of the first conquistadors to enter the province of Teguzgalpa (Squier, 1860). 20 Although Palacio himself did not enter this region, he had access to men and reports now gone forever.21 The remaining tongue associated with the Atlantic littoral of Nicaragua is Misquito, a term often combined with Zambo and not seen before the 18th century. At this period, Misquitos appeared to have occupied the coast from the Black River to the San Juan basin (Fernández, 1881-1907, 9:187, 189, 327; Peralta, 1898, pp. 43, 120) holding in subordination the Ulvas, Taguacas, etc., who lived farther inland (García Peláez, 1852, 3:156). Corobici extended from Costa Rica into the Solentiname Islands and may have been the original tongue of the Isthmus of Rivas and Ometepe Island (Ponce, 1873, 1:369). The most common idiom of Pacific Nicaragua at the time of the conquest, however, was Mangue, the speech of the Chorotega. It was used in parts of the province of Nagrando (Fernández de Oviedo, 1855, 4:33, 61), Masaya, Nindiri, Mateare, Nagarote, Mabiti, and Subtiaba (Ponce, 1873, 1:356-69), as well as Nicoya, Costa Rica, and a small enclave by Rio de los Mangues. Maribio was spoken in Cinandega, Pozolteca, Chichigalpa, and Mazatega (ibid., 1:354-56). Between Cinandega and Subtiaba, however, lay the town of Yacacoyana distinguished by its own language called Tacacho which appears to have been the aboriginal tongue of this section of Nicaragua (ibid., 1:356).

The rest of the west coast was inhabited by Nahua-speaking peoples. This included besides some of the province of Nagrando (Fernández de Oviedo, 1851-55, 4:37), sites such as El Viejo and Chinandega in the north, and in the south, the Isthmus of Rivas where in the vicinity of Granada, 19 This probably was a result of close communication with Mexican tribes in the forest habitat of Taguzgalpa (Stone, 1957, pp. 7 7 - 7 8 ) . Sumu is a word applied by the Misquitos to all jungle Indians (Lehmann, 1920, 1:471). The term is not found in the early sources. 20 The spelling of the term for the entire area and the separate provinces is frequently confused (Vázquez, 1944, 4:78, 202). For the sake of clarity, I use the terms Taguzgalpa and Tologalpa for the two provinces and Teguzgalpa for the combination of them both. 21 In this connection, although the ancient province of San Miguel, eastern El Salvador, does not fall within the limits of this paper, I point out that Palacio's statement that here Poton, Taulepa, and Ulva were spoken is in keeping, as far as the first and last name are concerned, with the report of Ponce, who passed through this section eight years later. Taulepa, a term Squier notes was taken from the old name of Lake Yojoa, Honduras, can be classified as Ulua-Maya. This was not a "true" Maya people but one culturally mixed and including both Mesoamerican and non-Mesoamerican stocks. The dominant element seems to have been a Maya branch heavily influenced by certain Teotihuacan theological concepts. This is corroborated by archaeological finds lying in eastern El Salvador and Lake Yojoa and extending into the islands of the Bay of Fonseca. It is possible that at one time the Ulua-Maya inhabited these regions but did not survive, as did the Ulva and the Poton, a "non-Mayanized" Lenca people. In this connection it should be remembered that Ponce remarks (1873, 1:386) that the Indians in this southern area "son pocos," which suggests a drastic change in population at the time of or right before the Spanish conquest. I believe that by combining data furnished by later chroniclers, such as Vázquez, and the observations of linguists, we have evidence which indicates that Palacio was correct in counting the Poton tongue as one found in Nicaragua. I classify this language as Lenca (Stone, 1957, pp. 84-86) and suggest that it might have been spoken in the department of Nueva Segovia where Lenca peoples such as the Parakas appear to have penetrated in preconquest times. Lehmann indicates such an extension when he associates certain place names in Nueva Segovia with the Lenca (Lehmann, 1920, 2:1013). See also names such as Molaguina in de Vilaplana (1763, p. 116) and Consequina by the Bay of Fonseca.

213

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

Mangue and Nahua met (Ponce, 1873, 1:352, 354). 22 Fernández de Oviedo claims that the town of Managua was Chorotega (i.e. Mangue)-speaking (Fernández de Oviedo, 1851-55, 4:67). Ponce, however, who journeyed through this region in 1584, states that Nahua was spoken in Managua (Ponce, 1873, 1:359).23 There were two important provinces associated with the eastern Caribbean littoral of Honduras: Taia (Mártir de Anglería, 1944, p. 228) and Taguzgalpa, which we have already examined.24 Both were touched by Columbus, who found that the interpreter he took on board at the Bay Island of Bonocco was useful on the mainland near the Aguan River. Paya was spoken from Trujillo to the Patuca River and inland as far as the valley of Olancho where Lenca predominated. It also extended from the northern border of Jamastran northward into the forest land of the upper Patuca. The Taguaca extended from the Guayambre valley north into the same forest land and east into Nicaragua, whereas Lenca was heard throughout this inland section (see Serrano y Sanz, 1908, pp. 318, 352, 373; Vázquez, 1937-44, 4:196). 25 Toponymical evidence suggests the presence at one time of the Matagalpa in this region, but it may be that the Ulva extended here from Nicaragua (see Ulva above). Coastal Taguzgalpa was associated with the Misquito at least from the 18th century on, as we have noticed before. In this whole territory, Mexican tongues had infiltrated. In the vicinity of Trujillo, "the language of Culua" (Nahua) was in use (MacNutt, 1908, pp. 317-18). Farther east in the woodland, Mexican speech was of great importance until these people were dominated by uprisings of the Lenca and Taguacas in the early 17th century (Vázquez, 1937-44, 4:105-07). Documents show that in the 16th century Nahua was spoken on the shore of Fonseca Bay, although this site was abandoned by the end 214

of the century, the inhabitants moving to El Viejo in present-day Nicaragua (Ponce, 1873, 1:343-52). Ciuatepetl, in the Bay also, had a Nahua-speaking community (ibid., 1:379). Later, in the 17th century, Nahua served as a lingua franca in this southern province of Choluteca Malalaca (Vázquez, 1937-44, 4:31-32, 62-64). At the time of Palacio and Ponce, the dominant tongues in Pacific Honduras were Mangue (Squier, 1860; Ponce, 1873, 1:33738), Poton,26 and Ulva (ibid., 1:339-40). Ponce notes that from the Goascoran River on the southern Honduran mainland to Nicarahego near Nacaome, Mangue was spoken. From Ola through Colama, Santiago Lamaciuy (Namasigüe), Zazacali, Condega on the Negro River, to Somoto in Nicaragua, the language was Ulva. The Poton tongue was common in eastern El Salvador and used in the mainland port of Amapal as well as on the following islands in the Bay of Fonseca: Quetzaltepetl or Meangola (Meanguera), La Teca or Chonchagua, and Mazatepetl. Vázquez apparently is in accord with these designations as he names Lenca in connection with the convents of Nacaome and Amapal (Vázquez, 1937-44, 4:31-32, 62-64). SETTLEMENTS AND DWELLINGS

Community Patterns In spite of the fact that extensive areas were under cultivation (see Agriculture), 22

Lehmann, 1920, 1:910; 2:1002. Lehmann has pointed out that Fernández de Oviedo frequently confused Nicarao ( N a h u a ) and Mangue (Lehmann, 1920, 2:1003). Managua lay in Mangue-speaking territory but following the indications of Lehmann and the observations of the friars who were language-conscious, I feel that both languages may have been spoken there. The fact that the Nicarao were advancing in Chorotega-Mangue territory at the time of the Spaniards' arrival strengthens this point of view. 24 For discussion of Taia see Stone, 1941, p. 9. 25 Stone, 1957, pp. 74-79. 26 Ponce, 1873, 1:381-83. Ponton and Ulva should be considered under Chontal as given by Palacio. Chontal is not indicated by Ponce on the mainland. Vázquez, a later source, designates only Lenca (Poton) and Mexican on the mainland. 23

ETHNOHISTORY: LOWER CENTRAL AMERICA

there were no villages in the European sense of the word on the entire Caribbean coast. Ferdinand Columbus noted this when he referred to the "town" of Veraguas: " . . . and because I call it a town or village, it is to be observed that in those parts their houses are not close together, but they live as in Biscay, at some distance from one another" (Colón, 1947, chap. 97, p. 300). Similar comments are found throughout the early documents. Attention is called to the custom of group movements depending on game and fish (Fernández de Oviedo, 185155, 3:132), and that people moved about like Arabs (López de Gomara, 1941, 1:162). Throughout what is now Panama, it was customary for ranking Indians to have houses in different sections of their territory which were used according to the season. This manner of living continued through the Caribbean watershed of Costa Rica. All reports emphasize the temporary aspect of settlements which at most were two or three large habitations grouped according to kinship (Peralta, 1883, p. 71; Fernández, 18811907, 5:372) leaving the Spaniards with the problem of forming towns or "reducciones" (de Vilaplana, 1763, p. 40). As in Panama, dwelling sites were frequently moved (Benzoni, 1857, p. 127). 27 A similar pattern extends throughout Caribbean Nicaragua and northeastern Honduras. In 1699 the Misquito village at Sandy Bank was formed by 12 habitations which housed 400 people (M. W., 1752, p. 301). In Taguzgalpa the chroniclers noted that huts were called a town (Vázquez, 1937-44, 4:145). 28 In western Pacific Panama the first change in living pattern is suggested. Despite the fact that Oviedo makes it plain that permanent villages are lacking (Fernández de Oviedo, 1851-55, 3:132), his account of some 45 or 50 houses leads to the conclusion that this nucleus could literally be interpreted as a village. In addition, near Nata, in the province of Tobre and Trota where four different

languages were spoken, there were palisaded settlements with walls of thorny thistle-like plants closely interwoven (Herrera, 1726, dec. 4, p. 20; Relación . . . Espinosa, 1892, 2:494-95, 508). Concentrations for safety reasons appeared also on the Pacific slope of Costa Rica. Some had double walls separated by a dry moat (Fernández Guardia, 1908, p. 50). All sites were in strategic places near water and fields with walks, plazas, and a ceremonial area joined to the fort (ibid., pp. 35, 49-50). In the flood plain, and along the coast, fenced communities do not seem to have existed (Fernández de Oviedo, 1851-55, 3:99). Habitations in some places were in water, probably as a safety measure (Peralta, 1883, p. 6). The Nicoya Peninsula was the first place toward the north where many houses were grouped within sight of one another and not fortified. Here and in Pacific Nicaragua were towns with squares, ritual centers, dwelling quarters, market places, planted trees. They were associated with the Chorotega and the Nicarao. The Nicarao plan seems to have had a more elaborate living area for the chief and the ranking people and special quarters for goldsmiths, a detail lacking in Chorotega sites (Fernández de Oviedo, 1851-55, 2:109-11; López de Gómara, 1941, 2:216). 29 27 Archaeological evidence of a ceremonial or perhaps residential center consists of stone-covered earth mounds at Last Mercedes and sites with circular stone and earth enclosures, all on the Línea Vieja, Costa Rica, a region where excavated artifacts yield many articles of obviously foreign origin and indicate an important trading zone. 28 Some sites historically associated with the Paya, such as Tonjagua in the Agalta valley, Honduras, have pyramidal platform mounds which differ from the familiar system of eastern Central America and open the question as to who the original builders were. Near the present San Marcos de Colon there is also archaeological evidence of fortified hill sites which suggest the Lenca (Stone, 1941, p. 14; 1957, p. 8 1 ) . 29 Cf. Fernández de Oviedo, 1855, 5:109-10, and López de Gomara, 1941, p. 216, with his pp. 80-100. The description of Managua without a

215

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

Throughout this region, however, were small communities lacking market places and probably elaborate ceremonial and living centers. The majority were politically associated with the Nicarao (see Social Organization). In southern Honduras, Ulva and Poton sites are described as towns or a collection of huts (Ponce, 1873, 1:339-40, 381). In other words, houses were more closely grouped together than on the Caribbean side, suggesting northern influence. House Types On the Caribbean littoral of eastern Panama Columbus saw tree houses (Colón, 1947, pp. 291-92). There were both enormous houses supported by posts of black palm, at the mouth of the San Johan River (Fernández de Oviedo, 1851-55, 3:50), and tree houses. In Darien, "warhouses" with holes for shooting arrows were prevalent (Wafer, 1934, p. 90), and some habitations had a stone base (Mártir de Anglería, 1944) but most dwellings were of cane or wood roofed with grass. A few were grouped together, although the majority were scattered over a large territory under distinct chieftains (see Villages and Social Organization). At times bijagua (Cafathea insignis Petersen) leaves were used for roofing (Fernández de Oviedo, 1851-55, 1:276). Fernández de Oviedo states (3:131-32) that these items were "materials . . . at hand wherever they go." Similar construction and distribution continued into the Atlantic watershed and highland region of Costa Rica (Fernández, 1881-1907, 3:37; 5:370). The only allusion to the use of mortar was made by Ferdinand Columbus, who writes that main plaza but with many plazas puts it in a category apart similar to a "mushroom" town, unlike Nicoya, a Chorotega-Mangue site, and unlike Tecoatega (El Viejo) of the Nicarao. Cf. ibid., pp. 67, 109. 30 Sapper, 1904, p. 27. 31 The high roof and the clay object are reminiscent of the dwellings of the present-day Guaymi of Panama and the Bribri of Costa Rica.

216

they saw "a great mass of stucco," apparently of stone and lime, beyond the eastern end of the Valiente peninsula at Cubiga (Colón, 1947, p. 286). A separate dwelling for the chief is typical of Panama, Costa Rica except fortified sites in the south, Nicaragua, and the Lenca in Honduras. Special mention is made of the chiefs house in Suerre, Costa Rica. It was an egg-shaped construction of reeds roofed with closely interwoven palm branches. The few other houses were "of a common sort" (Benzoni, 1857, p. 127). The earliest reports of dwellings in nonPacific Nicaragua refer to those in trees (López de Gomara, 1941, 2:216). Isolated large communal habitations in small groups, with roofs of palm leaves instead of grass, predominated. Both the Sumu-Taguacas and the Misquito had multiple-family houses.30 In northeastern Honduras, house types generally followed the construction and location indicated above, but no large communal ones are reported. Leaf roofs with wood and cane walls were usual (Serrano y Sanz, 1908, p. 362). The Pacific slope of lower Central America was less uniform. At the eastern periphery, were tree houses in lagoons (Fernández de Oviedo, 1851-55, 3:46; Benzoni, 1857, pp. 237-38). Circular multiple-family habitations with vertical walls similar to the Caribbean area continued, although rectangular structures appeared in western Panama (Mártir de Anglería, 1944, p. 220). In the vicinity of Nata houses were noted for more than one door and a high peak formed by the grass roof, at the apex of which was a long-necked baked clay object (Fernández de Oviedo, 1851-55, 3:13132). 31 The fortress houses of southeastern Costa Rica were circular, of wood and grass, and with extremely high peaked roofs resembling spires. The majority were large enough for 25 people, but some could hold 400. Their unique feature was that they were

ETHNOHISTORY: LOWER CENTRAL AMERICA

raised above ground a half-yard. They also had windows designed for shooting (Fernández Guardia, 1908, pp. 49-50). 32 Multifamily houses supported by posts in the delta section of southern Costa Rica likewise were slightly different (Peralta, 1883, p. 6; Fernández de Oviedo, 1851-55, 3:99). The first single-family houses appear in the Chorotega and Nicarao settlements. They were rectangular, wood and cane, grass-covered constructions without floors. The Nicaraos in particular had kitchens, storage buildings, council houses which also served as dormitories for unmarried men. A raised mound base distinguished the chiefs house from the other dwellings (ibid., 4:109-10). Mud coating was applied to house walls in southern Honduras. Adobe construction, terraces, and flat roofs probably extended south from eastern El Salvador (Ponce, 1873, 1:385). House Furnishings Ordinary house furnishings throughout Panama were few (Fernández de Oviedo, 1851-55, 3:132) and remained scarce in all of lower Central America. Except in southern Honduras and among the Nicarao and Chorotega, the hammock, often of solid cloth (ibid., 3:131), was characteristic. Hung from house posts and carried on journeys in havas (ibid., 3:132), it was used for sleeping in parts of Panama, particularly on the Pacific (ibid., 3:27, 131), and in most of Costa Rica (Fernández Guardia, 1908, p. 51; Fernández, 18811907, 5:374). In other sections of Panama people slept on "barbacoas" of sticks, on straw, cotton cloth, palm leaves, or piled-up earth (Fernández de Oviedo, 1851-55, 3:140; Herrera, 1726, dec. 4, lib. 1, p. 18). Sleeping on leaves or tree bark was customary on the Caribbean watershed of Costa Rica, Nicaragua, and much of eastern Honduras (Fernández, 1881-1907, 5:374; Benzoni, 1857, p. 218). A platform with matting and cotton sheets was used by the

Nicarao (Fernández de Oviedo, 1851-55, 4:109), and a bed of sticks, plaited straw, and woven mat by the Lenca (Herrera, 1726, dec. 4, lib. 8, cap. 5, p. 159). Wooden head rests were common among the Cueva and seem to have been customary in Suerre, Costa Rica (Benzoni, 1857, p. 128), reappearing among the Nicarao (Fernández de Oviedo, 1851-55, 4:109) and Lenca (Herrera, 1726, dec. 4, lib. 8, cap. 5, p. 159). Low wooden seats were used throughout the Caribbean slope. They also formed part of the ranking Chorotega and Nicarao furnishings (Fernández de Oviedo, 1851-55, 4:96, 109). Other common household items were gourd vessels, baskets, planting tools, hunting and fishing equipment. SUBSISTENCE

Hunting Hunting differed principally between the Cueva and Nicarao. The Cueva chief assigned people for this task; the chief and nobles of the Nicarao considered deer hunting, in particular, a special privilege. The usual means were with bow, arrows (including the stunning arrow), lances, hennequen nets, pits covered with leaves and branches, loops of vine, cords, traps, battue, and resin glues. Blowguns with clay pellets were used in Costa Rica and among the Paya and Lenca. Fire was a favorite method to drive game into the open where it was shot with arrows or beaten with sticks, especially in Panama and Honduras. Dogs took part in the chase. In eastern El Salvador and southern Honduras, birds, after feeding, were pursued until exhausted, then caught alive. Bird decoys were common (Fernández de Oviedo, 1851-55, 1:408, 411-12, 417; 4:55; Herrera, 1726, dec. 4, p. 156; Ponce, 1873, 1:331-80; Fernández de Navarrete, 1945, 3:396, 399). 33 32

Cf. with Darien "war houses." See also Sapper, 1902, pp. 80-81; Lehmann, 1920, 2:728-29. 33

217

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

Fishing Tortoiseshell hooks, such as seen in Haiti and other islands, were noticed by Ferdinand Columbus in Caribbean Veraguas.34 Drag- and dip-nets made of cotton, majagua (Hibiscus tiliaceas), silk-grass (probably Karatas plumieri or Furcraea cabuya), and hennequen were used in Panama, and large nets in the San Juan River between Costa Rica and Nicaragua (Colón, 1947, p. 298; Wafer, 1934, pp. 57-59, 77; Fernández de Oviedo, 1851-55, 1:425; 3:136; López de Gomara, 1941, 1:163; Serrano y Sanz, 1908, p. 95; Peralta, 1883, p. 732). 35 Dams or corrals were made with stakes on reefs (Fernández de Oviedo, 1851-55, 1:424); on the Pacific side of Panama and Nicaragua, sticks were used to club fish until dead (ibid., 4:64, 119). The Cueva chief appointed men for fishing (ibid., 3:133). Community participation was characteristic of the Caribbean area for river fishing. There were many methods. Nets were used. Certain vines were crushed in the water releasing poisons which stupified the fish. Mats were placed in a canoe and rustled to make such noise that the fish jumped into the vessel. Fish were also grabbed by hand or shot with bow and arrows with wooden barbed points (ibid., 1:424; Wafer, 1934, p. 77; Ponce, 1873, 1:378). Turtles were turned on their back, and manatí were caught in nets and harpooned (Fernández de Oviedo, 1851-55, 1:432-34). Agriculture The digging stick was universal—and sometimes served as a war implement (see Arms). In Nicoya and Chira Island spades were made by tying a mother-of-pearl shell to a stick with fine twisted cotton thread (ibid., 1:607; 3:110). Extensive agriculture was practiced in Panama, southeastern Costa Rica, parts of the Nicoya Peninsula, and Pacific Nicaragua. In Panama maize, yuca (Manihot utilissima Pohl.), and pineapples were 218

grown on the Caribbean side, coconuts on the Pacific, especially near Burica (Colón, 1947, p. 299; Mártir de Anglería, 1944, dec. 8, lib. 6, pp. 596-98; Fernández de Oviedo, 1851-55, 1:335; 3:133). At least three varieties of pineapples were raised; coconuts were not used for fiber, as in Oceania, because maguey and cotton served this need. Extending from Panama up to the Aguan valley in Honduras, the pejibaye palm (Guilielma gasipaes Bailey) was an essential article of food (Colón, 1947, p. 299). Although planted all over this area, it was of special importance among the Guaymi of Panama and the Talamancan tribes of Costa Rica (Serrano y Sanz, 1908, pp. 86-87), where from 30,000 to 50,000 trees in one stand were reported (Fernández, 1881— 1907, 6:180, 183, 185) and where it was taken over to an extent by Nahua-speaking peoples.36 From Veraguas, Panama, to Matina (Chirripo), Costa Rica, cacao was abundant (Fernández, 1881-1907, 5:157, 370). The Nicarao introduced it from Mexico and monopolized its cultivation in Pacific Nicaragua (Fernández de Oviedo, 1851-55, 1:315; 4:61) where they irrigated and shaded it with "Madera negra" (Gliricidia maculata H. B. & K.), which was pruned to grow straight and tall (ibid., 1:317-18). The harvest lasted from February through April, the beans being dried. The Nicarao also had pole beans and grew a 40-day maize by means of daily irrigation in the dry season. Likewise they cultivated large areas of cotton, maguey, and cabuya (Agave sp.) (ibid., 1:276, 285; 4:36, 105). In the province of Nagrando, Nicaragua, where Mangue speech predominated and 34 Archaeology reveals gold hooks from Panama through southeastern Costa Rica, and historical data place them off the Yucatecan coast. Copper hooks have been found in Nicaragua and central Honduras. 35 Later reports mention silk-grass nets and bone fishhooks in Caribbean Nicaragua. Conzemius, 1932, pp. 69-70. 36 Stone, 1956b.

ETHNOHISTORY: LOWER CENTRAL AMERICA

where the people were to a certain extent under northern influence, beans of many types including yellow and spotted ones were produced in quantity and intensely cultivated (ibid., 1:285) as well as yuca and maize. Red and white beans were procured in quantity by Columbus in northeastern Honduras (Colón, 1947, p. 278). Cultivated plots and gardens throughout lower Central America contained varied plants and trees. These plots were fenced in Panama and among the Nicarao (Fernández de Oviedo, 1851-55, 1:317). Sweet and hot peppers received great care (ibid., 1:275), particularly in Panama and Pacific Nicaragua. Tubers such as yams (Discorea sp.) and small yellow sweet potatoes (especially in Darien, Pacora, and Careta; ibid., 1:273; Serrano y Sanz, 1908, p. 86) were grown. Papayas (Carica papaya L.) and gourd trees also were planted (Fernández de Oviedo, 1851-55, 1:323). Certain sections were noted for special cultivations. These included the American almond (Coumarouna panamensis Pittieri; Dipteryxoleijera Benth) in Castillo del Oro (ibid., 1:355) and dye-yield trees, both red and black, for tattoo or body painting, in Panama, northeastern Honduras, and probably in non-Mexican Nicaragua (Mártir de Anglería, 1944, dec. 3, lib. 4, p. 229). A drug plant was used in Veraguas and in Pacific Nicaragua by the Nicarao (Colón, 1947, pp. 285, 296-97; Fernández de Oviedo, 1851-55, 1:206-07). Fernández de Oviedo, identifying the Nicaraguan plant with one grown in Venezuela and Peru, calls it coca and says it was raised in many places. The identification of this plant with coca is confirmed by Las Casas (n. d., ch. 244, p. 584). A plant called "dactos" (Lemaireocereus griseus (Haworth) Britton and Rose) which was wild in Venezuela was also grown by the Nicarao who seem to have been the only people in Central America to do so (Fernández de Oviedo, 1851-55, 1:311-12). These people also raised alligator pears, guavas, and anonas

(ibid., 1:353, 304-05, 317). The ChorotegaMangue had a monopoly on nisperos (Sapota Zapotilla Coville) (ibid., 1:308; 4:61). Slash-burn preparation of land was usual everywhere. Planting and weeding was done by women in Darien and southeastern and northeastern Costa Rica and by men in Pacific Nicaragua (Wafer, 1934, pp. 9293; Fernández Guardia, 1908, p. 50; Fernández, 1881-1907, 5:373; Fernández de Oviedo, 1851-55, 4:39). Maize (purple, red, white, and yellow) was planted at new moon and the beginning of the rainy season (ibid., 1:268). The seed, dampened several days before planting, was carried in gourds tied around the waist or neck. The planters walked in line, making 4-7-cm. holes with the digging stick, placing four or five grains in each and refilling the hole with the foot (ibid., 1:264-65). The shoots were weeded while still small. Poisonous yuca was raised only in Darien (Wafer, 1934, pp. 62-63). Nonpoisonous yuca was cultivated on the mainland and planted from the full to the new moon. Of the two methods, the commonest was to place six or more cuttings in round mounds; in level ground, two or more cuttings were placed together without mounding them (Fernández de Oviedo, 1851-55, 1:269-70). Food and Food Preparation Food was roasted, stewed, cooked in soup, smoked, or salted. The three basic foodstuffs were maize, tubers (of which yuca was the most important), and pejibaye (see Agriculture). All were prepared by boiling, roasting, or made into a kind of beer (see Beverages). Pejibaye was also dehydrated to last between seasons (Serrano y Sanz, 1908, p. 87); maize was kept in husks. Women ground maize with water on a stone, made a loaf of this paste, wrapped it in the husk of a leaf, and boiled or roasted it. The Mexican tribes introduced the tortilla or thin maize cake into Central America, at least as far as the Nicoya Pen219

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

insula on the Pacific and possibly to Taguzgalpa on the Caribbean. It differed from the loaves in that the hull was removed and it was easier on the teeth (Fernández de Oviedo, 1851-55, 1:266-67, 270). On the Pacific side of Nicaragua and in Darien, maize flour was mixed with water for trips (Wafer, 1934, p. 91). Yuca was also used for soup. Bread was made only in Darien from the poisonous yuca (see Agriculture)37 (Fernández de Oviedo, 1851-55, 1:270-72; Wafer, 1934, pp. 62-63). Chili peppers were mixed with fish, meat, soup, and the leaves put in sauce; sweet peppers were cooked; and honey was used for sweetening. Diet also included a variety of fruits and vegetables including cooked flowers, soups of greens with calabashes and peppers, mamey, jocotes, sour sap, roasted maguey roots, and papayas (Fernández de Oviedo, 1851-55; 1:278-79, 303-06, 308, 323). In Pacific Nicaragua, Nicoya, Chira island, Chiriqui, and Talamanca, cacao was both eaten and drunk. The Nicarao painted the face with the butter, taking it off bit by bit with their fingers and sucking it (ibid., 1:318). The Talamancas used it in the form of tablets (de Vilaplana, 1763, p. 46). Cacao was prepared three ways. The Nicarao took 30 toasted beans and ground them with a pint of water and achiote (Bixa orellana L.) to give a blood color, boiled, stirred, and then poured the mixture from one gourd to another to make foam (Fernández de Oviedo, 1851-55, 1:316-17). Cacao with achiote was used from Poton territory through Nicoya (Ponce, 1873, 1:392). The Chorotega used less water, very slow fire, a swizzle stick, and strained the oil through a woven palm cloth (Fernández de Oviedo, 1851-55, 1:319). In Chiriqui, the pod and the beans were some37 This may be why graters for yuca do not appear archaeologically. Poisonous yuca may have been introduced into Darien by the Carib or the Arawak.

220

times eaten raw. To make paste or butter, the beans were toasted, the hull removed, and the beans ground without water. This bean meal was cooked on a slow fire with a gourd of boiling water, stirred with a stick or cane until it did not stick to this, beaten until the oil appeared, and then the oil removed bit by bit with a spoon. The remaining cacao was stirred some more and as the oil came up it was taken off with a feather (ibid., 1:321). Fish, shellfish, fowl, and meat were consumed everywhere except in the province of Paris in Panama. Here the warriors ate only iguana and fish; the farmers ate meat (Fernández de Navarrete, 1945, 3:404; Herrera, 1726, dec. 4, p. 20). Frogs were roasted from Darien through Veraguas, Nicoya, and Nicaragua (Fernández de Oviedo, 1851-55, 1:355; 3:110; Wafer, 1934, p. 103). Fowl was generally wild and included quail, pigeons, and curassow (ibid., pp. 6970; López de Gomara, 1941, 1:125; Ponce, 1873, 1:380). Meat meant armadillo, rabbit, deer, monkey, and wild pig. Danta was also eaten by most people. Lard was obtained from jaguar fat by the Nicarao (Fernández de Oviedo, 1851-55, 1:402-12). In Pacific Nicaragua turkeys (ibid., 1:49) and barkless dogs were bred to be eaten. The consumption of human flesh for pleasure was limited to the Nicarao and Chorotega, where people were sold in the market for this purpose. The head was cut off, the body chopped into small pieces and put into pots to stew with salt and chili peppers. It was eaten with maize (ibid., 4:14, 37, 5 1 52). These people also had ceremonial cannibalism (see Religion). So did the Taguacas of northeastern Honduras (Serrano y Sanz, 1908). Salt generally was made by boiling or evaporating sea water (Fernández de Oviedo, 1851-55, 1:173; Fernández de Navarrete, 1945, 3:400; Wafer, 1934, p. 77). The Paya in northeastern Honduras made

ETHNOHISTORY: LOWER CENTRAL AMERICA

salt from the ashes of a wild palm, which they chipped and burned (Serrano y Sanz, 1908, p. 368) similarly to the Amazonian method (Staden, 1928, pt. 2, ch. 11). The Sumu burned wood, dipped the hot brand into the sea, and wiped off the condensed crystals on a leaf (M. W., 1852, p. 302).

55, 4:96) and in Darien. In the latter region the lighted portion of the cigar was stuck in the mouth of a boy who blew the smoke through the cigar into the faces of seated men who collected the smoke in their hands to inhale it (Wafer, 1934, p. 63). 39

BEVERAGES AND NARCOTICS

From the Cuevas to the province of Paris, Pacific Panama, young girls wore no clothes. Through the Asuero peninsula lower-class women's dress consisted of a cotton skirt from waist to knee; that of the upper classes reached to the ankles. In Paris only, these skirts were painted. 40 Men, except in Coiba (Jijon y Caamaño, 1938, 2:17-18) (see Trade), wore a penis cover of shell, calabash, knot of wood, or gold belted by a string tied in two holes. For festivals they wore a tunic-like fringed robe often studded with gold plaques or a small apron with feathers (Fernández de Oviedo, 1851-55, 3:126; Herrera, 1726, dec. 4, lib. 1, p. 18; Wafer, 1934, pp. 84-85). Warriors on Coiba island wore exceptionally thick cotton corselets from the shoulder to below the knee. West of the Asuero peninsula through Burica men went naked when they wished to, and women wore loincloths; otherwise dress was similar to the Chorotega. In southeastern Costa Rica, very fine cotton cloth with openwork was customary. Women wore a parka-like garment which covered the head and reached the feet; men wore cotton breechclouts with a small apron in front (Fernández, 1881-1907, 5:374; Fernández Guardia, 1908, p. 51). Cotton cloth was also used in the central plateau (ibid., p. 13). West of the plateau, on the Pacific mainland, bark cloth was worn by women and tunics of bark by men 41 (cf. the Caribbean coast). Cockburn noted later (1779, pp. 107-08) similar dress on Chira island. This island, when first seen by the Spaniards, was reported to have been inhabited only by warriors. It was the one place in the Nicoya region where the

In the Caribbean regions pejibaye was boiled, strained, and drunk without fermentation (Fernández, 1881-1907, 5:382), but an alcoholic drink was also made from it (ibid., 6:307; Fernández de Oviedo, 185155, 1:334; López de Gómara, 1941, 1:160). Yuca beverages, both alcoholic and nonalcoholic, seem to have been universal (Fernández de Oviedo, 1851-55, 1:272). Fermented drinks were everywhere made from maize; in the Caribbean area from pineapples; and in Nicaragua from jocotes (Spondias purpurea L.) and guasuma (Guazuma ulmifolia Lam.) (ibid., 1:298, 307). Among the Nicarao, to drink chocolate was the exclusive privilege of the chief and nobles (ibid., 1:316-17; see Food and Food Preparation). Chewing an herb identified as coca (see Agriculture) was prevalent in Veraguas, where it was mixed with a powder. Among the Nicarao the same herb was used with ground seashells. Mixing was done with a little stick and the mixture carried in a small gourd or bag around the neck (Colón, 1947, pp. 285, 296-97; Fernández de Oviedo, 1851-55, 1:206-07),38 a method practiced in Colombia also (Matienzo, 1910, chs. 44-51). Cigars were smoked by the Chorotega (Fernández de Oviedo, 185138 This herb was chewed in Venezuela and Peru. 39 Archaeological evidence yields single- and double-pronged snuffers in Costa Rica which recalls the Haitian receiver for cohiba mentioned by Fernández de Oviedo, 1851-55, 1:130. See Stone, 1958, fig. 19,a-b. 40 Painted cloth was used in many places in Panama and in Costa Rica to cover the dead. See Death and Burial. 41 Bancroft, 1882, 1:751.

DRESS AND ADORNMENT

221

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

penis hung free (Fernández de Oviedo, 1851-55, 3:108, 110). Female Corobici used cotton loincloths, but males went naked (Mártir de Anglería, 1944, p. 231) with the penis tied up tightly with many strings (Fernández de Oviedo, 1851-55, 3:109) or used breechclouts. In the provinces of the Chorotega, Nicarao, and Nagrando (in part Maribio) men went naked when they chose (ibid., 4:108; Las Casas, n.d., ch. 242, p. 577; Fernández de Navarrete, 1945, 3:399). Dress consisted of multicolored fine cotton sleeveless shirts and white belts, as wide as the palm of the hand (Fernández de Oviedo, 1851-55, 4:98) (see above: west of the Asuero peninsula). These belts were twisted and wrapped around the body many times from breast to hips, with the end brought between the legs and fastened to the belt to act as a loincloth (ibid., 4:38-39, 55, 98). Often a cloak was carried under the arm. Footgear was a kind of "alpargato" made of double deerhide tied to the ankles with cotton cords passing between the toes. Warriors used sleeveless cotton doublets and headgear in Pacific Nicaragua (ibid., 3:104). Chorotega women and those from the province of Nagrando wore an elaborate variation of loincloth arranged as an apron tied with a single string at the back, passed between the legs, and inserted beneath the string in front (cf. Fernández de Oviedo, 1851-55, 4:108; López de Gomara, 1941, 2:217). 42 Similar dress supposedly extended to the Asuero peninsula, Panama (see Fernández de Oviedo, 1851-55, 4:108). Nicarao women wore skirts like the Cuevas but used cotton ruffs to cover the breasts. They also wore alpargatos (ibid., 4:38; López de Gomara, 1941, 2:217). Ponce, who was in Nicaragua a little over half a century after Fernández de Oviedo, found that women's dress from the coast of southern Honduras to Granada was a black squared short sleeveless cassock worn with 42 Lehmann, 1920, p. 797.

222

a point in front and back and held in place by wide belts and a colored knee-length skirt (Ponce, 1873, 1:352; cf. Herrera, 1726, dec. 4, lib. 8, first p. 155, who gives a description of Lenca dress). On the Caribbean side, bark breechclouts served as dress of both sexes from Veraguas through Guaymi territory (Serrano y Sanz, 1908, p. 96). Bark cloth continued in use in northeastern Costa Rica. It was made into breechclouts and a short narrow waistcoat-like garment for men and into knee-length skirts for women. Cotton skirts were reserved for gala occasions (Colón, 1947, p. 280; Fernández, 1881-1907, 5:373; 10:501). Long bark-cloth breechclouts with a bark tunic-like shirt were common to the Sumu, Misquito, and probably the other inhabitants of Caribbean Nicaragua (M. W., 1752, pp. 307-08). Sometimes clouts were cotton with colored designs. For festive occasions the cotton tunics of ranking men reached to the knees and were embroidered with feathers (see Cueva ceremonial dress above) and worn with a long belt. Women wore bark knee-length skirts tucked in at the waist. In northeastern Honduras, by Cape Gracias a Dios, most people went naked (Colón, 1947, p. 277), although warriors had quilted cotton armor. The Paya used bark breechclouts; and for ceremonial occasions they had belted robes of bark with sleeves and, sometimes, a mitre-like bonnet (Serrano y Sanz, 1908, pp. 356-57). Other peoples had colored cotton loincloths and sleeveless shirts reaching to the navel. Individuals of high rank put red and white cloths on their head. In eastern Central America, gold was favored as an adornment, particularly in Darien, where ranking women used gold bars to hold up the breasts and men wore gold helmets. Gold headbands, ear rods, nose pendants, cuffs, bracelets, rings, and many necklaces were worn for festivals. Gold ornaments continued through Costa

ETHNOHISTORY: LOWER CENTRAL AMERICA

Rica (Colón, 1947, pp. 282, 284; Fernández, 1881-1907, 5:317; 7:387; Fernández Guardia, 1908, pp. 22, 36) into Caribbean Nicaragua where gold disks were worn by the Misquito (M. W., 1752, p. 308). Some Chorotegan men had removable hammered gold or bone buttons between the chin and lower lip; women wore gold disks (Fernández de Oviedo, 1851-55, 3:103; 4:98). The Nicarao are said to have put great importance on gold work (see Social Organization), but no jewelry is described. Feathers (in Caribbean Nicaragua dyed from secretion of poisonous frogs), feathered headbands, stone pendants, whale teeth (in Darien), animal and fish teeth, and bone and shell ornaments and beads were common to lower Central America. Removable lower-lip ornaments of shell were used by the Misquito (Dampier, 1699, 1:32; M. W., 1752, p. 308). The Nicarao and Caribbean Nicaraguans wore jadeite beads and labrets. The Nicarao had nose rings, the others had nose rods (Fernández de Oviedo, 1851-55, 4:38, 256; López de Gomara, 1941, 2:217). Enormous earplugs were customary among the Nicarao and in Caribbean Nicaragua and Honduras. The Spaniards named the latter coast "Coast of the Ears" because of this (Colón, 1947, pp. 277-78). Body painting, especially with red and black, was of prime importance for ceremonies, galas, and war throughout the region (Fernández de Oviedo, 1851-55, 1:204; 3:138-39; 4:111; Wafer, 1934, p. 83; Fernández de Navarrete, 1945, 3:404; López de Gomara, 1941, 2:163; 2:217; Colón, 1947, p. 278; Benzoni, 1857, pp. 13233; Mártir de Anglería, 1944, p. 229; Serrano y Sanz, 1908, p. 90; Exquemelin, 1893, p. 252). Tattooing by scarification and paint, with each chiefs personal symbol used by his followers, was current in Darien and among the Chorotega and Nicarao (Wafer, 1934, p. 83; Fernández de Oviedo, 1851-55, 1:204; 3:111, 138-39; 4:38). Tattooing by

cautery was practiced by the Nicaraguan Sumu and in northeastern Honduras (Colón, 1947, p. 278; Pirn and Seeman, 1869). In general, body hair was removed by pincers; but in parts of Darien it was done with a concoction of herbs and powdered ants (López de Gomara, 1941, 1:162; Wafer, 1934, p. 79). The only reports of beards refer to the warriors of Escoria on the Asuero peninsula and the Nicarao (Relación . . . Espinosa, 1892, 2:479; Fernández de Oviedo, 1851-55, 4:111). The care of the hair was important among the Cueva, Nicarao, Chorotega, and Misquito. All used combs (ibid., 4:108-09; Exquemelin, 1686, 2:268-69). In Darien, both sexes took pride in their long black hair, the women tying it just behind the head. In Caribbean Costa Rica, men's hair was long, braided and rolled around the head with bangs, that of women was cut short (Colón, 1947, p. 280). In Nicaragua and the Nicoya peninsula, males shaved half the head from front to back leaving a band from ear to ear behind the crown. Nicarao, Chorotega, and Chontal (see Languages) warriors who triumphed in bodyto-body fighting left on top a crown about 2.5 cm. high; in the center of this remained a tassel of hair higher than the crown. Such warriors were considered as nobles (see Social Organization) (Fernández de Oviedo, 1851-55, 4:38), suggesting "Nicaraozation." Non-warrior Chorotegan men wore a cotton band with a plait hanging down the back or gathered on top of the crown. The Chorotega women parted the hair in the center, combing it in two stiff plaits (ibid., 3:109; López de Gomara, 1941, 2:217). TRANSPORTATION AND COMMUNICATION

Burdens were carried by means of a tumpline except eastern Panama where a carrying pole was used (Wafer, 1934, fig. facing p. 140). Cueva men and women of rank were carried in a hammock litter along tree223

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

bordered roads (Fernández de Oviedo, 1851-55, 3:31, 126). The Caribs were famous sailors in the province of Cueva. Their large sea-going canoes carried more than 70 men and were propelled by paddles and cotton sails (ibid., 3:159; Mártir de Anglería, 1944, p. 612) while a coxswain sang in time to the oars. Crafts 45 feet long and rowed by 12 oarsmen seated in pairs navigated the San Juan River and Caribbean coast of Suerre (Peralta, 1883, p. 734). The Misquito also had sea-going canoes (Raveneau de Lussan, 1869, p. 440). In Fonseca Bay, pointedprow dugouts, wider at bottom than top for protection against waves and manned by eight paddlers, were common. These had cotton or woven grass sails. Passengers sat under individual mat covers. Paddles were long with an oval base and manipulated by men standing between each passenger (Ponce, 1873, 1:375-76). Although canoes were used in the Gulf of Nicoya, the inhabitants of Chara and Pocosi islands only had rafts.43 These were made of four or six logs tied with grass rope (vines) at the ends and in the middle; narrower logs were tied crosswise over them. The base of the paddles was of mother-of-pearl (Fernández de Oviedo, 1851-55, 3:110). In Chiriqui, Pacific Panama, the small bark canoes used in rivers (Cockburn, 1779, pp. 143-44) were portaged by women. On the Caribbean side, balsa rafts, made in the manner just described but tied with fiber cord and pegged with wood, were employed for fishing and for crossing rivers (Wafer, 1934, pp. 59-60). Wooden dugouts and hard black wood paddles were employed for river travel in Panama and among the Misquito, Sumu, and Paya in Honduras and eastern Nicaragua (Serrano y Sanz, 1908, pp. 370-73).

43 These rafts could have been of any wood; the Spanish merely termed the logs "maderos."

224

MANUFACTURES

Ropes, string, and hammocks were made of hennequen and cabuya in Panama, Pacific Nicaragua, and Costa Rica (Fernández de Oviedo, 1851-55, 1:277-78). Cotton hammocks were also made in Panama. Baskets for clothes and salt were fashioned by women from bijagua stalks, woven double or interwoven with the leaves. Cups and vessels of gourds were made throughout the area. Eastern Panama added gold handles (ibid., 1:295-96). Lance poles were likewise famous in Panama. Excellent cotton cloth (see Dress and Adornment) was made by men in southwestern Costa Rica (Fernández Guardia, 1908, p. 51). Among the Chorotega and Nicarao women wove fine cotton material; Lenca weaving was used for tribute to the Spaniard (Fernández de Oviedo, 1851-55, 4:36; López de Gomara, 1941, 2:217; Herrera, 1726, dec. 4, lib. 8, pp. 155, 161; Fernández Guardia, 1908, p. 51). Southeastern Costa Rica and the Chorotegans were noted for pottery (see Trade) (Fernández de Oviedo, 1851-55, 4:105, 108; Fernández Guardia, 1908, p. 51). From Darien through Suerre on the Caribbean, and on the Pacific through southeastern Costa Rica particularly, gold working was famous. The metal was mixed with copper and silver and often gilded by using plant juice (Fernández de Oviedo, 1851-55, 1:189). The art appeared to be new among the Nicarao as special rank was given to goldsmiths (see Social Organization ) . LIFE CYCLE

Birth and Puberty Taboos veiled birth and puberty. Birth generally took place in an isolated shelter, always near a stream. Taboo had to be terminated by a priest-doctor before mother and newborn could come into contact with other people. In eastern Panama, right after the event women did not exercise

ETHNOHISTORY: LOWER CENTRAL AMERICA

and stopped eating for fear of getting fat. Another woman, probably a priest-doctor, was usually present at or immediately after the birth and carried the mother to the river, washing her and the child (Fernández de Oviedo, 1851-55, 3:134; Wafer, 1934, p. 93; Herrera, 1726, dec. 3, p. 160; Cockburn, 1779, p. 104). 44 Among many of these people the umbilical cord represented a tie with the earth. It was cut or mashed by the mother and secretly buried at the spot of birth. Weaning was late (Fernández de Oviedo, 1851-55, 3:110; Herrera, 1726, dec. 3, p. 160). 45 From Darien westward, female puberty was a period of taboo, with slightly varying customs such as removal to the forest and haircutting. The aid of a priest-doctor seems to have been necessary to restore the individual to normal life (Serrano y Sanz, 1908, p. 131). 46 Physical Deformation Artificial deformation of the head was practiced by the Sumu, Ulva, and Nicarao (Fernández de Oviedo, 1851-55, 4:54; Μ. W., 1752, pp. 304-05; Collinson, 1870, pp. 149-50). Las Casas cites artificial flattening of the head east of Darien, and this recalls Wafer's statement that the new-born in Darien were strapped to a board for the first month to make the child "grow very straight" (Las Casas, n.d., vol. 2, ch. 244, p. 586; Wafer, 1934, p. 94). 47 The female Sumu employed a kind of tourniquet above the ankle and below the knee to bulge the calf muscles (Dampier, 1699, 1:32). Marriage Prearranged marriages were customary. The first degree of consanguinity was taboo. Most peoples were polygamous, the Paya on the Caribbean and the Nicarao on the Pacific excepted. The Nicaraos, however, did not practice what their law implied, generally having two wives with an un-

limited number for the chief (Fernández de Oviedo, 1851-55, 3:133). When of age all lived in a single dwelling with the principal wife holding precedence over others (Herrera, 1726, dec. 4, lib. 8, pp. 158-59). Change of wife was not frowned upon but had to take place during her menstrual period and usually not before two years after marriage. There are indications that in certain sections, such as the Voto-Corobici area, polyandry might have been the rule at some remote period. The custom in the independent towns of Pacific Nicaragua and the Nicoya Peninsula for women to pick their husbands from a group of young men who ate together suggests this (López de Gomara, 1941, 2:217-18) as does the license given Nicarao married women to sleep one night a year with whomever they wished (Fernández de Oviedo, 1851-55, 4:102). 48 Marriage ceremonies seem to have consisted in one of the contracting parties placing a hand on some part of the other, this gesture then followed by a feast. The Nicarao joined fingers. They placed such importance on the marriage ceremony that even to repeat it courted the risk of exile and loss of inheritance. Jus primae noctis was practiced by Chorotega and Nicarao chieftains (ibid., 4: 37, 97; Herrera, 1726,

44 See also Gabb in Fernández, 1881-1907, 3:341-42. 45 No early source mentions the couvade, but vestiges of this are reported among the Boruca and Guatuso of Costa Rica and the Sumu and Misquito of Honduras and Nicaragua. See Pittier, 1895, p. 20; Sapper, 1902, p. 231; Conzemius, 1932, p. 151. 46 Stone, 1962. 47 Cf. Conzemius, 1932, p. 28. 48 The vestige of polyandry noted by Sapper (1902, p. 231) among the modern Guatuso also backs such a supposition. It must be remembered that the Corobici territory extended from the Nicoya Penisula to at least Ometepe Island in Lake Nicaragua (see Language) and that the present-day Guatuso have Corobici and Voto blood. See Stone, 1956a, pp. 503-04.

225

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

dec. 4, lib. 1, p. 18; Serrano y Sanz, 1908, p. 132). 49 In eastern Panama, some wives were buried with the chief (see Death and Burial). Among the Guaymi, the heir, be he chief or not, took the wives of the deceased as his own. The sister of a dead wife was also married by the widower. This and the marriage of the widow to her brotherin-law were customary over the entire Caribbean littoral (Fernández de Oviedo, 1851-55, 3:154; Serrano y Sanz, 1908, p. 29). 50 Abortion and Sterility Abortion through consuming certain herbs or fruits was common from Cueva westward. Very significant was the use of herbs to produce temporary sterility (Fernández de Oviedo, 1851-55, 3:134). 51 Death and Burial Through most of lower Central America the sick were left to die alone, either in the forest or in special shelters. The bodies of mourners were painted red and black, songs recounted the deeds of the deceased, dances were held, and food and drink were provided—all as parts of the ritual taking place at the time of death (rarely), at the end of a year after death (commonly), or at both times (see Fernández de Navarrete, 1945, 3:394-95; Herrera, 1726, dec. 4, lib. 1, p. 19; Fernández de Oviedo, 1851-55, 1:12830; 3:156; Serrano y Sanz, 1908, pp. 92-93). In Panama commoners were buried with maize, wine, and cloth, a custom indicated by archaeology to have been widespread. Some, probably the poor or slaves, just wandered away and died, the body left to birds or beasts of prey (López de Gomara, 1941,2:202-03). Food, drink, and personal property were generally buried with the corpse, although those who did not believe in immortality put nothing in the grave. Goods included painted cloth52 and gold ornaments in Panama and Costa Rica. The use of gold with 226

the dead extended into eastern Nicaragua, where gold dust was mixed with clay and applied as a mask or mummy face-covering or put on burial urns (O. W. Roberts, 1827, pp. 299-300). Dehydration of the corpse over fire and housing of the dried body was the treatment given a dead chief in many parts of Panama, Caribbean Costa Rica, Nicaragua, and in the extreme western limit of northeastern Honduras, the Olancho valley. Housing without drying was customary to the Honduran Paya, who in addition placed a canoe on top of the dead (Fernández de Navarrete, 1945, 3:394-95; Fernández de Oviedo, 1851-55, 3:155; Colón, 1947, p. 282; National Archives of Seville, Guatemala II, Doc. 38 (2); Serrano y Sanz, 1908, pp. 361-62). Burial of chiefs seated on stools in woodlined graves was also known in Panama. Wives, retainers (see Social Organization), and slaves were frequently interred with the ruler, either voluntarily or forcibly. Drunkenness and suicide through poison helped this end (Fernández de Oviedo, 1851-55, 1:134; 3:154-56). The custom of burying slaves or retainers with the chief extended into Caribbean Nicaragua (Fernández Guardia, 1908, pp. 30, 44, 50; Exquemelin, 1686, 2:275). Bone burial was practiced from Panama through the Taguaca (Sumu) territory of Honduras (Serrano y Sanz, 1908, pp. 93, 135; Fernández, 1881-1907, 5:373; Exquemelin, 1686, 2:274). Cremation with personal effects was known in Panama. Among the Chorotega 49

See also Alvarez Rubiano, 1944, p. 231; Conzemius, 1932, p. 147. For archaeological evidence in Costa Rica see Stone, 1962. 50 See also Conzemius, 1932, p. 146. 51 In 1894 this was prevalent in Talamanca and was noted during the 19th century among the Misquito and possibly the Sumu. Stone, 1949, pp. 26-27; Conzemius, 1932, p. 152. 52 The wrapping of bodies, particularly those of personages, in painted cloth was practiced until the beginning of this century in Talamanca and Boruca, Costa Rica.

ETHNOHISTORY: LOWER CENTRAL AMERICA

on Chira island, Costa Rica, the body was wrapped in leaves and burned to the accompaniment of singing and dancing, and the ashes were buried in a hole (Cockburn, 1779, p. 108); among the Nicarao the ashes of chiefs were buried in urns in front of the dwelling (Fernández de Oviedo, 1851-55, 4:48-49). The Nicarao wrapped dead children in cloth and buried them with food at the door of the house. If a person died childless, his personal property went into the grave with him; otherwise possessions were left for the offspring (see Inheritance). The death ritual was the one common to lower Central America, but the dead were remembered for 20 to 30 days instead of a year and clay figures were broken on top of the grave (ibid., 4:48-49, 93-95). SOCIAL, POLITICAL, AND ECONOMIC

killed him. Constables administered justice to those of lower rank. The hands of thieves were cut off and hung around the neck (ibid., 3:129-42). The Nicarao thief was tied and brought to the place of theft, where his hair was cut and he stayed bound until he paid, made restitution for the felony, or his hair grew out (ibid., 4:51). Boys stoned male perverts. A rape offender was tied in the victim's father's house until he paid or made up satisfactorily for the wrong; otherwise he became the slave of her or her parents. Male adulterers lost their inheritance. Females taken in adultery were returned to their parents, but if a male slept with the master's or noble's daughter, the pair was buried alive by their relatives. A free man who killed another had to pay the parents or wife with a slave, clothes, or possessions.

ORGANIZATION

Social Organization

Land Tenure and Inheritance In spite of reference to fenced plots in Panama (see Agriculture), land tenure is not mentioned in the documents except for the Nicarao. Elsewhere land for cultivation seems to have been allotted by the chief or was used by a given group within tribal boundaries. In Panama, rank and possessions were inherited by the eldest son or, if none, by the daughter. If, however, there were only girls, the sons of the second daughter inherited (ibid., 3:133). The Nicarao individually owned land but it was not negotiable. If a man left his community, neither his land (ibid., 4:55) nor house (López de Gomara, 1941, 2:218) could be sold but had to be given to his relatives. All possessions went to the children if legitimate and born at home (Fernández de Oviedo, 1851-55, 4:48 (see Death and Burial).

The Cuevas had a chief, men of high rank, those of less rank, commoners, and slaves (ibid., 3:129-31; Herrera, 1726, dec. 4, lib. 1, p. 18). A person of low rank and occasionally commoners could ascend through marriage. Generally, commoners could change social position only if lucky in war. The rigidness of the system is made obvious by the belief that the only way a commoner could reach heaven was to commit suicide at the death of a chief (Fernandez de Oviedo, 1851-55, 3:154). Male perversion was important among the ruling class, and prostitution common (ibid., 3:134). There were four distinct names for chief in this province. At least one, "guaxiro," was Arawak, equivalent to the Santo Domingan term "cacique" (ibid., 3:129). There was only one term, "queví," to designate the very highest. It is this last who participated in religious rituals, received service but not tribute, but could also fish, hunt, and plant when not at war (idid., 3:130-33). 53

Crime and Justice In eastern Panama, the chief announced the misdeed of a person of rank and gave him the first blow on the head, but others

53

All this suggests that the Cuevas had con227

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

Parts of Cueva lacked chiefs but had war captains for battle (Serrano y Sanz, 1908, pp. 116, 130). This arrangement occurred in modified form among the Guaymi and the Talamanca peoples of Costa Rica, where chiefs had no power except during war. The Guaymi likewise had a lesser authority, "cabra," and commoners (Serrano y Sanz, 1908, pp. 103-04). Independent populations without rulers appear at the western periphery of the Cuevas on the Pacific. This includes Purulata, Chiame, and Coyba (Coiba), places where the "language of Cuba" was spoken (Herrera, 1730, dec. 1, lib. 8, p. 20; 1726, dec. 4, lib. 1, p. 18; Fernández de Oviedo, 1851-55, 3:12). The Cuevas permitted only the male to inherit position (see Land (ibid., 3:133). Tenure and Inheritance) Chape, on the Pacific slope, with a chieftainess demonstrates a concept foreign to "Cuevarized" Panama. A female ruler was also seen among the Voto in Costa Rica (Fernández Guardia, 1908, p. 18). Independent towns governed by a council of elders chosen by vote and assembled from different communities seem to have been limited to the province of Nagrando in Pacific Nicaragua. It is a form of administration familiar to the highlands of western Central America (Herrera, 1726, dec. 4, lib. 8, p. 158). Nagrando is associated historically with the Chorotega (Fernández de Oviedo, 1851-55, 4:33, 61; Torquemada, 1943, 1:323), Nicaraos and Maribio (Fernández de Oviedo, 1851-55, 4:37); but "tacacho," claimed to have been the mother tongue of this region (see Language), was also spoken. This suggests an acculturation which necessitated some social changes. A stratified society, common to both the quered earlier aborigines, who were originally more like lowland South America in their social organization but had slowly become "Cuevarized." 54 Social stratification in the Caribbean area was based on the rank of the clan, not of the individual. Priests and chieftains could come only from given groups and certain tasks fell to set clans.

228

Chorotega-Mangue and Nicarao, was maintained (see Hunting); but in governmental matters the council, elected from a number of communities, held the power and chose a captain-general for war (ibid., 4:36). Secret councils of hereditary chiefs to obtain from oracles advice and prophecies regarding the welfare of the people were current among the non-"Nicaraoized" Chorotega-Mangue (ibid., 4:74-75, 82). These people were noted for being under the dominance of their women (ibid., 4:60). Hereditary rulers who spent a year in the temple and governed with an elected secret council of "nobles," which changed every four moons, were characteristic of the Nicarao. This society was highly stratified with vassal rulers, nobles, officials, priests, commoners, prostitutes (in houses run by "madams"), and slaves. Each class was subdivided into distinct categories. The extraction of tribute was also typical of the Nicarao. Women held a subordinate position (ibid., 3:134; 4:37, 61, 102). The Nicarao seem to have asserted great influence over some of the Chorotega-Mangue, at least in Nicoya (ibid., 4:37-38, 134). Matrilineal clans were the dominant form of society from Chiriqui lagoon through most of Atlantic Costa Rica, and they are reflected in large multifamily habitations. Reports of missionaries and conquerors repeatedly indicate that an entire family group inhabited a dwelling and followed rigid taboos on mingling (Fernández, 18811907, 9:23; Peralta, 1883, pp. 395-432). The same system was common to the Pacific highlands where concentration in fortified sites for safety emphasized the power of the chief more than on the Caribbean (see Habitations: Villages) (Fernández Guardia, 1908, p. 49; Fernández, 1881-1907, 5:373; Serrano y Sanz, 1908, p. 95; Peralta, 1883, p. β). 54 Clans continued in Caribbean Nicaragua where Vázquez notes that the only distinction between divers nations were families and family groupings (Vázquez, 1937-44,

ETHNOHISTORY: LOWER CENTRAL AMERICA

4:202). In 1715, the multifamily houses of the Misquito also suggest clan organization (Peralta, 1898, p. 80). Caribbean Nicaragua and northeastern Honduras appear to have had chiefs only in war time, when they were generally chosen by a council of elders (Vázquez, 1937-44, 4:121). 55 Trade Interchange of goods without money within a designated area or through itinerant traders (Peralta, 1883, p. 735) was of utmost importance in most of this region. Columbus found a 50-league trading tract from Chiriqui lagoon eastward (Colón, 1947, p. 286; Fernández, 1881-1907, 5:157). In Darien, when not warring, life was dominated by trading (Fernández de Oviedo, 1851-55, 3:140; López de Gomara, 1941, 1:163) which suggests a matriarchal society in contrast to the Nicarao. Goods were transported great distances in canoes or on backs of slaves.56 Even the Lenca had trading areas during periods of peace (Herrera, 1726, dec. 4, lib. 8, p. 156). Items of exchange were shell penis covers (in eastern Panama), pottery, raw cotton, cotton thread, woven and painted cloth, shell beads, gold figures, hammocks, slaves, dogs, salt, salted fish, dantas, wild pigs, birds, feathers, achiote, maize, cacao, resin, and danta-hide shields (Fernández de Oviedo, 1851-55, 3:140; Peralta, 1883, p. 54; Torquemada, 1943, 1:335; Fernández, 1881-1907, 5:157-58; Herrera, 1726, dec. 4, lib. 8, p. 156; Informe, 1697). The Nicarao and Chorotega had markets located in the principal towns (Fernández de Oviedo, 1851-55, 4:37). These were the only peoples who used a medium of exchange, cacao beans, which set a value on 55 The exceptions were probably the Lenca and Mexican tribes who penetrated this woodland area. The Lenca were also found in the region of Fonseca Bay where they may have retained the organization noted for highland Honduras (see Herrera, 1726, dec. 4, lib. 8, p. 159). 56 For example, in Panama, the Spaniards were told about Ecuadorean balsa rafts and llamas.

commercial items (ibid., 1:316). All negotiations were made by women although the supreme market authority was a man chosen by the secret council (see Social Organization). However, virgin boys, men of like tongue, or allies could enter the market (ibid., 4:37, 54, 104). Articles of sale were children (for cannibalism), gold, cloth, game, and agricultural products. Black dye was made by the Chontales and exchanged for goods or cacao beans in the market. Warfare and Arms Although wars were primarily fought over land boundaries, in Costa Rica the need for sacrificial victims was a predominant motive (ibid., 3:129; 4:53; Fernández, 1881-1907, 5:157-58). The Nicarao also fought for slaves and for the sake of seeing who was the more able (López de Gomara, 1941, 2:218). Contrary to the rest of the area where the strength of a chieftain depended on his ability in battle, the Nicarao chief did not participate in fighting but substituted instead a special captain. Each soldier had the privilege of keeping what he took in battle (Fernández de Oviedo, 1881-55, 4:53). Fighting arms throughout eastern Central America, with the exception of Tojar island in Chiriqui Lagoon, consisted of macanas, stones for hurling, lances, and darts. On Tojar, only lances are reported. Throwing sticks for darts were found in eastern Panama and southeastern Costa Rica. Generally, the darts were of cane with barbs of black palm or animal or fish bones. Among the Cueva there was a dart type with holes which whistled when flying through the air. Poisoned darts were found in Caribbean Nicaragua and Honduras among the Sumu-Taguaca and Lenca, poisoned arrows on the Pacific border of Costa Rica-Panama at Burica. Herrera even talks of a poisoned macana in Honduras. Nonpoisonous barbed black palm arrows were used for war in Burica, in Talamanca, where fire arrows were likewise known, and 229

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

by the Chorotega, Nicarao, Paya, and Lenca. The Nicarao in addition had cane arrows with stone and fishbone points and wooden swords with stone teeth. Shields of wood covered with woven cords and of danta (tapir) hide were common in northwestern Panama and eastern and southeastern Costa Rica where women accompanied their men to war. In Chontales, Nicaragua, there were hide shields; in Honduras, woven cane and hide; and the Nicarao used bark and wooden shields covered with cotton and feathers (Fernández de Oviedo, 1851-55, 3:104-05, 111, 12730; 4:37, 39, 53; Mártir de Anglería, 1944, p. 229; Herrera, 1726, dec. 4, lib. 8, p. 154; Fernández, 1881-1907, 2:131; 5:157, 374, 410; Fernández Guardia, 1908, pp. 34, 47, 50; Serrano y Sanz, 1908, pp. 352, 359; Vázquez, 1937-44, 4:121-22; Cockburn, 1779, p. 14; Peralta, 1898, pp. 59, 84). Religion The fundamental religious concept reaching from Cueva into northeastern Honduras was that of a universal unapproachable god, with minor gods to do his bidding. He was so distant that he was beyond the need of sacrifice or homage. The lesser deities, usually celestial bodies, animals, and reptiles, received the homage. Animism was a general pattern with spirits associated with mountains, pools, rocks, and woodland habitats (cf. Serrano y Sanz, 1908, p. 125; Fernández, 1881-1907, 5:372-73; Exquemelin, 1893, p. 251; Herrera, 1726, dec. 4, lib. 8, p. 157). The Chorotega-Mangue shared to an extent the animism of their neighbors. They gave the volcano of Masaya special reverence. They also had a festival celebrating the maize harvest, and in this fleeing women were pursued by husband and relatives. The woman who fled farthest was most appreciated (Fernández de Oviedo, 1851-55, 5:98). 57 Likewise, the adoption of eastern customs into the accepted ritual of the Nicarao 230

is apparent in the execution of the stick dance by itinerant and paid performers at the death festival of the chief of Agateyte (ibid., 4:94-95). This was a Guaymi dance to celebrate the pejibaye harvest, and was described in 1640 as "one of the most famous games which the Indians of this province (Guaymi) have . . . " (Serrano y Sanz, 1908, p. 94). 58 The Nicarao harvest festival for cacao emphasized the "volador" and a "see-saw" game called "comelagatoazte" of northern or Mexican origin (Fernández de Oviedo, 1851-55, 4:93-95). Lower Central American ceremonies were principally consultations by priest or priestess in ceremonial enclosures, caves, or designated sites (de Vilaplana, 1763, p. 47). These concerned war or the elements, such as earthquakes or hurricanes. Most idols were of wood, clay, or stone, although in Cueva and eastern Costa Rica there were prayer houses of perishable material with gold images (see Fernández, 1881-1907, 2:152). The Nicarao appear to have been in the process of adopting cast gold idols (Fernández de Oviedo, 1851-55, 3:103; 4:47, 103, 111; López de Gomara, 1941, 2:210). (Note also the position of goldsmiths; see Social Organization.) Mounds formed part of the religious constructions of the Nicarao, Chorotega, Maribio, and Lenca (Herrera, 1726, dec. 4, lib. 8, p. 158; Fernández de Oviedo, 1851-55, 4:43). Sections of the eastern periphery, however, must have differed. These were the regions where "the dances, rites, and religion" were said to be like Española (Haiti) and Cuba (López de Gómara, 1941, 2:202; Serrano y Sanz, 1908, p. 69) (see Language; Beliefs). Human sacrifice took place periodically from Panama through the Taguaca area in Honduras. The usual methods consisted in cutting off the head or slitting the throat (Fernández de Oviedo, 1851-55, 3:127; de 57 This is similar to the yearly fleeing of the Guatuso women (Stone, 1956a), a custom which may have been derived from the Corobici. 58 It is still a favorite of the Guaymi.

ETHNOHISTORY: LOWER CENTRAL AMERICA

Vilaplana, 1763, p. 116; Fernández, 18811907, 5:156; Fernández Guardia, 1908, pp. 35, 38; Serrano y Sanz, 1908, p. 372). The Nicarao, Chorotega-Mangue, and Maribio extracted the heart, and the Nicarao saved the heads on trees in front of temples (Fernández de Oviedo, 1851-55, 4:40-46; López de Gómara, 1941, 2:220). This last practice might have developed from the domestic ritual of assuring good hunting by saving animal skulls from the chase on poles by the dwelling (Fernández de Oviedo, 185155, 4:93). 59 The placing of human skulls of sacrificed victims near the temple is, of course, a Mesoamerican trait. The Maribio practiced the Xipe cult and used the flayed skins of their victims (ibid., 4:101). Ceremonial cannibalism seems confined to the upper-ranking Nicarao and Chorotega and to Sumu-Taguacas and Ulvas (ibid., 4:14, 37, 46-47, 51-52; Serrano y Sanz, 1908, p. 372; Exquemelin, 1893, p. 114; M. W., 1752, p. 305). The Guaymi, Paya, Lenca, Nicarao, and Chorotega sacrificed blood from the tongue, ears, or genitals (Serrano y Sanz, 1908, p. 361; Fernández de Oviedo, 1851-55, 4:47, 98; Squier, 1860, pp. 113-15). Baptism was practiced in Darien (Serrano y Sanz, 1908, pp. 130-31. 60 The Nicarao alone had confession, heard by an unmarried layman designated for one year who also pronounced penance. They practiced such domestic rituals as the making of offerings to household gods (Fernández de Oviedo, 1851-55, 4:55-56). Priest-Doctors and

Witch-Doctors

There were two kinds of paid male and female priests who were not necessarily celibates: a supreme priest-doctor to perform rituals, administer to idols or sacred spots, talk to spirits, serve as an oracle, and cure with herbs or by bleeding and sucking; and a witch-doctor associated with black magic who also cured but was shrouded in evil and considered to have the power to convert himself into animals and

to kill or bring sickness through remote control (Serrano y Sanz, 1908, pp. 88-91, 361; Fernández de Oviedo, 1851-55, 3:12627, 138, 160; Herrera, 1726, dec. 4, lib. 8, p. 159; Fernández, 1881-1907, 5:156-57; Las Casas, n.d., vol. 3, ch. 245, pp. 589-92; de Vilaplana, 1763, pp. 47, 101; Wafer, 1934, pp. 18, 24-26). The priest-doctor had his own paraphernalia, which generally included stones, feathers, leaves, and tobacco smoking. This Central American complex of priestdoctors and witch-doctors persisted among the Nicarao and the Chorotega. Both peoples had male and female witch-doctors who practiced black magic and were believed to have power to convert themselves into animals. Certain old women were credited with curing by blowing a concoction through a cane on the patient. Females, however, were barred from temples. Celibate priest-nobles who owned no property but performed rituals held an important rank in the Nicarao social hierarchy (Fernández de Oviedo, 1851-55, 4:37, 55, 107; López de Gómara, 1941, 2:219). Beliefs Belief in the creation of man from seed was current from the Guaymi northward through Costa Rica (Serrano y Sanz, 1908, p. 88; Fernández, 1881-1907, 5:372). 61 The Nicarao and Lenca believed in procreation by a god and goddess (Herrera, 1726, dec. 4, lib. 8, p. 157; Fernández de Oviedo, 185155, 4:41-42). 62 The idea of a nagual or 59 The custom of saving animal skulls and stringing them on the walls of houses for such a purpose is typical of eastern Central America today. 60 The Talamanca tribes of Costa Rica have their own baptism. See Stone, 1962. 61 Seed might have been a tuber. See Stone, 1956b. 62 It has been noted that parts of Panama had the "rites and religion" of Cuba and Haiti. In the latter, man was allegedly created from a human form by a bird (Mártir de Anglería, 1944, pp. 9 7 - 9 8 ) . This belief is evident in the archaeology of the Caribbean watershed of Costa Rica

231

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

guardian spirit was prevalent among the Lenca, Maribio, and probably the Nicarao (Herrera, 1726, dec. 4, lib. 8, p. 157; Fernández de Oviedo, 1851-55, 1:255-56; 4:107). 63 ARTS

Songs, Dances, and Musical

Instruments

Song was used for work, love, and recounting of histories and genealogies. Among the Cueva, even orders were given through song (Fernández de Oviedo, 185155, 3:137). Singing was usually in chorus at death, harvest festivals, divination, homage to the gods, marriages, on the return of victorious warriors, and at work reunions where group service was required. Dances were ceremonial and secular. (Balser, 1955, pp. 3 8 4 - 8 7 ) . For more recent documentation on the procreation of man in eastern Central America see Lehmann, 1910, pp. 717-18; Conzemius, 1932, p. 130. 63 See also Lehmann, 1920, 2:910. There is a similarity in the folklore of much of the Caribbean area of eastern Central America. In the period with which we are dealing there is documentation concerning the creation of the sea from a tree and that of the association of the procuring of water by crabs in Darien (Serrano y Sanz, 1908, pp. 126-27). Similar myths are known in Talamanca, Costa Rica (Stone, 1962). Animal owners or masters, a water jaguar responsible for drowning, the concept of Thunder living as a man on earth are some of the mythological themes common to the Talamancas and Borucas of Costa Rica (the Borucas are a mixture of aboriginal tribes from the Pacific side) and the Sumu and Misquito of Nicaragua and Honduras (Stone, 1949, 1962; Conzemius, 1932, pp. 127, 165, 169). Conzemius notes (ibid., p. 16) the identical belief regarding celestial bodies and their mode of travel in canoes among the Sumu and the Cuna. The Paya, Sumu, Misquito, and Rama have tales referring to an apelike creature which has been noted in South America (ibid., 168). The Sumu, Misquito, and Ulva, on the other hand, have myths resembling Mexican Quetzalcoatl, Ometecutli, and Omeciuatl pointing to a strong northern influence (Lehmann, 1920, 1:462). 64 Archaeology and ethnology were two types in Panama and had a wooden tongue, rested on was beaten with a stick similar drum. The other had hide and while carried. See Stone, 1962.

232

show that there Costa Rica. One the ground, and to the Antillean could be played

They were performed by large groups in circles or, less frequently, in a straight line. Dancers interlocked arms, danced hand in hand, or free and acrobatic. Both sexes or a single sex participated. The time was kept by foot movement and sometimes by jumps accompanied by song and often drums. A kind of clown or solo performer was common. Heavy drinking usually formed part of death, victory, or marriage celebrations but not at work reunions. The Quepos of Pacific Costa Rica were famous for never getting drunk (ibid., 1:128-30; 3:137, 142; 4:48, 70, 96-98; Serrano y Sanz, 1908, pp. 133-34, 361; Herrera, 1726, dec. 4, lib. 8, pp. 155-56; Mártir de Anglería, 1944, pp. 609-10; Las Casas, n.d., vol. 3, ch. 243, p. 587; Fernández de Navarrete, 1945, 3:392; Fernández Guardia, 1908, p. 46). Drums were used throughout the area.64 In Panama and Pacific Costa Rica there were gold trumpets and bells. Pacific Nicaragua also had trumpets (Fernández de Oviedo, 1851-55, 3:103, 181). Reed and bone flutes, shell horns, clay ocarinas, bone and shell bracelets and anklets which jingled, and rattles were common to the Caribbean area (Las Casas, n.d., vol. 3, ch. 243, p, 582; Mártir de Anglería, 1944, p. 611). GAMES

The only reports of games refer to the province of Darien where the rubber ball game and chance games were known to Panamanian Guaymi and their stick dance game, and to the two Mexican games already mentioned for the Nicarao (see Religion ). Sports such as swimming and shooting were common throughout (López de Gomara, 1941, 1:163; 2:202; Fernández de Oviedo, 1851-55, 3:159, 160; 4:93-95; Serrano y Sanz, 1908, pp. 94-95). BOOKS

Books of deer hide with painted figures appeared only in Nicaragua. They contained records of memorable events, rites, and land

ETHNOHISTORY: LOWER CENTRAL AMERICA

boundaries and were used by the "guegue" council to decide arguments (Fernández de Oviedo, 1851-55, 4:36; López de Gómara, 1941, 2:219-20, Mártir de Anglería, 1944, dec. 3, ch. 10). 65 MEASUREMENTS OF T I M E

The moon calendar and two major periods, "dry" and "wet," were characteristic of 65 Here we consider them Nicarao possessions because of the word "guegue." Their presence in the province of Nagrando, however, raises the question whether or not both the Chorotega Mangue and the Nicarao had books. It is also important to consider Peter Martyr's statement of the Indian questioned by the Spaniard in Darien. He may have been Mangue (see Language).

much of this territory (Wafer, 1934, p. 105; Fernández, 1881-1907, 5:156; Serrano y Sanz, 1908, p. 93). The Lenca, Nicarao, and probably the Chorotega divided their year into months of 20-day periods. The Nicarao and Lenca observed the first day of each month as a date of religious and other festivals. The Nicarao included the symbol of Quetzalcoatl, Acat, in their calendrical hierarchy of gods. Only three special days are mentioned for the Chorotega (Fernández de Oviedo, 1851-55, 4:52; Herrera, 1726, dec. 4, lib. 8, p. 160). The Sumu and Misquito had a timecounting device suggestive of the Peruvian quipu. This consisted of a string with knots (O. W. Roberts, 1827, p. 270).

REFERENCES Alvarez, 1944 Balser, 1955 Bancroft, 1882 Benzoni, 1857 Candolle, 1959 Cockburn, 1779 Collinson, 1870 Colón, 1947 Conzemius, 1928, 1932 Costa Rica-Panama, 1913 Dampier, 1699 Espinosa, C , 1519 Espinosa, I. F., 1742 Exquemelin, 1686, 1893 Fernández, 1881-1907 Fernández de Navarrete, 1945 Fernández de Oviedo y Valdés, 1851-55 Fernández Guardia, 1908 García Peláez, 1851-52 Habel, 1878 Herrera, 1726-30 Informe de los Padres, 1697 Jijón y Caamaño, 1938 Las Casas, n.d. Lehmann, W., 1910, 1920

López de Gómara, 1941 Lothrop, 1942 MacNutt, 1908 Mártir de Anglería, 1944 Matienzo, 1910 National Archives of Seville, 1600 Peralta, 1883, 1890, 1898 Pim and Seemann, 1869 Pittier, 1895 Ponee, 1873 Raveneau, 1869 Relación Breve, see Ponce, 1873 Relación . . . Espinosa, 1892 Roberts, O. W., 1827 Sapper, 1902, 1904 Serrano y Sanz, 1908 Squier, 1860 Staden, 1928 Stone, 1941, 1949, 1956a, 1956b, 1957, 1958, 1962 Strangeways, 1822 Torquemada, 1943 Vázquez, 1937-44 Vilaplana, 1763 W., M., 1752 Wafer, 1934

233

11. Mesoamerica and the Eastern Caribbean Area

IRVING

from Yucatan in MesoTamerica to Cuba in the West Indies HE PASSAGE

is only 120 miles wide. From it, the Indies stretch eastward past the southern tip of Florida to form the northern side of the Caribbean Basin, and then curve southward along the eastern side of the Caribbean to Venezuela and the Guianas in South America. Thus, they provide a series of convenient stepping stones whereby Indians might have traveled from Mesoamerica to the southeastern part of the United States and to the northeastern part of South America. The prevailing winds and currents, however, favor movement in the opposite direction, up from South America (Ricketson, 1940). The West Indies were inhabited in the time of Columbus by three major groups of Indians, the Ciboney, the Island Arawak, and the Island Carib. (Authorities differ as to the number and names of these groups. For a synonymy of the Ciboney, see Hahn, 1961, Table 1; and of the Arawak and Carib, Rouse, 1948, pp. 521-22, 549.) All three groups had counterparts in northeastern South America—the Warrau, the mainland Arawak, and the mainland Carib, re234

ROUSE

spectively—and it is probable that they all moved up from South America, although the Ciboney may have had a multiple origin, in Middle and North America as well as in the south (Rouse, 1960a). The Ciboney were the first to reach the Antilles. A radiocarbon date of 3770 B.C. for related people on the north coast of Venezuela (Rouse and Cruxent, 1963, p. 155) suggests a time when they might have moved out into the Antilles. They were a fishing folk, who lacked agriculture and pottery, and seem to have had a relatively simple form of social organization and religion. Their remains have been found in Trinidad and the Virgin Islands, at either end of the Lesser Antilles, and on the islands of Puerto Rico, Hispaniola, and Cuba in the Greater Antilles. The subsequent groups of Indians pushed them back into peripheral regions, adjacent to Mesoamerica, and by the time of Columbus they were limited to the southwestern peninsula of Haiti and the southern and western parts of Cuba. By radiocarbon dates (ibid.; Rouse, 1964), the Arawak must have started moving out into the Antilles from the north coast

MESOAMERICA & EASTERN CARIBBEAN

of Venezuela shortly before the time of Christ. By A.D. 1000 they had completed the movement, pushing the Ciboney back into the peripheral position they occupied in the time of Columbus. The Arawak were agriculturalists, relying primarily on South American root crops, especially bitter manioc, rather than on the characteristic Mesoamerican crops, such as maize (Fewkes, 1922, pp. 57-58); and they made pottery, which was likewise South American in type (Lovén, 1935, pp. 333-34). During contact time in the Greater Antilles, they had hereditary chieftains, a series of social classes, and a relatively elaborate religion, with priests and deities known as zemis (Rouse, 1948, pp. 521-46). The presence of these traits has led Steward and Faron (1959, pp. 239-51) to classify the island Arawak with a number of other tribes in Venezuela as "theocratic chiefdoms." The Carib were the last of the three groups to arrive in the Antilles. According to their traditions, they invaded the small islands of the Lesser Antilles, exclusive of Trinidad at the southern end, only shortly before the time of Columbus, killed off the men among the Arawak who then occupied the islands, and married the women. Nevertheless, it was the women's language which survived, for the Carib now speak an Arawakan language rather than Cariban (D. M. Taylor, 1951, p. 41). Like the Arawak, the Carib practiced agriculture as well as fishing and made pottery, but they were more democratic and less aristocratic, having only elected chiefs, and they emphasized warfare rather than religion. They were accustomed to raid the more peaceful Arawak for captives to eat; our term "cannibal" is derived from their name (Rouse, 1948, pp. 547-64). They are compared by Steward and Faron (1959, pp. 322-25) with the "tropical-forest villagers" of South America. Given origin of almost all the aborigines of the West Indies in South America, one is led to inquire whether they received any influences from the southeastern United

States via Florida or from Mesoamerica via Yucatan. Sturtevant (1960) has recently surveyed the evidence for the former and has come to an essentially negative conclusion. The problem of contact between Mesoamerica and the West Indies is more difficult to resolve. Three previous authors have discussed this problem in detail: Lovén (1924, 1935), Berlin (1940), and Morales Patino (1949, 1950). All agree that both the Mesoamericans and the West Indians possessed the technical capacity to travel between the two regions. But documentary evidences of travel are rare, and artifacts which might have been carried from one region to the other are difficult to identify. The three authors differ as to how much contact can be inferred from the cultural similarities between the two regions. Lovén concludes that contact was considerable, since a number of Mesoamerican and Antillean traits are so much alike that they must have diffused from the mainland to the islands (though not vice versa). Berlin and Morales Patiño argue instead that contact was probably limited to an occasional canoe party blown to the islands by storms. They consider most of the resemblances cited by Lovén to be fortuitous. Perhaps the best way to summarize the situation is to say that the evidences for Meso- and Central American influences in the West Indies are suggestive but not conclusive. One probable case of diffusion in the opposite direction will also be discussed below. DOCUMENTARY EVIDENCES OF CONTACT

The previous authors cite, as evidence of Mesoamerican ability to travel to the Antilles, the large trading canoe encountered by Columbus in the Bay Islands of Spanish Honduras during his fourth voyage in 1502. Las Casas (1951, 2:274) speculated that this canoe was on a voyage from Yucatan, but Morison (1942, 2:331-33) argues, less convincingly, that it was engaged in local 235

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

trade between the Bay Islands and the adjacent mainland of Spanish Honduras. Be that as it may, the canoe is impressive: it "must have held about 40 persons, for to the 25 men must be added the women and children. The cargo, too, must have been fairly bulky to justify the running costs of a voyage with that number of merchants and crewmen" (J. E. S. Thompson, 1951, p. 69). Comparable canoes and trading voyages are known from the Antilles. According to Oviedo (1959a, 67:149), the Arawak of Hispaniola had canoes which could carry 40-50 men and which were supplied with both oars and sails. Columbus (1930, p. 124) encountered an Arawak canoe in Jamaica which was 96 feet long and 8 feet wide; presumably it was built of planks, since no tree trunk would have been that large. Such large canoes "were maintained by the caciques for trade and prestige" (McKusick, 1960, p. 7). The Carib had similar canoes, though without sails, which they used to raid northeastern South America and the Greater Antilles (ibid., pp. 4-7). The nature of Ciboney canoes is not known. Despite the quality of this equipment— and the shortness of the passage between Yucatan and Cuba—we possess only a single record of pre-Columbian travel between the two. The Book of Chilam Balam of Chumayel reports the arrival in Yucatan during the year A.D. 1359 of unclothed strangers looking for people to eat (Roys, 1933, p. 142; Berlin, 1940, p. 159). Presumably, these were Carib raiders, who had extended their normal, West Indian range of operation to Yucatan. There is, in addition, some evidence that the Arawak knew of Meso- or Central America, though Berlin (1940, pp. 156-58) considers it unreliable. The inhabitants of northern Hispaniola informed Columbus during his first voyage that Jamaica and Hispaniola were only ten days by canoe from the mainland where, unlike themselves, the natives wore clothes (Navarrete, 236

1825-37, p. 127). Indians encountered in central Cuba during Columbus' second voyage similarly talked of a man clothed in a white tunic, which extended to his feet (Las Casas, 1951, 1:388); other contemporary sources elaborate this reference into a "race that wore clothes, under a mighty prince, who was also clothed" (Lovén, 1935, p. 59). The linguistic evidence for pre-Columbian contact between Mesoamerica and the West Indies is less satisfactory. Wolf (1959, p. 13) states that the West Indians "answered his [Columbus'] interminable questions about lands to the east by repeating 'Colhua, Colhua!'," which is an Aztec name; however, I have been unable to verify this reference. Other authors cite the presence in Mesoamerica of such Arawakan terms as cacique, huracán, and hamaca, and in Cuba of the name Maya. Morales Patiño (1950, pp. 86, 96) argues that the Spaniards introduced these terms during colonial time, and he is probably right in most cases. Huracán does have a prehistoric counterpart in Hurakán among the Quiche of Guatemala (Ortiz, 1947, p. 467), but, since the former word refers to a storm and the latter to a creator deity, the resemblance is probably fortuitous. Hamaca is the term most likely to have passed from the Antilles to Mesoamerica before the arrival of Columbus; the evidence for this will be discussed below. Finally, the documents record one instance of accidental contact between Mesoamerica and the West Indies, resulting from storms. In 1518, during the discovery and conquest of Yucatan, Juan de Grijalva encountered a Jamaican Indian girl on the island of Cozumel, off the coast. This girl said that "she had sailed from Jamaica two years before in a large canoe with ten of her countrymen, to fish at certain small islands, and that the current had driven them hither, where the natives had killed and sacrificed her husband and all her companions . . ." (Díaz del Castillo, 1927, 1:40-

MESOAMERICA & EASTERN CARIBBEAN

41). Berlin (1940, p. 156) cites a second instance, in which Spaniards were blown from Jamaica to the mainland, but this is a misquotation; the Spaniards were on their way from Darien to Santo Domingo when they were blown to Yucatan (Díaz del Castillo, 1927, 1:265). OBJECTS OF FOREIGN ORIGIN

Columbus, on his first voyage, and Las Casas, a short time after the conquest of Cuba, observed honeycombs on the island of Cuba, the former in Arawak territory and the latter among the Ciboney. Las Casas (1951, 1:245-46) concluded that these honeycombs must have originated in Yucatan, since bees were domesticated there but not in the Antilles, and suggested two alternative ways in which they might have reached Cuba: (1) local Indians might have voyaged to the mainland for them, or (2) Yucatan traders might have accidentally been blown by storms to Cuba. Lovén (1935, pp. 59-60) accepts the former hypothesis, Berlin (1940, pp. 156-57) prefers the latter, while Morales Patiño (1950, p. 95) suggests alternatively that the honeycombs may have floated with the currents from Central America! Archaeological evidences of trade or other contact are rare. Maurice Ries (personal communication) reports finding a Classic Maya potsherd and several pieces of obsidian on the beach at Cabo San Antonio, Cuba, across the channel from Yucatan. I visited this beach in 1941 in an effort to follow up Ries' report, but found only Ciboney remains. Covarrubias (1957, fig. 33, pp. 70-71) illustrates a "jadeite celt with incised design, purchased from a Cuban Negro witchdoctor," which is now in the Museo Nacional de Antropología in Mexico. This is Olmec in style, and so it must date from the first millennium B.C. Whether it reached Cuba at the time it was made or after the arrival of the Spaniards is uncertain; if the former, it would have been too early to have

influenced the development of the cult of the celt which is discussed below. Five stone metates of Central American type have been found in the Greater Antilles, one in Cuba (Morales Patiño, 1950, p. 77), three in Jamaica (Lovén, 1935, pp. 379-80), and one in Puerto Rico (Richardson, 1942, p. 96). The exact nature of the Cuban metate is not known. Deurden (1897, figs. 5, 6) has illustrated two of the Jamaican specimens. From these illustrations, M. D. Coe (personal communication) concludes that the specimens were probably made in northwestern Costa Rica or adjacent regions during the Early Polychrome Β period, dating from A.D. 500 to 750. Richardson (1942, p. 96) similarly compares the Puerto Rican metate with one from Honduras, the area and age of which are unknown. Previous commentators (e.g., Morales Patiño, 1950, p. 81) have assumed that the five metates were brought to the Antilles by Indian slaves whom the Spaniards imported from the mainland during colonial time. According to Coe (personal communication), this is not necessarily inconsistent with the early date of the two Costa Rican specimens, since similar artifacts are still being dug up and re-used by the inhabitants of that country. If, on the other hand, the two specimens reached Jamaica during the period of their manufacture, they predated the Arawak occupation of the island, which apparently took place after A.D. 700 (Rouse, 1964, fig. 5 ) . Another object which almost certainly reached the Antilles from Central America, in either prehistoric or historic time, is a sculptured pumice figure from Ouanaminthe, Haiti, which is now at the Museum of the American Indian in New York. According to Richardson (1942, p. 96), this human figure "finds its only parallel in Nicaragua . . . . So identical in technique and material is [it] to those from Nicaragua that one is forced to consider it an object of trade." 237

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

CULTURAL RESEMBLANCES: CIBONEY

Of the three major groups of West Indians, the Carib is the only one for which no cultural resemblances with Mesoamerica have been noted. Berlin (1940, p. 145) likewise denies the existence of resemblances between the Ciboney and Mesoamerica, but his denial has been outdated by subsequent research. Actually, the Ciboney had the greatest opportunity for contact, both because they were on the islands for the longest period of time and because they were the sole occupants of western Cuba, adjacent to Yucatan. On the other hand, they would have had the greatest difficulty in accepting influences since, as Berlin (ibid.) points out, they were furthest below the Mesoamerican level of cultural development. Hahn (1961) has recently been able to set up a chronology for the Ciboney cultures of Cuba consisting of five periods, numbered from I to V. The first and the last of these are hypothetical. The culture of Period II, dating perhaps from A.D. 1 to 350, bears no resemblance to Mesoamerica. The similarities begin during Period III, i.e., between A.D. 350 and 1100, with the appearance of stone balls and dagger-shaped ceremonial stones known locally as dagolitos. Others are added during Period IV, from A.D. 1100 to 1500, including stone rings and spoons. These objects are all relatively simple and therefore could have been invented independently in Cuba, especially since they form part of a more extensive development of ground stonework during Periods III and IV. On the other hand, it may be significant that the resemblances are multiple, even though they do not seem to have been functionally interrelated. Haiti is further from Mesoamerica than Cuba, but it contains a more complex and therefore possibly a more significant resemblance. Both of its Ciboney complexes, Couri in the north and Cabaret in the south, 238

are characterized by large flint blades, rechipped only on the edges, if at all. Some of these are tanged, and may have served as spearheads or daggers; others appear to have been knives; and still others, scrapers. W. R. Coe (1957) has noted the widespread occurrence of similar artifacts in Meso- and Central America, especially in British Honduras, and Epstein (personal communication) has commented that both the Haitian and Mesoamerican blades tend to have plain striking platforms, whereas South American platforms are faceted. If the Ciboney blades of Haiti are derived from Mesoamerica, it is surprising that none has been found on the intervening islands of Jamaica and Cuba, except for a few simpler blades on the latter island (Hahn, 1961). I have suggested (1960a, p. 24) the possibility that a group of preceramic Mesoamericans may accidentally have been blown from the mainland directly to Haiti, bypassing the other two islands. CULTURAL RESEMBLANCES: ARAWAK

The Arawak arrived in the West Indies during Period II of the Ciboney chronology discussed above. By the end of that period, they had traveled northward as far as Puerto Rico, the easternmost island of the Greater Antilles. It was not until Period III, however, that they penetrated the westernmost islands, where contact with Mesoamerica became possible (Cruxent and Rouse, 1958-59, fig. 4). Period III also happens to be the time when resemblances with Mesoamerica make their first appearance. Before that time, the Arawak seem to have had a relatively simple culture, like the tropical-forest villagers of South America, but during the latter part of Period III they began to construct ball courts and to make the idols (zemis) which they worshipped in the time of Columbus, and presumably they also developed hereditary chieftains and social classes. These developments became more

MESOAMERICA & EASTERN CARIBBEAN

elaborate during Period IV and reached their peak as the Spaniards arrived (Rouse, 1953). It must have been largely coincidence that the developments took place just as the Arawak reached the position to make contact with Mesoamerica. The resemblances to the latter area are relatively superficial. Moreover, had they been inspired from Mesoamerica, one would expect them to occur both among the Arawak and the Ciboney, who lived nearer to Mesoamerica, and they do not; except for stone balls, the Ciboney similarities are different. Nevertheless, there are enough specific, relatively complex resemblances between Arawakan and Mesoamerican culture to raise the question whether the latter exerted influence on the former. Maize As already noted, bitter manioc was the main crop of the Arawak, in contrast to maize in Mesoamerica. Nevertheless, maize also occurs among the Arawak, and it is reasonable to inquire whether it spread directly from Mesoamerica (Gower, 1927, p. 46). There are two reasons for assuming that it did not. First, Brown (1960) has recently made a classification of modern Greater Antillean maize into seven races. He concludes that only three of these go back to prehistoric time and that all three are native to South America rather than to Mesoamerica. Second, the aboriginal method of preparing maize in the islands was different from that in Mesoamerica. If we are to believe Oviedo (1959b, p. 15), the Arawak roasted the ears in the fire, instead of preparing corn in the Mesoamerican fashion by soaking it overnight in lime water and grinding the swollen kernels into dough for making tamales and tortillas. However, the historic Indians did grind corn meal (Lovén, 1935, p. 376) and the prevalence of pestles, though not metates, in the sites

of Period IV suggests the possibility that corn meal may go back to prehistoric time. Hammock In the time of Columbus, the Arawak normally slept in hammocks and the Maya in beds (Tozzer, 1941, pp. 86-87). Nevertheless, there is evidence, both for prehistoric and early historic time, that the Maya used the hammock secondarily. A painting on the Late Classic "Tabasco vase," which is probably from Jaina Island, shows a chief sitting on a jaguar-skin hammock (Cook de Leonard, 1954, pp. 96-102; Covarrubias, 1957, plate facing p. 228). Martyr (1912, 2:27) indicates that Cortez found the hammock in use on Cozumel Island during the conquest in 1519 and states that the local natives called it by the same name as in the Antilles, ammaca. Oviedo (1959a, 3:403) likewise describes its use by a Maya chief whom Montejo met in northern Yucatan in 1528. On the other hand, Landa describes only beds for the period when he was in Yucatan during the latter part of the 16th century, though the hammock is mentioned in several lesser sources of the time (Tozzer, 1941, pp. 86-87). Subsequently, its use increased, until it has become universal among the natives of Yucatan (ibid.).1 The authorities disagree as to when and how the hammock reached Yucatan. J. E. S. Thompson (1930, p. 94) concludes that it was prehistoric, but prefers to derive it from lower Central America rather than from the West Indies because of certain modern resemblances with the former area. Ε. Η. Thompson considered it a local development in Yucatan (Tozzer, 1907, p. 56). Molina (1896, pp. 17, 246-47) argued that the Spaniards introduced it from the West Indies during the 15th century. I prefer an hypothesis of direct, prehistoric diffusion from the Indies, because of the evidence for 1

I am indebted to Michael D. Coe for calling to my attention several resemblances on the Mesoamerican side, especially that of the hammock.

239

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

pre-Columbian contact cited above and the fact that the three places where the hammock is reported in preconquest and conquest time happen to be precisely where the contact took place—on Cozumel Island and in northern Yucatan. Chiefs'

Insignia

Lovén (1935, pp. 467, 484, 507-14, 656, 677-78, 691) believes that certain insignia of the chiefs and "nobles" of Haiti and/or Jamaica are derived from Mesoamerica. These include standards, presumably of cotton cloth; feather crowns and hoods; diadems composed of wooden (?) wings inlaid with gold, or of stone and gold beads; similarly beaded girdles; and aprons woven of cotton cloth with stone beads embroidered on them. Masklike ornaments, likewise apparently made of wood and inlaid with gold, were worn on the girdles in Haiti as among the Classic Maya. None of the other writers comments on these resemblances. Ball Game The relation of the Mesoamerican rubberball game to the one played by Arawak has been the subject of controversy. Lovén (1935, p. 680), Rouse (1953, p. 197; 1954), and Ekholm (1961, p. 18) conclude that the ball game diffused directly from Yucatan to the Greater Antilles. On the other hand, a series of authors, of whom T. Stern (1948, p. 100; 1954) and Alegría (1951, p. 348) are the latest, assume a contrary spread from South America into the Antilles. (Some of the latter authors would derive the South American form of the ball court from Mesoamerica via Panama and others assume diffusion in the opposite direction, but this is irrelevant to our argument.) The argument depends on two points: the degree of similarity between the ball game in Mesoamerica, on the islands, and in South America; and the distribution of games in each place. The advocates of direct, Mesoamerican-Antillean connections 240

note that the Arawak courts are bordered with earth and/or slab walls, which also occur in Mesoamerica but not in South America (Lovén, 1935, p. 60); and are accompanied by stone collars, which have a similar distribution. Ekholm (1946a, 1961) and Alegría (1951, p. 351) interpret the collars as effigies of the wooden belts worn by the ball players, as was true of the comparable stone yokes in Mesoamerica. On the other hand, Morales Patiño (1950, pp. 8889) notes that the games themselves were different; he compares the Maya game to our basketball and the Arawak game to our volley ball. T. Stern (1954) adds that the Arawak game is most like that played by the Otomac Indians in the middle of the Orinoco Valley, Venezuela. The fully developed Arawak game, with slab-lined courts, petroglyphs on the slabs, collars, and other associated stone objects, such as carved pillars and balls, and threepointed and elbow stones, is limited to the Virgin Islands, Puerto Rico, and the Dominican Republic in the very center of the Antilles. Haiti and eastern Cuba contain a simpler form of the game, whereas it is completely absent from the rest of Cuba and from Jamaica, which are the parts of the Antilles closest to Mesoamerica. Berlin (1940, p. 153) terms this gap a "missing link" and argues that it precludes the possibility of direct connection between the Arawak and Mayan games; but he fails to note the existence of a far larger gap between the Antillean game and its closest occurrence in South America. The game is not known from the Lesser Antilles or from eastern Venezuela; the closest record we have of it is among the Otomac Indians, mentioned above, who are believed to have migrated eastward from Colombia at a relatively late date (Rouse, 1954). If one accepts the gaps as evidence against connection, therefore, one is forced to assume that Arawak invented the game independently in the Greater Antilles. I prefer to believe that the Arawak game is

MESOAMERICA & EASTERN CARIBBEAN

derived directly from Mesoamerica, if only as the result of shipwreck of a Maya canoe, such as has been discussed above. This would explain the gap in distribution in the western part of the Antilles, since the canoeists could have been driven by the winds past Cuba and Jamaica to Hispaniola; and it would also account for the relative simplicity of the Arawak game, since a shipwrecked group would presumably have been too small to impose the game in all of its original complexity. Religion Arawak religion has a distribution similar to that of the ball game; it was most highly developed in Puerto Rico and the Dominican Republic and shades off as one moves eastward into the Virgin Islands and the Lesser Antilles or westward towards Mesoamerica. The greatest resemblances to Mesoamerica are to be found in the central part of the area, as in the case of the ball game. Indeed, Arawak idols (zemis), e.g., the "three-pointed stones," are frequently found in ball courts, although there is no documentary evidence that ceremonies were conducted in connection with the game, as they were in Mesoamerica (Alegría, 1951, p. 352). Lovén (1935, pp. 613-14, 655-56, 68182, 693) concludes that "it is probably from Yucatan that the Tainos [Arawak of Puerto Rico and Hispaniola] learnt to make idols of wood or stone, as well as other religious stone sculpture." He is able to cite only one specific resemblance in deities, however, a wind god or rain god from Jamaica (Joyce, 1916, pp. 184-85). On the other hand, the Lothrops (1927) note the occurrence of plaster on a carved stone head from Puerto Rico, traces of paint on several other Puerto Rican carvings, and shell and gold inlays

throughout the Greater Antilles. All these traits are distinctively Mesoamerican, although inlaying also occurs in northern South America. Counterbalancing these resemblances is the fact that freestanding clay figurines, so common in Mesoamerica, occur only rarely in the West Indies (Lovén, 1935, p. 614). The Arawak did portray their deities (zemis) as adornos on pottery vessels, but these are more reminiscent of Barrancoid pottery in eastern Venezuela than of Mesoamerican figurines (Cruxent and Rouse, 1958-59, 1:26). The carving of deities on stone celts would link the Arawak more closely with Mesoamerica than with Venezuela, were it not for the fact, already noted, that this is an Olmec trait in Mesoamerica and therefore too early for the occurrence of the trait in the Antilles. Finally, Lovén (1935, pp. 656, 681, 692) would derive the Taino myth of an underworld with its own ruler from the Maya. Morales Patiño (1950, p. 90) argues, on the contrary, that Taino mythology is completely different from that in Mesoamerica. Miscellaneous

Traits

A number of other unrelated traits have been cited as possible evidences of contact between Mesoamerica and the Arawak. These include wooden gongs (Joyce, 1916, p. 204), clay stamps (Ries, 1932, pp. 44054), earplugs (Lovén, 1935, pp. 487-88), monolithic axes (ibid., pp. 160-61), and various weapons (Morales Patiño, 1950, pp. 78-80). In my opinion, these are all too general or too widely distributed to be significant. It is to Arawak social organization and religion, rather than to the more mundane aspects of their life, that we must look for Mesoamerican influences, if any.

241

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

REFERENCES Alegría, 1951 Berlin, 1940 Brown, 1960 Coe, W. R., 1957 Columbus, 1930 Cook de Leonard, 1954 Covarrubias, 1957 Cruxent and Rouse, 1958-59 Deurden, 1897 Diaz del Castillo, 1927 Ekholm, 1946a, 1961 Fewkes, 1922 Gower, 1927 Hahn, 1961 Joyce, 1916 Las Casas, 1951 Lothrop, R. W. and S. K , 1927 Lovén, 1924, 1935 McKusick, 1960 Martyr, 1912

242

Molina Solis, 1896 Morales Patiño, 1949, 1950 Morison, 1942 Navarrete, M. F. de, 1825-37 Ortiz, 1947 Oviedo y Valdés, 1959a, 1959b Richardson, 1942 Ricketson, 1940 Ries, 1932 Rouse, 1948, 1953, 1954, 1960a, 1964 and Cruxent, 1963 Roys, 1933 Stern, T., 1948, 1954 Steward and Faron, 1959 Sturtevant, 1960 Taylor, D. M., 1951 Thompson, J. E. S., 1930, 1951 Tozzer, 1907, 1941 Wolf, 1959

12. Mesoamerica and Ecuador

CLIFFORD EVANS and BETTY J. MEGGERS

between Mesoamerica T relationships and Ecuador dates back more than HE POSTULATION of direct influences or

half a century, beginning with the writings of Saville (1907, 1909, 1910) and Jijón y Caamaño (1914, 1930, 1951). Unfortunately, stratigraphic sequences were poorly known in both areas at that time, and the conclusions were of necessity vaguely oriented chronologically, making them easy to attack. If Uhle, who devoted considerable attention to this question (1922, 1923a, 1923b, 1927, 1931), had used a term with less specific cultural connotations than "Mayoid," recent stratigraphic work in both areas would have proven many of his points of relationships to be correct and he would have been spared the ridicule of many contemporary colleagues, some recent specialists in the area, and certain reviewers of his works. In recent years, there have been a few additional contributions to the problem of interrelationships between the two areas: Henri Lehmann (1951, pp. 291-98; 1953, pp. 78-80) has published various articles on figurines, especially those in beds or cradles; Brainerd (1953, pp. 14-17) wrote

on possible speech-scrolls on a cylindrical stamp from Ecuador; Nicholson (1953, pp. 164-66) discussed the problem of fine orange ware; Willey (1955a, pp. 35-42, 45; 1958, pp. 353-78) has generalized on the subject in summary articles; Evans and Meggers (1957, pp. 235-47) suggested strong relationships at the Formative period as a result of deep stratigraphic excavations made in 1956 in the Guayas Basin; Estrada (1957c, pp. 141-56; Estrada and Evans, 1963, pp. 80-81, 83-84) has pointed out similarities of various Ecuadorian traits to Mesoamerica; Borhegyi (1959, pp. 141-56) has published an inventory of similarities and called attention to such specific items as three-pronged incense burners (1960, pp. 157-64); M. D. Coe (1960b, pp. 387-417) and Estrada and Evans (1963, pp. 80-84) have outlined specific similarities between Ecuador and coastal Guatemala during the Formative period; and Meggers (1963, pp. 132-45) has discussed theoretical problems arising from the inference of connections between Ecuador and Mesoamerica at various times. A dependable evaluation of the strength 243

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

FIG. 1—OBSIDIAN ARTIFACTS OF EARLY HUNTING CULTURE, HIGHLAND ECUADOR. El Inga and Loson sites.

and significance of aboriginal connections between Mesoamerica and Ecuador cannot be made without detailed chronological information on the appearance of the traits in both regions, so that contemporaneity or relative antiquity can be established. The absence of such information, particularly in Ecuador, has handicapped investigators in the past. In the last few years, however, intensive archaeological investigations, particularly by Estrada (1956; 1957a,b,c; 1958; 1962), have gone far to remove this obstacle. With many of the Ecuadorian traits pinpointed chronologically, a new look at the situation is in order. In keeping with the improved orientation of the data, the pres244

entation will be in terms of broad stages of cultural development, beginning with the Pre-agricultural (Paleo-Indian, Hunting, Fishing, and Gathering) period, continuing through Early and Middle-Late Formative (Preclassic), Regional Developmental (Florescent, Classic) periods, and culminating in the Integration (Empire and Conquest, Postclassic) period. The relationships will be viewed from Ecuador; if a trait or complex does not exist in a clearly defined cultural or time position in Ecuador, no comment on its Mesoamerican affiliations will be made. Traits that are widely diffused in other parts of South America will not be included in this analysis.

MESOAMERICA & ECUADOR

PRE-AGRICULTURAL

( PALEO-INDIAN,

HUNT-

ING, FISHING, AND GATHERING) PERIOD

In spite of an increasing number of clearly defined early lithic complexes in Mexico associated with extinct Pleistocene fauna, nothing of comparable nature has shown up in Ecuador. The Punin calvarium (Sullivan and Hellman, 1925) lacks reliable geological provenience and was not associated with artifacts. On the Ecuadorian coast no early hunting, fishing, and gathering sites without pottery and comparable to the early Huaca Prieta or Pampa de los Fosiles sites in Peru have been excavated. The most promising indications of relationships between some of South America's earliest immigrants and early hunting-andgathering groups from North America are the sites of El Inga and Loson near Quito reported by Mayer-Oakes and Bell (1960, pp. 1805-06; Bell, 1960, pp. 102-06; MayerOakes, 1963, pp. 116-28). The El Inga deposit represents a camp and workshop site. Refuse, roughly 45 cm. deep, occupies a badly eroded hillslope flanking Ilalo Mountain near the town of Tumbaco in the Province of Pichincha. The Loson site is approximately 2 km. away in a similar geographical situation. The majority of the artifacts are obsidian, and include a variety of projectile points, side and end scrapers, ovate blades, gravers, burins, drills, prismatic blades, microblades, and small hemispheric polyhedral cores (fig. 1). Basalt artifacts are generally triangular, but a few well-made scrapers, choppers, flakes, and cores do occur. Several projectile points closely resemble Mesoamerican types. The double-pointed, laurel-leaf-shaped points are very similar to the Mexican Lerma points, found with the second Santa Isabel Iztapan mammoth (Aveleyra A. de Anda, 1956, fig. 7-2) and defined from the stratigraphic sequence of the Sierra de Tamaulipas, where the Lerma horizon has a carbon-14 date of 7312 B.C. ± 500 years

(MacNeish, 1958, pp. 52, 152-53, 194; sample M-499; Wormington, 1957, pp. 99, 202). Preliminary testing of a few El Inga specimens by the obsidian method of dating gives a wide range, but all the specimens have a hydration rim of 5.7-10.0 microns, giving a probable minimum antiquity of 6000-7000 years. There is no doubt that an early wave of migrants from North America into the South American continent passed through the highlands of Ecuador. EARLY

FORMATIVE

(INCIPIENT

AGRICUL-

TURAL) PERIOD

The earliest pottery-producing culture in Ecuador is Valdivia, represented by sites along the south Ecuadorian coast in Guayas Province at Buena Vista, Posorja, Punta Arenas de Posorja, Palmar, Valdivia (Estrada, 1956; Evans, Meggers, and Estrada, 1959), and at San Pablo (Zevallos and Holm, 1960). The main food resource was shellfish, supporting a fairly sedentary population occupying sites long enough to accumulate midden refuse several meters in depth. C14 dates based on three shell and five charcoal samples from the middle to lower part of the sequence give a range of dates of 2090 B.C. to 3190 B.C. (See U.S. Geological Survey samples W-630, W-631, W-632, and University of Michigan samples M-1317, M-1318, M-1320, M-1321, M-1322, for detailed discussion of provenience and cultural associations.) Stratigraphic excavations and an analysis of the change in pottery and stone artifact types throughout the sequence establish several subperiods, where new traits appear and others are replaced. The percussion-made stone tool complex and incised pottery tradition is more comparable to the Formative period cultures of Guañape on the north coast of Peru, Barlovento near Cartagena in Colombia, and Monagrillo on the Pacific coast of Parita Bay, Panama, than it is to Mesoamerican cultures. None of these cultures, however, show as close similarities with the 245

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

FIG. 2—IRIDESCENT PAINT, α-c, Chorrera phase, Ecuador, d-g, Ocos phase, Guatemala.

total ceramic complex as does early Valdivia and late-early to early-middle Jomon of western Japan. Period Β of Valdivia culture witnesses the appearance of small handmade pottery figurines and the elaboration of decorative techniques such as excision, complex motifs of broad-line incision, pebble polishing, brushing, appliqué fillets, and rocker-stamping (some of which may have been also present in Period A). Comparison of these traits with pottery of other Formative period cultures in Mesoamerica and South America shows the closest resemblance to Tlatilco in the Valley of Mexico. Although the pottery figurines are distinct from both sites, comparison shows a large number of unusual features in common (Table 1). Since the Tlatilco culture is more elaborate in all details than Valdivia and the C14 dates are about 1000 years younger, the unique features of Period Β cannot be derived from Tlatilco or sites of a similar complex in Mesoamerica. Perhaps the simi246

larities result from diffusion from coastal Ecuador both northward to Mesoamerica and southward along the Peruvian coast, where local modifications were made. At the same time, roughly 2000-1500 B.C., another shellfishing group appears on part

TABLE 1—COMPARISON O F VALDIVIA PERIOD Β AND TLATILCO FIGURINES Figurine Female Male Bisexual Standing Pregnant Seated Baby-in-arms Nude Clothed Elaborate hairdress Jewelry No jewelry Solid Handmade Two heads

Valdivia X Rare X X Rare Rare Rare X Rare X X X X X

Tlatilco X X ... X X X X X X X X X X X X

MESOAMERICA & ECUADOR

of the coast of Ecuador, known as the Machalilla culture (Meggers and Evans, 1962). Principally hunters, fishers, and gatherers, they have a pottery complex completely unlike Valdivia, with thin walls, highly polished surfaces, stirrup spouts and strongly carinated bowls. No similar complex has been reported in Mesoamerica or other parts of South America. These two cultures lived side by side, during Period C of Valdivia culture, trading back and forth some of their pottery apparently without grossly affecting their ways of life. About 1500 B.C., for reasons unexplained at present, the Valdivia culture becomes extinct; a series of new traits comes to the Machalilla culture, amalgamates with it, and produces a complex known as the Chorrera culture. LATE

FORMATIVE

(DEVELOPED

AGRICUL-

TURAL, PRECLASSIC) STAGE1

Stratigraphic excavations at La Victoria site, near the Mexican-Guatemalan border not far from La Victoria, Guatemala (M. D. Coe, 1960b), and the work of the New World Archaeological Foundation in the state of Chiapas, Mexico (Dixon, 1959; Lowe and Navarrete, 1959; C. Navarrete, 1960) produced ceramics that correlate closely with Late Formative period materials from the Chorrera phase of the Guayas Province of Ecuador (Evans and Meggers, 1957; Estrada, 1958). The absence of many of these traits south of Ecuador has suggested a general filtering out of cultural features during a movement southward out of Mesoamerica. However, the possibility of a two-way exchange, in which some traits traveled northward out of Ecuador to Mesoamerica, cannot be ruled out at our present state of knowledge. Specific items found in Ecuador in the Chorrera phase or the transitional ChorreraTejar phase can be compared with traits 1 Ed. note: "Late Formative" here would correspond chronologically to what Phillips (Art. 15) refers to as "Early Formative" in Peru, i.e. the Chavin horizon.

FIG. 3—ROCKER STAMPING. a, Chorrera phase, Ecuador. b-d, Ocos phase, Guatemala. e, Preclassic period (Chiapa I and I I ) , Chiapas, Mexico (after Dixon, 1959, fig. 5 5 ) . f-h, Playa de los Muertos horizon, Honduras.

from various parts of Mesoamerica, where the Preclassic period has been clearly defined. Iridescent, lustrous painting of the Chorrera complex, varying from a steelgray metallic appearance to a pinkish hue when applied too thickly, and executed in parallel stripes, often running in diagonal lines interspersed with dots, is identical in appearance, design, technique of application, and all general features with materials from the oldest phase, Ocos, at La Victoria, Guatemala (fig. 2). The principal difference lies in the greater thickness of the vessel walls of the Guatemala sherds; vessel shapes, surface treatment, irregular rims, and zoning of the iridescent paint by bordering it with incising are closely equivalent in the two areas. Although sherds from Phase I at Mirador site in the Municipality of Jiquipilas, northeast of Colonia Vincent Guerrero in western Chiapas, have been reported to have iridescent paint on the 247

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

rims of incurved, small-mouthed bowls (Navarrete, Peterson, McNeish, personal communications), these sherds do not have the typical stripes or dots and have not been personally inspected by us or by Coe, so that identification is not absolute. If the Mirador Phase I materials are the same, this would be evidence of more widespread occurrence of the technique in Mesoamerica on an early time horizon because Mirador I equates with Chiapa I at Chiapa de Corzo site, with Ocos phase at La Victoria site and with the first Preclassic horizon of Santa Marta Cave, Chiapas. No other Mesoamerican sites to date appear to have produced this material. The technique died out at the end of the Preclassic period on coastal Guatemala, but persisted on the Ecuadorian coast into later archaeological periods. Fingernail punctate or gouging is not common in either Mesoamerica or Ecuador, but occurs in both the Chorrera phase of Ecuador and the Ocos phase of Guatemala (M. D. Coe, 1960b, cf. figs. 2,a-d; 3,a-c) and in Chiapa I, or the Cotorra phase, of the Early Preclassic period of Chiapa de Corzo and Frailesca regions of Chiapas, Mexico (Dixon, 1959, fig. 52,b-d; C. Navarrete, 1960, p. 24, fig. 22,d,g). Rocker-stamping of the plain variety is present but not common in the Chorrera phase (fig. 3); in the Ocos phase plain rocker-stamping also occurs but the shell dentate variety is the more common type (fig. 3,b,c,d; M. D. Coe, 1960b, cf. figs. 2,e-g; 3,d). Of interest is the presence of rocker-stamping in other Mesoamerican Formative or Preclassic cultures, such as Chiapa I at Chiapa de Corzo (Dixon, 1959, pLs. 52,p-t; 55); Cotorra phase in the Frailesca subregion of Chiapas (C. Navarrete, 1960, p. 25, fig. 23); Tlatilco (M. N. Porter, 1953, p. 37-38, pls. 9,H; 11,F,G; Piña Chan, 1958, p. 92, pls. 6, 12, Table 5, Chart 2); the bichrome horizon of the Playa de los Muertos culture of Honduras (Strong, Kidder, and Paul, 1938, pl. 9,c,e); the Monte Fresco phase of Tamarin248

do Bay, Costa Rica (M. D. Coe, personal communication); the Catalina phase of the Middle Tempisque, Costa Rica (Baudez, personal communication); the Olmec culture of La Venta site (Drucker, 1952, pp. 231-32); and the Preclassic site of El Trapiche near Cempoala, Veracruz (García Payón, 1950, pl. 12, no. 5). Zoned incised with red and black paint, a typical Formative period trait of the early Chavin and Cupisnique horizons of Peru, is found in the Chorrera period of Ecuador (Evans and Meggers, 1957, fig. 2,i,j), as well as the Ocos and Conchas phases of coastal Guatemala (Coe, personal communication), and in the Chombo phase on Santa Elena Peninsula, Costa Rica (Coe, personal communication). In addition, zoned red "cuspidor"shaped vessels of the Chorrera and Conchas phases are identical in shape, proportions, and zoned decoration (fig. 4). Rim and body exteriors are red-slipped on a natural colored buff-tan-orange. Ornament consists of vertical incised lines, punctate, rockerstamping, or plain areas on the unslipped buff to light orange neck (Coe, 1960b, p. 369, figs. 4,n,o; 5,n,o). Sherds from the early bichrome horizon at Playa de los Muertos, Honduras, are similar. A distinctive Mesoamerican trait found in Formative period sites in Ecuador is a small, thin-walled, highly polished, pottery napkin-ring earspool slightly flared at one end. These objects, a distinctive trait in the Chorrera phase in Ecuador, are common in the Conchas phase at La Victoria site, Guatemala (Coe, 1960b, cf. figs. 4,a-h; 5,a-f). In fact the specimens are so identical in shape, technique of manufacture, size (1.4-2.7 cm. in length, 2.0-3.0 mm. in body wall thickness, and 2.7-4.8 cm. in diameter at the larger end) and surface finish that separation of a group of napkin-ring earplugs by area depended entirely on the catalogue numbers (fig. 5). Although not found in other Mesoamerican sites quite as early as the Conchas phase, the trait has

MESOAMERICA & ECUADOR

FIG. 4—CUSPIDOR-SHAPED VESSELS, a-b, Conchas phase, Guatemala, c-f, Chorrera phase, Ecuador.

been reported from Formative period horizons at Kaminaljuyu (Kidder, Jennings, and Shook, 1946, p. 215, fig. 91), the Mamom and Chicanei horizon at Uaxactun (Ricketson and Ricketson, 1937, pl. 696), the Middle period at Zacatenco (Vaillant, 1931, pl. 82), Periods I and II of El Arbolillo (Vaillant, 1935, pp. 237-39, fig. 25), and the Pavon period of Pavon site near Panuco, Veracruz (Ekholm, 1944, pp. 467-69, fig. 47,h,k). A few additional traits must be mentioned before we leave the Late Formative period of Ecuador. Plain, polished monochromes in themselves are not diagnostic but, as one of many traits characteristic of the Formative period of Mesoamerica and South America, they add one more bit of evidence to the case for interrelationships at this early period. Polished plainwares constitute 56 per cent of the pottery of the Chorrera period as a whole (Evans and Meggers, 1957, p. 237) and constitute a majority of the plain pottery from the following Mesoamerican complexes: Chiapa I and II of the Chiapa de Corzo site, Chiapas;

the Cotorra phase and Dili phase of the Frailesca region of Chiapas; the Ocos and Conchas phases of La Victoria, coastal Guatemala; the Chombo phase of the Santa Elena Peninsula and the Monte Fresco phase of Tamarindo Bay, Costa Rica; and Tlatilco in the Valley of Mexico (Μ. Ν. Porter, 1953, p. 35, pls. 7-11; Piña Chan, 1958, pp. 35-52, 56-70, 73-91); Lower Tres Zapotes from sites in Veracruz (Drucker, 1943a, p. 47-69); and Playa de los Muertos from the Ulua River drainage, Honduras (Strong, Kidder, and Paul, 1938, pp. 7273). Wide-everted rims with grooves along the upper surface are typical of the polished monochromes, some shapes from the Chorrera phase in Ecuador (Coe, 1960b, fig. 5,k-m) being so similar to specimens from the Conchas phase in Guatemala (Coe, 1960b, fig. 4,m), the Chiapa I and II periods of Chiapa de Corzo, Chiapas (Dixon, 1959, p. 37, fig. 49), and the Escalera and Francesca phases of the Frailesca region of Chiapas (C. Navarrete, 1960, figs. 27,g; 28) that they cannot be separated easily when the sherds are mixed (fig. 6). 249

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

FIG. 5—NAPKIN-RING POTTERY EARPLUGS, a, Chorrera phase, Ecuador. b, Conchas phase, Guatemala.

As we continue to approach the analysis from the point of view of Ecuador, a series of traits can be discerned in the transitional period from Chorrera to Tejar that have distinct affiliations to early horizons in Mesoamerica. They are: zoned red slip or paint on polished and unpolished surfaces without incised lines bordering the decoration, zoned red slip or paint on polished surfaces bordered by broad-line incisions, fine incised lines on polished surfaces, broad-line incisions on polished surfaces, polished red and white slips applied in separate areas on polished surfaces, and polished streaky white slip. The zoned red slip or paint on a polished and sometimes on an unpolished surface, where the red is applied in broad bands and patches, is distinctive of the Conchas phase of Guatemala (Coe, personal communication), the Chombo phase of Santa Elena Peninsula and the Monte Fresco phase of Tamarindo Bay, Costa Rica (Coe, personal communication), Playa de los Muertos, Honduras (Strong, Kidder, and Paul, 1938, pp. 70, 74), Tlatilco in the Valley of Mexico (Porter, 1953, pp. 35-36, figs. 6, 8; Piña Chan, 1958, pp. 44-46, 85, figs. 15, 43), Chiapa I and Chiapa II at Chiapa de Corzo, Chiapas (Dixon, 1959, pp. 12-16, 32-33), 250

and Cotorra and Dili phases in Frailesca region, Chiapas (C. Navarette, 1960, pp. 2426). The polished red and white slips applied independently in broad bands and patches and the polished streaky white slip (fig. 7) are so uniform that thickness of body wall is the principal feature by which sherds of the Chorrera and Tejar phases of Ecuador can be distinguished from those of the Conchas phase in Guatemala (Coe, 1960b, cf. figs. 6,g-j; 7,g-j); the streaky red on a thick white slip of the Monte Fresco phase of Tamarindo Bay, Costa Rica (Coe, personal communication); Playa de los Muertos Bichrome sherds, Honduras (Strong, Kidder, and Paul, 1938, pl. ll,a-e,l,m); Chila White from Period I in the Tampico-Panuco sequences (Ekholm, 1944, pp. 341-43, 42325) and Progreso White, especially the later variants ancestral to Chila White in the Pavon period of Site VC-2 near Panuco (MacNeish, 1954, pp. 566-67); White Monochromes from Chiapa I and Chiapa II at Chiapa de Corzo site, Chiapas (Dixon, 1959, pp. 7-12, 23-31); White Slip from Dili phase in the Frailesca region of Chiapas (Navarrete, 1960, pp. 25-26); white wares from the early periods in the Valley of Mexico sites of Zacatenco (Vaillant, 1930,

FIG. 6—WIDE-GROOVED, phase, Guatemala.

WIDE-EVERTED

RIMS, a-d, Chorrera phase, Ecuador, e-g,

Conchas

FIG. 7—POLISHED R E D AND W H I T E SLIPS, a-f, Chorrera and Tejar phases, Ecuador, g-j, Conchas phase, Guatemala, k-l, Playa de los Muertos horizon, Honduras.

FIG. 8—STRIATED P O L I S H E D POTTERY, α-c, Chorrera phase, Ecuador. d,e, Conchas phase, Guatemala.

FIG. 9—NEGATIVE OR "RESIST" PAINTING, α-c, Tejar phase, Ecuador, d-f, Conchas phase, Guatemala.

FIG. 10—GRATER BOWLS. a,b, Tejar phase, Ecuador. c, Conchas phase, Guatemala.

FIG. 11—POTTERY MASKS FROM ECUADOR AND MESOAMERICA. Whereas all the specimens in Ecuador are from one general region and the Esmeraldas period, those in Mesoamerica are more widespread and developed over a longer period, a, La Tolita, Esmeraldas Province, Ecuador (in Museo Arqueológico Victor Emilio Estrada). b,c, Esmeraldas Province, Ecuador (after d'Harcourt, 1942, pls. 52, 53). d, Las Charcas, Guatemala (after Borhegyi, 1955, fig. 2,a). e, Chipoc, Alta Verapaz, Guatemala (after Borhegyi, 1955, fig. 2,b). f-h, Cerro de las Mesas, Veracruz, Mexico (after Drucker, 1943b, pl. 4 3 ) . cm.

MESOAMERICA & ECUADOR

FIG. 12—FIGURES IN A BED. α-c, La Tolita, Esmeraldas Province, Ecuador (after H. Lehmann, 1951, figs. 1-3). d-f, Valley of Mexico (after H. Lehmann, 1951, figs. 9, 10, 12).

pp. 82-83), Ticoman (Vaillant, 1931, pp. 386-87), and El Arbolillo (Vaillant, 1935, pp. 227-31). Fine- and broad-line incisions on polished to well-smoothed monochrome surfaces constitute another link between the Chorrera period of coastal Ecuador and Preciassic period complexes in Mesoamerica, such as Ocos and Conchas phases, coastal Guatemala; Chiapa I and II from Chiapa de Corzo, Chiapas; Cotorra and Dili phases from Frailesca region, Chiapas; Chombo phase from Santa Elena Peninsula, Costa Rica; Playa de los Muertos horizon in Honduras; and various early horizons from sites in the Valley of Mexico, such as Tlatilco, El Arbolillo, Zacatenco. The incised tradition alone is not conclusive evidence but, linked with the other traits, it assumes greater meaning. Striated polished (sometimes called "burnished lines") on unpolished and polished surfaces is a technique (fig. 8) that increases in popularity in the Chorrera period as the fully polished monochrome plain types decline in frequency. This distinctive method of treating the pottery surface also occurs early in Mesoamerica. Detailed examination of the actual sherds (plain types are too infrequently illustrated to rely on publications alone) shows that this technique is characteristic of the Conchas phase of coastal Guatemala, is present in Chiapa I and Chiapa II of the Chiapa de Corzo

site of Chiapas, is common in Playa de los Muertos materials from Honduras, is a diagnostic pottery feature of the Monte Fresco phase of Tamarindo Bay, Costa Rica, occurs in the early horizons of El Arbolillo and Zacatenco I in the Valley of Mexico, in Pavon in the Huasteca region, and in Lower Tres Zapotes in Veracruz. REGIONAL DEVELOPMENTAL

(REGIONAL

FLORESCENT, CLASSIC) PERIOD

As has been pointed out, this survey of Mesoamerican-Ecuadorian connections is being undertaken primarily from the point of view of Ecuador. In the Ecuadorian framework of archaeological periods (see Estrada, 1958, pp. 7-20, Chart 1; Estrada and Evans, 1963, fig. 10), the Formative has been defined as ending with the end of the Chorrera phase, which initiates the development of a group of regional complexes. These share certain features as a result of a common ancestry, but differ in other details. Some of these details have Mesoamerican parallels. Two of the parallels are negative or resist painting and a form of pottery vessel designated as a "grater bowl." Negative painting begins in Ecuador as simple line-and-dot designs on well-polished, buff, cream, or light tan surfaces in the Tejar phase. The technique and motifs (fig. 9) are comparable to negative on White-to-Buff sherds of the Conchas phase of coastal Guatemala 253

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

FIG. 13—FIGURINES WEARING FEATHER CLOAKS OR COSTUMES REPRESENTING BIRDS, a, Probably represents a warrior. La Tolita, Ecuador. Height 22.5 cm. b, Mold for figurine manufacture. La Tolita, Ecuador. Height 11.7 cm. c, Eagle costume. Valley of Mexico. Height 13 cm.

(Coe, 1960b, fig. 7,a-f), to resist-painted sherds from Chiapa III and IV of Chiapa de Corzo, and to negative-painted sherds from burials at Tlatilco (Porter, 1953, p. 27, fig. 3 ) . Pottery grater bowls of the Tejar phase are of two forms: large, open, round-bottomed bowls and broad, flat-bottomed rectanguloid bowls with high, out-slanting walls and a mouth or spout at one end (fig. 10). The shapes are not typical of the Conchas phase, but the deep grooving or scoring on the inside of the vessel bottoms is paralleled on specimens of Conchas White-to-Buff (Coe, 1960b, cf. figs. 6,k,l; 7,k-m). The deep, grooved grater bowl does not persist in Mesoamerica but evolves into the shallow, Mexican "chili-grinder" widespread in the Postclassic horizons. In Ecua254

dor the shallow variety of grater vessel also appears, but does not completely replace the deep, grooved variety. Coe (1960b, pp. 370-71) suggests the possibility that this appearance of the grater bowl in the Ecuadorian sequence might reflect the introduction of chili grinding from Guatemala; unfortunately the wet climates of the two regions do not allow preservation of the vegetable remains to prove this theory. The Guangala phase of the north coast of Guayas Province, Ecuador, shares negative painting with the Tejar phase, and exhibits some other features with Mesoamerican parallels. One is a style of whiteon-red painting with bold geometric designs that resembles closely, not only in decoration but also in vessel shape, examples of white-on-red of the Matapalo

MESOAMERICA & ECUADOR

FIG. 14—HUMAN FIGURES, PROBABLY WARRIORS, WITH HEAD APPEARING FROM OPEN JAWS OF ANIMAL MASK. a,b, Mexico, c-f, Ecuador.

phase of Tamarindo Bay, Guanacaste region, Costa Rica (Coe and Baudez, 1961, p. 508; Baudez and Coe, 1962, p. 368; Coe, 1962a, p. 363-65). This may be an example of an introduction of a trait from Ecuador, since it seems at our present state of knowledge to be more widespread there than in Mesoamerica. Polychrome painting in bold geometric motifs on a white-slipped surface appears in the late Guangala phase, apparently without direct local antecedents.

Polychrome painting has been stratigraphically defined on the coast of Costa Rica by Coe and Baudez (1961, pp. 505-15; Baudez, 1963, pp. 46-47, fig. 6, Table 2; Baudez and Coe, 1962) at an early enough time level to be the source of the Ecuadorian occurrence. The general outlines of sequences in the Provinces of Manabi and Esmeraldas are well enough known to indicate that a large series of traits of the Regional Develop255

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

FIG. 15—THREE-PRONGED INCENSE BURNERS, ESMERALDAS PROVINCE, ECUADOR. Height of a,b, 26.7 cm.; d, 17.3 cm. (After Borhegyi, 1959, fig. l,a-c; 1960, fig. 1.)

FJG. 16—BAT-JAGUAR FIGURES, a, La Tolita, Esmeraldas Province, Ecuador. Height 46 cm. Museo Arqueológico Victor Emilio Estrada, b, Monte Alban, Oaxaca, Mexico. Height 22 cm. (after Caso and Bernal, 1952, fig. 113,b).

256

mental period (500 B.C.-A.D. 500) are more closely related to Mesoamerica than to South America. These traits are concentrated in the region from Bahia de Caraquez northward, including such well-known sites as Atacames, Esmeraldas, and La Tolita along the coast of the Province of Esmeraldas. Comparable material has been reported across the border in Colombia around Tumaco in the state of Nariño (Cubillos, 1955). The time period involved is probably around A.D. 0-500. A listing of such traits, with emphasis on their Mesoamerican origin, is not new; the new evidence is the presence of detailed stratigraphic sequences for the coast of Ecuador. These sequences clearly indicate that the appearance of this complex of traits is not the result of evolutionary development from indigenous pre-existing Ecuadorian elements, but rather the result of the introduction of a series of non-South American elements into a local situation (Estrada and Evans, 1963, pp. 82-84). Reference to Borhegyi's various inventories (1959, 1960) is worthwhile because only a few most striking features will be summarized here. Significant traits include: an abundance of moldmade zoomorphic and anthropomorphic pottery figures; threepronged pottery incense burners; pottery figures "tied to a bed"; effigy whistles; figurines with the human head protruding from the open jaw of an animal; pottery masks; flat and roller pottery stamps; figurines dressed in feather capes; full-size pottery human heads; yellow, orange, green, and asphaltum paint on pottery figurines or pottery appendages; bat and jaguar figures sometimes represented with a human body; fertility-cult figurines with widespread legs; representation of an old man with wrinkled face; the bifurcated tongue in zoomorphic figures; and mammiform legs of vessels. The closeness of the similarities can be appreciated by referring to items on the same figure, noting in the caption which

FIG. 17—MOLDMADE ZOOMORPHIC AND ANTHROPOMORPHIC FIGURINES, ECUADOR. a,b, Northern Manabi Province, Museo Arqueológico Victor Emilio Estrada. c-g, Mubuque, northern Manabi Province. U.S. National Museum. f, 8.3 cm. tall, g, 12.8 cm. tall.

FIG. 18—FLAT AND CYLINDRICAL POTTERY STAMPS FROM REGIONAL DEVELOPMENTAL CULTURES OF NORTHERN MANABI AND SOUTHERN ESMERALDAS PROVINCES, ECUADOR, a-v, Same scale. Length of a, 6 cm. Length of w, 7.5 cm. x,y,gg-jj, Same scale. Height of x, 6 cm. z,αα, Same scale. Height of z, 7 cm. bb-ff, Same scale. Height of bb, 5.8 cm.

MESOAMERICA & ECUADOR

specimens are Ecuadorian and which are Mesoamerican: pottery masks (fig. 11), reclining human figures on a cradle or bed (fig. 12), human figurines in feather cloaks (fig. 13), figurines with human head protruding from the open jaw of an animal (fig. 14), three-pronged pottery incense burners (fig. 15), bat and jaguar figures (fig. 16), moldmade zoomorphic and anthropomorphic figurines (fig. 17), cylinder and flat stamps (fig. 18), the old man with wrinkled face (fig. 19), and figurines with movable arms (fig. 20). Extensive combing of the literature and examination of museum collections in the United States, Latin America, and Europe relate this complex of traits on the north coast of Ecuador and the south coast of Colombia to no one specific site in Mesoamerica, but do narrow it down to several distinct regions. All other areas at the present state of knowledge can be eliminated. The traits appear about the beginning of the Christian Era in areas of Veracruz, the Valley of Mexico, and Oaxaca (Piña Chan, 1963, fig. 4). Many occur in the Upper Tres Zapotes period, represented by such sites as Tres Zapotes and Cerro de las Mesas, in Veracruz; the Monte Alban III-IV periods of Oaxaca; and the Toltec and Aztec periods of the Valley of Mexico. This was a time when these Mesoamerican cultures were expanding their territorial control and it is not inconceivable that movements may have been planned as far away as South America, establishing trade with northern Ecuador, perhaps developing outpost communities that were amalgamated into the local aboriginal cultures of this part of South America. INTEGRATION ( POSTCLASSIC, EMPIRE, AND CONQUEST) PERIOD

Shaft tombs make their appearance in Ecuador about A.D. 400-500, at the end of the Regional Developmental and the beginning of the Integration period. They consist of a vertical neck or shaft 0.75-1 m. long, capped

FIG. 19—OLD MEN W I T H W R I N K L E D FACES. a, Agua Amarga, northern Manabi Province, Ecuador. c, Viche, Manabi Province, Ecuador, e, Esmeraldas, Esmeraldas Province, Ecuador, b, Mexico. 5.8 cm. tall. d,f, Tres Zapotes, Veracruz, Mexico. Height of d, 9.4 cm. a,b, U.S. National Museum. c,e, Museo Arqueológico Victor Emilio Estrada. d,f, after Drucker, 1943a, fig. 60,c,d.

about 1 m. below the surface with stone slabs. The enlarged chamber measures 2 2.5 m. in diameter at the floor, which is 2.5-4 m. below the surface (fig. 21). Tombs of this type occur at El Carmen in Canton Guaranda, Parroquia Guanujo, Province of Bolivar (Costales Samaniego, 1956); at Las 259

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

FIG. 20—HUMAN FIGURINES W I T H MOVABLE ARMS AND LEGS. a,b, Tisal, northern Manabi Province, Ecuador. U.S. National Museum, c, Tres Zapotes, Veracruz, Mexico. Length of last specimen on right: 7 cm. U.S. National Museum.

Tolas de Huaraqui, Canton Pedro Moncayo, Province of Pichincha (Guignabaudet, 1953); at the site of Chaupihuaca, in the same canton and province; at Santa Elena cemetery, Province of Tungurahua; at San Pablo cemetery, Parroquia San Andres Cachi-Huayco, Province of Chimborazo and at El Angel site, Hacienda Pucara, Province of Carchi; and in the intermontane valleys from the Province of Carchi to the Province of Loja (Verneau and Rivet, 1912, p. 124, figs. 17-22). Although comparatively rare

FIG. 21—BOTTLE-SHAPED TOMB, ECUADOR. Typical of highland Ecuador and southern Colombia. (After Costales Samaniego, 1956, fig. 4.)

260

on the Ecuadorian coast, bottle-shaped tombs are reported by Saville (1910, pp. 8285) at La Roma and near Cerro Jaboncillo on the Manta peninsula of Manabi Province. It is pertinent to note that the major occurrence of bottle-shaped and antechambered shaft tombs is in the Cauca Valley of Colombia (Bennett, 1946, pp. 834-36), where they appear to begin around A.D. 400-500 and continue until a few hundred years before European contact (fig. 22). The shaft tombs of western Mexico include both the bottle-shaped and the antechambered types. Bottle-shaped tombs have been described by Corona Núñez (1954, pp. 46-47) from the site of El Llano in the Municipio of San Blas, from sites in the Municipio of Santa Maria del Oro and Ixtlan in the state of Nayarit (fig. 23). Their shape, proportions, and measurements are very close to the Ecuadorian tombs. Shaft tombs with a large antechamber to one side (fig. 24) have been described by Noguera (1946, pp. 150-54; 1939, pp. 57486, fig. 14) from the site of El Opeño in the region of Zamora in the state of Michoacan, by Corona Núñez (1955, pp. 7-8, figs. 1, 2), from El Arenal near Etzatlan in the state of Jalisco, and also by Corona Núñez (1954, pp. 47-48, figs. 6, 7) from the site of Corral Falso, Municipio de Santa Maria del Oro, and the site of Los Chiqueros near Ixtlan, both in Nayarit. They are apparently associated with late styles. The limited distribution in Mesoamerica of both the bottleshaped shaft tomb and the antechambered shaft tomb to the three adjoining states of Nayarit, Jalisco, and Michoacan on the central Pacific coast of Mexico, and the rather widespread occurrence of the traits in the Ecuadorian highlands and southern Colombia suggest that the trait might have been introduced from South America to Mesoamerica (cf. Meggers, 1963, fig. 20). Drilled teeth inlaid with gold in the form of discs, pegs, or plates (fig. 25) are restricted in Ecuador to the Milagro period cultures of the Guayas Basin and the

MESOAMERICA & ECUADOR

FIG. 22—ANTECHAMBERED TOMBS, SOUTHERN COLOMBIA. Also typical of highland Ecuador. (After Bennett, 1946, fig. 92.)

"mound" cultures so well known from such sites as La Tolita, Atacames and Esmeraldas in Esmeraldas Province. Whereas in Mesoamerica filing the teeth and inlaying with iron pyrites and jadeites and other stones has a long history, the use of gold is very restricted (see Chart 12, Romero, 1958). Since the custom of tooth mutiliation is so widespread in time and space in Mesoamerica and restricted in South American archaeological remains to the coast of Ecuador, it can be concluded that the trait was introduced from Mesoamerica. Few skeletons have been saved in the past, but preserved Ecuadorian specimens of filled teeth include

FIG. 23—BOTTLE-SHAPED TOMB, NAYARIT, MEXICO. (After Corona Núñez, 1954, fig. 1.)

FIG. 24—ANTECHAMBERED TOMB, NAYARIT, MEXICO. (After Corona Núñez, 1954, fig. 6.)

261

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

FIG. 25—GOLD-FILLED T E E T H , ECUADOR. a, Gold discs from skull, Atacames, Esmeraldas Province, Ecuador (after Romero, 1958, pl. 7 ) . b, Cut teeth filled with gold plates, from skull, La Piedra, Esmeraldas Province, Ecuador (ibid., pl. 8 ) . c, Gold pegs with widened heads, from teeth, Site G-M-4, Elisita, Guayas Province, Ecuador.

the following: two incisors and one canine tooth showing gold pegs with expanded heads (Romero's Type E-4) from a chimney-urn burial of the Milagro period from Site G-M-5, Elisita, Guayas Province (Estrada, 1957a, fig. 11A); two central incisors 262

in an upper maxilla from Atacames, Ecuador, filled with discs of gold (Romero's Type E-l; Saville, 1913, pl. 16, fig. 56; Romero, 1958, pl. 7); and six teeth in the upper maxilla including the incisors, lateral incisors, and canines, with wide horizontal plates of gold (Romero's Type E-4) from the site of La Piedra on the Rio Esmeraldas in Esmeraldas Province, Ecuador (Saville, 1913, pl. 17, fig. 57; Romero, 1958, pl. 8). Dental mutilation of Romero's Type E-4 is not found in Mesoamerica and is unique in South America to the coast of Ecuador, in the area of the "mound" cultures. Type E-l is the most common technique of inlay in Mesoamerica but it is typically represented by pyrites, jadeite, or turquoise instead of gold. The distribution of Type E-l includes Cerro de las Remojadas and Mesas in Veracruz; Monte Alban, Monte Negro, Xoxo, and Yagul in Oaxaca; Jonuta in Tabasco; Jaina and other sites in Campeche; Palenque, Chiapa de Corzo, and Yoxiha in Chiapas; Tepeaca in Puebla; Teotihuacan and Tacuba of the Federal District; Baking Pot and San Jose in Belize, British Honduras; Uaxactun, Piedras Negras, Kaminaljuyu, and Holmul in Guatemala; and Copan and the Valley of Ulua, Honduras. However, the only instance of Type E-l in the hundreds of teeth examined by Romero (1958) in which gold was used is cat. no. TD-451 from the Monte Alban IV and V periods of Yagul, Oaxaca, on display in the Museo Arqueológico Regional de Oaxaca (verified by us in January 1960). As early as Saville's writings (1913), tooth mutilation in Ecuador was pointed out to be related to Middle America, but although this conclusion has been strengthened considerably by more recent work, the exact point of origin is still unknown. The exact time at which the introduction occurred is also uncertain, but it clearly took place sometime during the Integration period beginning not earlier than A.D. 500600 and extending until the time of Euro-

MESOAMERICA & ECUADOR

pean conquest. Romero's extensive studies (1952, 1958) and his compilation of the data on a time chart (1958, Chart 12) show the greatest concentration of tooth mutilation of Types E-l and E-4 to occur over the widest geographical area of Mesoamerica from around A.D. 500 to 1300. Of particular interest is the fact that in Mesoamerica tooth mutilation Type E-l has the longest history in the state of Oaxaca, suggesting that this region was the focus of development of this particular type of decorative tooth inlay and might well be the source of its introduction to coastal Ecuador. Another trait of the Milagro period of the Guayas Basin of Ecuador with Mesoamerican connections is copper ax-money. These thin, T-shaped, hammered pieces of copper with a crescent blade have flanged edges produced by hammering along the edge after flattening (fig. 26). They range in total length 6-9 cm., in width of curved blade 5-8 cm., and in width of butt end 1.8-2.2 cm. Large caches of ax-money have been found in the chimney-urn burials of the Milagro period, with over 1000 examples coming from a single burial in the site of Pedro Carbo (Site G-27) west of the town of Nobol on the road to Manabi in Guayas Province. Copper ax-money in the form of crescentshaped "knives" made of thinly beaten copper was one of the means of exchange used by the Aztecs (Vaillant, 1941, p. 128; Blom, 1934, pp. 423, 437). Archaeological distribution of the specimens in both the Valley of Mexico (USNM cat. nos. 215390, 306966) and in Oaxaca, at such sites as Zimatlan (USNM cat. no. 97785), suggests that the trait was also characteristic of the ZapotecMixtec tradition. CONCLUSION

The better known the archaeological record becomes, the more apparent is the conclusion that contact must have repeatedly taken place between Mesoamerica and the

FIG. 26-COPPER AX-MONEY, LAS PALMAS SITE, GUAYAS PROVINCE, ECUADOR.

west coast of South America. Beginning in the Early Formative period, communication was probably at first random and accidental, but in the centuries just prior to the arrival of Europeans it appears to have become planned and continuous. In South America the principal area of contact centers around the north coast of Ecuador, lying just north of the Equator. About this point, the southmoving ocean currents are replaced by the north-flowing Humboldt Current, which may have been a factor in impeding navigation farther southward. Although the archaeological evidence is extensive, it must be recalled that both in Mesoamerica and in coastal Ecuador climate destroys a major portion of what was once made and used by the aboriginal populations. It is probable that interchange of ideas and objects was on a far broader scale than the surviving evidence leads us to infer. Such a conclusion is supported by the description of a balsa raft encountered by the Spanish off the Ecuadorian coast, loaded with merchandise destined for a northern port, unfortunately unnamed (Ruiz, 1884). Of its large and varied cargo, little or nothing would have been preserved for an archaeologist to find. 263

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

REFERENCES Aveleyra, 1956 Baudez, 1963 and Coe, 1962 Bell, 1960 Bennett, 1946 Blom, 1934 Borhegyi, 1955, 1959, 1960 Brainerd, 1953 Caso and Bernal, 1952 Coe, M. D., 1960b, 1962a and Baudez, 1961 Corona Nunez, 1954, 1955 Costales Samaniego, 1956 Cubillos, 1955 D'Harcourt, 1942 Dixon, 1959 Drucker, 1943a, 1943b, 1952 Ekholm, 1944 Estrada, 1956, 1957a, 1957b, 1957c, 1958, 1962 and Evans, 1963 Evans and Meggers, 1957 , , and Estrada, 1959 García Payón, 1950 Guignabaudet, 1953 Jijón y Caamaño, 1914, 1930, 1951

264

Kidder, Jennings, and Shook, 1946 Lehmann, H., 1951, 1953 Lowe and Navarrete, 1959 MacNeish, 1954, 1958 Mayer-Oakes, 1963 and Bell, 1960 Meggers, 1963 and Evans, 1962 Navarrete, C , 1960 Nicholson, 1953 Noguera, 1939, 1946 Piña Chan, 1958, 1963 Porter, Μ. Ν., 1953 Ricketson and Ricketson, 1937 Romero, 1952, 1958 Ruiz, 1884 Saville, 1907, 1909, 1910, 1913 Strong, Kidder, and Paul, 1938 Sullivan and Hellman, 1925 Uhle, 1922, 1923a, 1923b, 1927, 1931 Vaillant, 1930, 1931, 1935, 1941 Verneau and Rivet, 1912 Willey, 1955a, 1958 Wormington, 1957 Zevallos Menéndez and Holm, 1960

13. Relationships between Mesoamerica and the Andean Areas

DONALD

state of knowledge conT cerning Mesoamerica and Peru does HE PRESENT

not allow us to discuss the cultural relationships between these two areas in a definitive or satisfactory way. A few years ago there was less specific information available, so it could be fitted into relatively simple models of cultural evolution. The vast recent increase in precise data concerning the sequences in Mesoamerica and Peru, and especially in the areas between them, appears to require more elaborate conceptual schemes if they are to be fitted into a coherent pattern. A major difficulty in discussing these problems is that much of the significant information is unpublished or is published only in a preliminary and cursory way. Although widely disseminated by word of mouth and correspondence among the specialists, it can not be fully documented in terms of published references. Reliance only on these would put us between five and ten years out of date. I have consequently made extensive use of personal communications, but it has been impossible in the time avail-

W.

LATERA?

able to contact all the people with relevant unpublished information.1 There is general agreement among most experts that a network of cultural relationships connecting all Nuclear America was an important factor in the cultural history 1

I am indebted to several people for assistance on this article, primarily to Edward P. Lanning, who personally instructed me on the latest unpublished material on the Early cultures of coastal Peru. He also made his Ph.D. dissertation available and helped in making Peruvian comparisons for a number of Mesoamerican specimens in the Museum of Anthropology of the University of California. His assistance warrants joint authorship in this paper, but he considers it premature to discuss these problems and wishes to disassociate himself from any general conclusions. Donald Collier brought the Haldas figurine illustrations to my attention and supplied copies of letters concerning the figurine from Michael D. Coe, Junius Bird, Edward P. Lanning, and Gordon R. Willey. Bird's letter is cited as personal communication. Coe's observations on the Peruvian figurines strengthen the case for Mesoamerican origins but are not included in the present paper. Discussions on problems of Early Mesoamerican chronology with René Millón and James Bennyhoff were of great assistance, as were general discussions of the general problems of the paper with John H. Rowe and Robert F. Heizer.

265

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

of this region. From a relatively early period up to contact times innovations in one section had an excellent chance of spreading through most parts of this large region, a point demonstrated by Willey (1955a). Beyond this point, however, and in terms of contact between specific areas of Nuclear America, such as the central Andes and Mesoamerica, it is difficult to speak with assurance. The nature of such contacts— whether the result of trade, individual-toindividual diffusion according to the model recently presented by Edmonson (1961), or actual migrations of sizable groups of people—becomes more elusive. Often it is difficult even to specify the direction of such cultural transmission. It is impossible here to list all alleged cultural similarities and present the history of thought on cultural relationships between the two areas. Instead of such an unevaluated catalogue of names and traits, I offer a few cases and some detail as to their cultural and temporal context. EARLY FORMATIVE CONNECTIONS

The idea that the two areas of high civilization in the New World arose from an essentially uniform cultural base is an old one in studies of New World culture history. It was presented in its most typical form by Spinden (1917b) and has frequently been referred to as the Spinden Hypothesis. It has fluctuated in popularity but during the middle and late fifties it had a great vogue; many discussions of cultural connections within Nuclear America have been framed in terms of some or all of the assumptions of the Spinden Hypothesis (G. and A. Reichel-Dolmatoff, 1956; Evans and Meggers, 1957; Evans, Meggers, and Estrada, 1959; Lathrap, 1958). Most simply stated, its basic ideas are: (1) the earliest ceramic traditions of South America represent a north-to-south spread out of Mesoamerica; (2) the earliest centers of Mesoamerican ceramics coincide with the early centers of seed-crop agriculture 266

(maize, beans, squash); (3) the spread of such basic agricultural patterns connects fairly closely with the spread of pottery; (4) many of the basic and significant cultural similarities shared by Nuclear America are the result of this expansion of a single, fairly uniform culture. What is involved here is essentially the spread of a neolithic economy. One can use the basic features of the Spinden Hypothesis without committing oneself to the mechanism of the spread. Evans, Meggers, and Estrada appear to favor migration as a major factor, as do the Reichel-Dolmatoffs in their Momil report; in so far as such a spread can be demonstrated, it seems to me that an outward growth of farming communities according to the generalized model presented by V. Gordon Childe (1942, pp. 66-67) is the most likely mechanism. Edmonson (1961) has interpreted the same data in terms of individual-to-individual communication through a stable network of people, an hypothesis which seems remarkably unlikely in terms of our ethnographic knowledge of contacts between agricultural and nonagricultural groups. Recent archaeological excavations in Nuclear America have tended not to support certain aspects of the Spinden Hypothesis. Work in Mesoamerica indicates that the basic pattern of seed-crop agriculture developed in the relatively arid highlands, and that it had a long history before it became associated with ceramics of any sort (MacNeish, 1958, 1961, 1962; Mangelsdorf, 1954; Mangelsdorf, MacNeish, and Galinat, 1964). Increasing data from northern South America and lower Central America indicate that ceramics are older here than in either Mesoamerica or Peru. The pottery of Valdivia in coastal Ecuador (Evans, Meggers, and Estrada, 1959), Monagrillo in Panama (Willey and McGimsey, 1954), Puerto Hormiga and Barlovento in coastal Colombia (ReichelDolmatoff, 1955, 1965), Momil I in northern Colombia (G. and A. Reichel-Dolmatoff, 1956); the earliest pottery in the

MESOAMERICA & ANDEAN AREAS

Saladoid tradition in the Orinoco Valley of Venezuela; and the earliest pottery in the Barrancoid tradition also in the Lower Orinoco (Cruxent and Rouse, 1958-59) all appear to be about as early as, or in some cases considerably earlier than, the earliest securely dated ceramics from Mesoamerica. None of these show convincing similarities to the early ceramic traditions of Mesoamerica. What is remarkable about these ceramics of northern South America is their relatively great diversity of form and decoration. Evans, Meggers, and Estrada (1959) have stressed the similarities among Valdivia, Barlovento, and Monagrillo, but by picking other individual "traits" one could make an equally strong argument for their diversity, whereas the other mentioned complexes show an even wider range of form and decoration. This ceramic diversity suggests a long period of experimentation behind these earliest known ceramic complexes dating from the third, second, and early first millennia B.C. At present northern South America appears a more likely hearth for ceramics than does Mesoamerica. There is good evidence that some of these ceramic complexes are associated with intensive root-crop agriculture (Cruxent and Rouse, 1958-59; G. and A. Reichel-Dolmatoff, 1956), and I have suggested (1962) that part of this proliferation and spread of ceramics is associated with a neolithic expansion based on root-crop agriculture. If these interpretations are correct, the expansion of a single Formative or Neolithic culture pattern is too simple a model to encompass the culture history of Nuclear America. Though the Spinden Hypothesis is apparently not completely true, it should not be entirely discarded. The spread through much of Nuclear America of seed-crop agriculture and certain ceramic elements, both from a Mesoamerican source, is supported by many of our data. One must strike a median between looking to Mesoamerica as the source of all early ceramic traditions in Nuclear America, and rejecting out of hand

the possibility of Mesoamerican influence on, or of origins for, some of the earlier ceramic traditions of northern South America, and especially of the central Andes. The relatively early existence of a pattern of rudimentary agriculture on the coast of Peru was first demonstrated by Junius Bird (1948). Its wide geographical spread has been indicated by Engel (1957a, 1957b). In this pattern the more important cultivated plants were used for textiles and for containers rather than for food; the few cultivated food plants seem to have made a relatively minor contribution to the diet. With the possible exceptions of Lagenaria and a squash Cucurbita moschata, there is no real evidence suggesting Mesoamerican connections for this early incipient agricultural pattern. The occurrence of the bottle gourd, Lagenaria, both in this context and in certain of the preceramic, early agricultural complexes of Mesoamerica is of interest (Bird, 1948; MacNeish, 1958; Cutler and Whitaker, 1961; Whitaker and Bird, 1949). The first indisputable indication of Mesoamerican influence in Peru is maize. There seems little doubt that maize was originally brought under cultivation in Mesoamerica (MacNeish, 1961, 1962; Mangelsdorf, MacNeish, and Galinat, 1964). This does not rule out the possibility of secondary centers of diversification in Peru or of the hybridization of the original maize with South American grasses (Grobman ef al., 1961). Initial appearances of maize in costal Peru show a curious pattern. In the northcentral coast we have what may well be its earliest appearance in a preceramic midden at Culebras (Lanning, 1960b, p. 586), and work at the Aspero site somewhat to the south seems to confirm the existence of maize in a preceramic context (Willey and Corbett, 1954, pp. 78, 151). Excavations farther north in the Chicama Valley suggest that maize does not appear until considerably later than the first introduction of pot267

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

tery (Bird, 1948, p. 27). Though other interpretations might be possible, the most economical one seems to be that maize appeared first somewhere south of the Casma Valley and only somewhat later spread to the north coast. The introduction of maize at this point on the Peruvian coast would require a route of diffusion through the Andes or Montaña or a very long sea journey bypassing the northern part of the Peruvian coastal strip. General considerations of practicality plus certain botanical evidence seem to favor the former alternative, and Collier (1959, p. 5) has recently argued for its acceptance. At present it would be difficult to specify the means of transmission from Mesoamerica to Peru. The earliest ceramics of coastal Peru offer other possible examples of Mesoamerican influence. Our knowledge of pre-Chavin pottery in this area has been greatly increased in recent years, mainly by the work of Lanning (1960a,b), according to whom it now appears that there was a considerable span of time between the initial introduction of pottery and the earliest of the major stylistic horizons in Peru, the Chavin horizon. The best available evidence indicates that ceramics were present on the north and north-central coast by 1400 B.C. and a recent carbon-14 date from Las Haldas suggests their presence there by 1600 B.C. (Tokyo Scientific Expedition, 1960, p. 518). The Initial period thus has a possible time span of 600 years.2 The Initial period pottery of coastal Peru shows some uniformity, and all the pre-Chavin ceramic complexes of coastal Peru from Chicama to the south-central coast appear to be members of a single ceramic tradition (Lanning, 1960a, p. 11). The most characteristic form in this ceramic tradition is the large neckless olla, which in most of these complexes occurs as the dominant form, and in some of the earliest complexes the only form. Such vessels typically have an internal thickening on the rim, giving them a characteristic "comma" profile. Lanning has recently discovered that 268

this particular rim form is associated with two vessel variants, a deep egg-shaped storage vessel with a subconical base, and a squatter, more globular cooking vessel. In Early Guañape, the most fully published of the Initial period complexes, such ollas are said to be associated with flat-bottomed, sloping-sided vessels with unmodified rims (Strong and Evans, 1952, fig. 35,1,4; tables, p. 255). Lanning, however, has cautioned me about the full acceptance of this association until Bird's excavations at Huaca Negra and at the early ceramic middens near Huaca Prieta, done by natural rather than arbitrary stratigraphy, are published. Elsewhere the Initial period ceramic complexes show the following association of vessel forms. In Haldas I and II, the earliest known complexes from the north-central coast in the vicinity of Casma Valley, the large neckless ollas are almost the only shapes occurring (Lanning, 1960b, p. 483). This is also true of the Chira complex of the central coast, probably of approximately the same age (Lanning, 1960b). At a somewhat later time on the central coast, but still far predating the Chavin horizon, the earlier occupations at the Curayacu site show the additions of shallow, sharply carinated bowls and single-spouted bottles. A number of decorative innovations at this point include: zoned red painting, the extensive use of white slips which are typically highly burnished, and the use of red and white bichrome schemes. Evans, Meggers, and Estrada (1959) have commented on the affiliations of the Initial period ceramics of Peru, making specific comparisons between Valdivia complex of Ecuador and Early Guañape. 3 I have already expressed the opinion (1960, p. 126) that these comparisons are not con2 This terminology is in accordance with that established by Rowe (1960) and somewhat modified by Lanning (1960b). 3 Evans and Meggers have modified their position since the above was written (Meggers and Evans, 1964).

MESOAMERICA & ANDEAN AREAS

vincing. Lanning (1960b, p. 550) and Collier (personal communication) have both indicated similar doubts. Most of the points of comparison which Evans, Meggers, and Estrada use are highly generalized. A small number of decorative techniques, especially notched appliqué, are shared by the two complexes. These few similarities seem of minor significance in the face of the fact that not one vessel form is shared by the two complexes, and that rim treatment in general is very different. Evans, Meggers, and Estrada achieve comparisons in rim form only by placing Valdivia bowl rims by the side of Early Guañape olla rims, a questionable procedure. If Early Guañape as defined by Strong and Evans can be accepted as a valid assemblage, then there are rather striking similarities between this ceramic complex and some of the earlier ceramic complexes of Mesoamerica. These similarities are mainly in terms of vessel form and in the percentages of various vessel forms in the total ceramic complex. Neckless ollas with round or subconical bottoms and internally thickened lips are the most common vessel form in Chiapa de Corzo I, which may well be one of the very earliest fully defined ceramic complexes known from Mesoamerica. The comparisons between olla rims of Chiapa I and those of Early Guañape, or of Haldas and Chira, yield very close formal similarities (Strong and Evans, 1952, fig. 1; Dixon, 1959, fig. 19; Lanning, 1960b; Tokyo Scientific Expedition, 1960). Flat-bottomed, outward-sloping vessels, and flat-bottomed, straight-sided vessels are also present in significant proportions. This pattern of preference in vessel forms also appears to be present in the Ocos material, M. D. Coe (1959, 1960b). This basic pattern of vessel forms is present in the ceramics of La Venta and other "Olmec" sites, though partially masked by the occurrence of a greater variety of more elaborate vessels. According to Drucker (1952, p. 117, fig. 39,a), at La Venta rim sherds from neckless ollas

with internally thickened rims are among the most common. I find these similarities in basic vessel form between certain of the early Mesoamerican ceramic complexes and the Initial period pottery of Peru far more impressive than the similarities which Evans, Meggers, and Estrada note between Valdivia and Early Guañape. Some of the decorative innovations in Curayacu I also have a general Early Mesoamerican look. The sharp carinations, the extensive use of burnished white slips, and the white and red color schemes can all be duplicated in very early complexes such as Chiapa I and Early Zacatenco. On the other hand, close parallels to the earliest ceramic groups in northern South America are lacking. It is nonetheless difficult to interpret the significance of these comparisons. The particular vessel shapes involved do not appear to be prominent in the other better-documented instances of Mesoamerican influence in northern South America. They are rare or absent in the Momil II materials of Colombia (G. and A. Reichel-Dolmatoff, 1956, fig. 7; table, p. 192), and are apparently not prominent in Chorrera in Ecuador (Estrada, 1958), though here a more complete publication of the Chorrera materials may rectify this impression (Coe, 1960b, p. 369). If the vessel shapes of the Initial period ceramics, and indeed the whole ceramic complex, are of Mesoamerican derivation, these influences represent a line of cultural contact independent of the Mesoamerican influences so far demonstrated in Colombia and Ecuador. One must invoke either a rapid transmission by boat, skipping the coast of Colombia and Ecuador, or a series of relatively rapid land migrations which have thus far escaped detection in northern South America. In the face of these difficulties the similarities in pottery shape should be given most careful scrutiny before they are granted acceptance as valid evidence of cultural contact. Solid, handmade figurines were never numerically significant in the early ceramic 269

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

complexes of the central Andes. In contrast to the profusion of such artifacts in most Mesoamerican "Formative" contexts, they are rare in Peru. Nonetheless, their nature and temporal distribution in this region may throw some light on the question of Mesoamerican-Peruvian contacts. Such figurines appear to be associated with both the Initial period ceramics and with Chavin horizon materials, though the second association is less certain than the first. The existence of these figurines has been known since Uhle's early work at Ancon (1906, fig. XVIII), and their possible Mesoamerican affinities have been remarked on from time to time (Strong, 1925, p. 154). Recent work at Curayacu and Las Haldas 4 has increased the size of the sample and provided more complete data on the context of the figurines. In particular, the recently published photographs of a figurine head from Haldas has impressed a number of specialists as looking markedly Mesoamerican (Tokyo Scientific Expedition, 1960, p. 103). This figurine shares several significant traits of manufacture with previously published examples from early coastal Peru. Since the Haldas figurine is the most clearly illustrated, we will begin by comparing it to possible Mesoamerican prototypes. The closest similarities would seem to lie with Drucker's Type I-A-l figurines from Tres Zapotes and La Venta 5 (Drucker, 1952, pl. 23; 1943a). These appear to be the earliest figurine style so far isolated on the Veracruz coast. What is shared is a profile rising to a rather sharp point with a straight or concave back of the head and a markedly sloping face and forehead. Appliqué is added to indicate headdress or hairdo, small deep punctation is extensively used to indicate features. The most specific similarities lie in the eye treatment, carefully 4

Ed. note: This site is also known as Las Aldas. This is my opinion. Several of the specialists writing to Collier expressed other opinions as to the precise type, but all looked to La Venta and Tres Zapotes to make their comparisons. 5

270

reconstructed by Drucker for the Tres Zapotes specimens (Drucker, 1952, pl. 24). Two curved, incised lines are first drawn, indicating the general shape of the eye. The incised lines have deep pits at either end. A very deep, central punctation is then made forming the pupil, and also forming a high ridge of clay around the pupil which separates the long incision into three distinct segments: the central circular pupil, and two triangular areas having deep pits at their outer apexes. It is not entirely clear from the illustration of the Haldas figurine whether the eyes are executed in the manner described above, or by three distinct punctations, one circular and two distinctly triangular. The description of the figurine in the recent Japanese publication states that the latter method was used (Tokyo Scientific Expedition, 1960, p. 446), but careful examination of the photograph by Lanning and by me suggested to both of us that the procedure was that used on the Tres Zapotes figurines. There is no question that this Mesoamerican pattern of rendering eyes was used on the figurine illustrated by Engel from the very lowest levels of Curayacu (Engel, 1956, fig. 9; Lanning, 1960b, p. 189), and in the figurines from Ancon illustrated by Uhle (1906, fig. 575) and Strong (1925, pl. 48,g). One of these I examined at the University of California Museum of Anthropology; here the use of an initial curved line with terminal pits followed by a central, deep punctation forming both the pupil and a ring of clay circling the pupil is clearly evident (U.C.M.A. 6341). The treatment of the nose on the Haldas figurine and certain of the Curayacu and Ancon figurines is quite similar to the early Tres Zapotes form, but in no case do the early Peruvian figurines show the deep pits at the outer corners of the mouth which are typical of the early Tres Zapotes style. It would be an exaggeration to state that these Peruvian solid figurines could be lost in any existing collection of Mesoamerican

MESOAMERICA & ANDEAN AREAS

specimens. Nonetheless, a general "Gestalt" of similarity is reinforced by certain specific and detailed analogies in manufacture which, to me, strongly suggest historicogenetic connections. What is evident is that these early Peruvian specimens are far closer to Mesoamerican examples than they are to any of the other early groups of figurines in South America. There is no resemblance to the Valdivia figurines (Evans, Meggers, and Estrada, 1959; Estrada, 1958; Zevallos and Holm, 1960), to the Chorrera figurines (Evans and Meggers, 1957, fig. 3,k), or to the Momil I figurines (G. and A. Reichel-Dolmatoff, 1956, pl. 22). On the basis of the evidence from Curayacu and Haldas this style of figurine is definitely established as pre-Chavin by a considerable margin of time; indeed, at Haldas the figurine is associated with some of the very earliest Initial period pottery (Lanning, personal communication). An apparently similar figurine so far not illustrated in publication was associated with the pre-Chavin pottery excavated by Bird near Huaca Prieta in the Chicama Valley. Bird has recently stated in personal communication that its eye treatment is like that of the Haldas specimen. Other examples, especially those from Ancon, appear to date from the maximum period of expansion of Chavin influence (Lanning, personal communication), but this point is less securely established, and they could possibly be earlier. Another trait which is widespread in early times in Mesoamerica and in northern South America is found with the Initial period pottery of Chicama Valley. This is the clay roller stamp (Bird, 1948, p. 27). On the basis of present evidence at least one trait, maize, of the Initial period in Peru, long predating the Chavin horizon, is certainly of ultimate Mesoamerican derivation. Other traits, especially the extensive use of large neckless ollas with internally thickened rims and solid, handmade figurines, have their closest similarities in

Mesoamerica, and one is left with the suspicion that a Mesoamerican origin for these early Peruvian traits is a distinct possibility. If all these traits appeared together in Peru at exactly the same time, the case for a site-unit intrusion of early Mesoamerican "Formative" culture would be a strong one, and such a group migration would come close to fulfilling the characteristics imagined by Spinden for the early spread of seed-crop agriculture in the New World. It has already been mentioned that these traits do not appear at exactly the same time in any sites so far excavated in Peru, pottery appearing first on the north coast, maize on the north-central coast. Lanning (personal communication) estimates that a hundred years or so might separate the appearance of maize and the appearance of pottery at Culebras. At present one could interpret these data in terms of repeated contact between Mesoamerica and Initial period Peru with the spread being due to long-distance communication, possibly by sea, or one could hypothesize that there are as yet undiscovered early sites in Peru where these traits do form a valid complex. Such a site could represent a true site-unit intrusion of Mesoamerican culture. The Mesoamerican source of such a migration would have to be a culture somewhat older and simpler than those so far described and showing the traits shared by such complexes as Chiapa de Corzo I, La Venta, and Ocos in simple, basic forms. PERUVIAN-MESOAMERICAN

CONNECTIONS

ON THE CHAVIN HORIZON

The question of connections between the Chavin horizon cultures of Peru on the one hand and such Mesoamerican sites as La Venta and Tlatilco on the other has received considerable attention in recent years. Μ. Ν. Porter's discussion (1953) of ceramic similarities between the two areas brings out many of the most important points, which there is no need to repeat in full. There are a number of traits of shape and decoration 271

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

which seem to form a sharp horizon in Peru, falling somewhere between 1000 and 600 B.C. (Lanning, 1960b). These traits also occur frequently at Tlatilco and as a smaller element in ceramic complexes from elsewhere in Mesoamerica. Among the more frequently cited traits are effigy vessels, stirrup-spout vessels, flat-bottomed, straight-sided bowls, vessels in the form of split masks emphasizing the idea of dualism. Such vessels are most frequently executed in dark smudged wares; the standard decorative technique involves the use of broadline incision separating the vessel surface into positive design areas typically highly burnished, and negative design areas typically dulled or roughened with a variety of techniques including textile impressing, rocker - stamping, nonrocked rouletting, scraping and brushing, punctating, or even the application of small pellets. In both Mesoamerica and Peru these ceramic elaborations appear mainly on the finer ceramics and are more likely to turn up in grave offerings than in midden deposits. The work at Ancon and Curayacu indicates that all these traits appear together in a tight-knit complex, as apparently at Tlatilco (Piña Chan, 1958). This suggests that the total complex was diffused from one of the two areas involved to the other in toto, and relatively rapidly. Some of these characteristics are present in the Ocos complex as defined by Coe, suggesting the possibility that it too falls in the 1000-800 B.c. range. If these similarities are significant, then the 1500 B.C estimate of Coe for the age of Ocos seems excessive (M. D. Coe, 1960b). Coe has recently suggested that the occurrence of stirrup spouts in the Valdivia complex in Ecuador indicates that at least one of these traits was earlier in Ecuador than in either Mesoamerica or Peru, whereas certain others of them almost certainly were developed in Mesoamerica (M. D. Coe, 1960b). I would suggest that until 272

stirrup spouts are found in fully documented cuts of Valdivia midden rather than from surface collections off sites containing Valdivia material this particular question should be allowed to remain open. At present there is no airtight argument which would indicate in which direction this complex of traits diffused, but the majority of evidence favors a movement from north to south. It cannot be emphasized too strongly that these elements are not part of the initial wave of Mesoamerican influence in Peru, and date around 600 years after that event. With the information at hand it is difficult to reconstruct the mode by which this complex of ceramic traits was spread. Bird (1948, p. 27), viewing the Chicama evidence only, suggests that the appearance of these traits is due to a migration of people into the area. Lanning (1960b, p. 570), viewing the evidence from sites somewhat farther south, suggests that no major movements of people were involved in this area, and that a relatively small number of religious leaders and a considerable amount of trade could explain the dispersal of these traits within Peru. Newman (1958) has recently pointed out that certain data of human genetics suggest sizable migrations out of Mesoamerica into the central Andes during the "Formative" stage, but such a movement of peoples could just as well be associated with the Initial period contacts discussed above. I would concur with Lanning's judgment on the intra-Peruvian spread of these elements, and it seems to me that their spread from Mesoamerica to Peru was most likely associated with the spread of religious ideas and a relatively small number of people rather than with a major movement of people. It is worthy of note that these ceramic elements are in both areas associated with the first evidence of major religious movements and the wide diffusion of religiously oriented art styles. In Peru the association is with the Chavin style and in Mesoamer-

MESOAMERICA & ANDEAN AREAS

ica with the La Venta-"Olmec" style. Certain people have seen close connections between these two major art styles. If one emphasizes subject matter, one can make a case for a close relationship (della Santa, 1959). The emphasis on large cats, raptorial birds, snakes, and fishes in both styles is a pronounced and diagnostic feature. If one compares the two styles as styles, it is hard to imagine two apparently contemporaneous groups of people treating the same subject matter in such completely dissimilar ways. It would be very hard to imagine either style evolving into something similar to the other in a matter of several hundred years, and each style is an apparently independent development peculiar to its own area. The Chavin style is an extremely complex manifestation of some duration. Any future attempts to derive Chavin style from Mesoamerica, or more specifically from La Venta style, must take cognizance of the refined stylistic and chronological analyses of Chavin style which have recently appeared (Rowe, 1962; Menzel, Rowe, and Dawson, 1964). Claims have been made concerning the influence of Chavin art on Mesoamerican sculpture or of some Mesoamerican style of this general period on Chavin art. These claims do not hold up under detailed scrutiny. The carved stone tablet from Guerrero which Covarrubias regarded as Chavinoid might possibly show some significant parallels with central Andean sculpture, but the more profitable comparisons would be with Pucara, Tiahuanaco, or Callejon de Huayles (Covarrubias, 1957, fig. 50). I am unable to find one specifically Chavinoid characteristic in this piece. I have heard that various people have commented on a general similarity between the not specifically Chavinoid, but still early, stone carvings of Cerro Sechin in the Casma Valley, Peru, and the sub-"Olmec" danzantes carvings at Monte Alban. On analysis, however, the general feeling of similarity

seems to relate to a general simplicity and a similar technique of execution rather than to any close stylistic affinities. If I had to select any group of Mesoamerican carvings which looked vaguely Chavinoid, it would be a couple of the altars at the site of Tonala in Chiapas (Ferdon, 1953, pls. 21, 22). Here the treatment of fangs, eyes, and eyebrows shows certain similarities to Chavin practice. I would hesitate to suggest contact on the basis of such slight evidence, however. ALLEGED EARLY MESOAMERICAN INFLUENCES IN THE PERUVIAN MONTAÑA

A group of ceramics known as Early Tutishcainyo are the oldest so far discovered in tropical forest Peru, and quite possibly in the whole Amazon Basin. If Early Tutishcainyo is reduced to a series of discreet traits rather than viewed as an integrated ceramic complex, one can present a list of ceramic elements which are more common in Mesoamerica viewed as a whole than in South America. Most prominent among these are broad labial flanges, grooving associated with broad labial flanges, elaborate composite silhouettes, exaggerated basal flanges, zoned crosshatched decoration, and the use of dry pigments in incision. On the basis of such a list one might conclude that Early Tutishcainyo represents a major focus of Mesoamerican influence (Lathrap, 1958). This position seems less likely when one observes that the closest Mesoamerican similarities in rim form are to be found in Chicanei, the late Formative of Copan, and complexes containing considerable Usulutan ware. The closest similarities in vessel profile are to Early Zacatenco, the closest similarities in basal flange are to Tzakol. To make Mesoamerican comparisons one must draw on complexes separated by as much as 1500 years. There is no known Mesoamerican complex in which the "Mesoamerican" traits found in Early Tutishcainyo occur together in the configuration characteristic of Early Tut273

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

ishcainyo. This particular case is especially instructive as to the dangers of discussing long-distance cultural connections without regard to strict chronological control at both ends. My present position on Early Tutishcainyo is that it is distantly allied with the earliest lowland ceramics of northern South America, specifically of Venezuela and Colombia. There is now substantial evidence, including radiocarbon dates, indicating that Early Tutishcainyo is far earlier than any Mesoamerican complexes showing the characteristics discussed. MISCELLANEOUS CERAMIC TRAITS SHARED BY MESOAMERICA AND PERU

Willey (1955a) sketched the distribution and probable direction of spread of several ceramic elements, among which were annular-based vessels, poly-pod vessels, and resist-decorated ware. On the basis of evidence now available his conclusions do not hold. Annular bases and vessel feet are now known to occur in the Initial period pottery of the central and south coast of Peru, so that a spread from Mesoamerica to Peru in relatively late times now appears impossible (Lanning, 1960b, pp. 78, 563). Both these traits appear early in lowland ceramics of northern South America, and I suspect have their source in this area rather than in Mesoamerica (G. and A. Reichel-Dolmatoff, 1956; Cruxent and Rouse, 1958-59). A spread into coastal Peru from the Montaña would fit better with their spotty distribution in this region than would a gradual spread from north to south along the coast or the Andes. At present the situation with regard to resist decoration appears extremely complex. In the coastal regions of Peru evidence is clear that the spread of this trait was from south to north. The earliest known examples are on the south coast, but there are other instances in Nuclear America, Momil I for example, where the trait may have equal or greater antiquity (Lanning, 274

1960b, p. 563; G. and A. Reichel-Dolmatoff, 1956, pp. 148-50). EVIDENCE OF CONTACT BETWEEN PERU AND MESOAMERICA IN THE CLASSIC AND POSTCLASSIC STAGES

The evidence cited by Willey indicates that Nuclear America continued to be a diffusion sphere in late precontact times. The available information becomes even more difficult to handle. At this time we appear to be dealing with the diffusion of individual disarticulated traits such as might be caused by trade. Evidence suggesting major movements of people between Mesoamerica and Peru, or suggesting the diffusion of total complexes or religious systems, is lacking. M. D. Coe (1960b) indicates that the sea was a major avenue of contact at relatively early times; West (1961) has documented the view that long-distance sea trade was normal and extensive at the time of the conquest. At the present time one must discuss the effects of such trade in terms of all Nuclear America, since the specific routes involved and the major centers for trading expeditions have not been established. That metallurgy in general spread from south to north in Nuclear America seems established, but the point or points from which it entered Mesoamerica is less clear (Willey, 1955a, p. 41). The piecemeal nature of this cultural transmission makes it difficult to evaluate similarities as definite evidence for cultural contact. Nonetheless, one gets the impression that the artisans of Mesoamerica in late times were frequently aware of the products of contemporary technology in Peru, and that the reverse situation was also true. Willey (1955a, p. 39) has discussed some possible examples, emphasizing particularly the use of molds in pottery manufacture. Covarrubias has suggested that certain of the Cholula Polychrome of early Postclassic Mexico shows specific influence of central or south coast Peruvian design (Covarrubias, 1957, plate

MESOAMERICA & ANDEAN AREAS

facing p. 196). The example he illustrates strikes me as quite convincing. Some of the polychrome pottery of northwestern Mexico, especially the early Culiacan series, could be interpreted as showing influence of coastal Peruvian ceramics and/or textile design of a time immediately predating the Tiahuanaco horizon. The dating would be compatible with such influence, but in no case are the similarities so specific or detailed as to make the case airtight. Certain of the Culiacan step designs, face designs, and the tendency of the Culiacan artist to square off the end of clusters of plumes have a distinctive coastal Peruvian look about them, and the Culiacan use of the step-fret design—what Kelly has called the "saw" design—is quite close to the step-fret design in coastal Peruvian tapestries of this time (Kelly, 1945, fig. 19). It has also been noted that certain practices in stone carving recently noted in the Mexican Plateau have close analogies in late Peruvian stonework (Wicke and Bullington, 1960). Comparisons have been made frequently between the cut stone mosaics of Mitla and Yagul and the cut clay mosaics of Chan Chan (Willey, 1955a, p. 43). With the possible exception of the Peruvian designs in Cholula Polychrome, none of these similarities are as convincing evidence for direct cultural contact as the similarities already discussed for earlier time periods. SUMMARY

The preceding discussion suggests chiefly that our data are still insufficient to discuss meaningfully the more interesting problems

concerning Mesoamerican-Peruvian contacts. One gets the impression that during the earlier part of the time span considered the major direction of influence was from Mesoamerica to Peru. This would seem to be related to the great significance of the introduction of seed-crop agriculture into the coastal areas of Peru. At later times there may have been more influence running the opposite direction, possibly a result of the more rapid and spectacular rise in technology in the central Andes. During the earlier periods, fairly elaborate cultural complexes appear to have been transmitted, more or less in toto, and there are possibilities of one or more site-unit intrusions from Mesoamerica into Peru. During later times, the influences transmitted seem to have been largely in terms of individual elements and traits, suggesting commerce but no major political or ideological impingement. With a reasonable amount of further work in Mesoamerica and Peru it should be possible to prove or disprove the hypotheses offered concerning contact on the "Formative" level, but the piecemeal nature of later cultural transmission will continue to make the certain identification of later cultural contacts between the two areas a difficult problem.6

6 This article was submitted July, 1961, since which time most of the significant literature on this subject has appeared, specifically: Izumi and Sono, 1963; Μ. D. Coe, 1962b, 1963; Lanning, 1961, 1963; Meggers and Evans, 1964; Heiser, 1965; Lathrap, 1963, 1965. Intensive programs of excavation are continuing at Las Haldas and Kotosh.

275

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

REFERENCES Bennett, 1948 Bird, 1948 Childe, 1942 Coe, M. D., 1959, 1960b, 1962b, 1963 Collier, 1959 Covarrubias, 1957 Cruxent and Rouse, 1958-59 Cutler and Whitaker, 1961 della Santa, 1959 Dixon, 1959 Drucker, 1943a, 1952 Edmonson, 1961 Engel, 1956, 1957a, 1957b Estrada, 1958 Evans and Meggers, 1957 , , and Estrada, 1959 Ferdon, 1953 Grobman et al., 1961 Heiser, 1965 Izumi and Sono, 1963 Kelly, 1945 Lanning, 1960a, 1960b, 1961, 1963 Lathrap, 1958, 1960, 1962, 1963, 1965 MacNeish, 1958, 1961, 1962

276

Mangelsdorf, 1954 , MacNeish, and Galinat, 1964 Meggers and Evans, 1955, 1964 Menzel, Rowe, and Dawson, 1964 Newman, 1958 Piña Chan, 1958 Porter, Μ. Ν., 1953 Reichel-Dolmatoff, 1955, 1965 and Reichel-Dolmatoff, 1956 Rowe, 1960, 1962 Spinden, 1917b Strong, 1925 and Evans, 1952 Thompson, R. H., 1958a Tokyo Scientific Expedition, 1960 Uhle, 1906 Wallace, 1960 West, 1961 Whitaker and Bird, 1949 Wicke and Bullington, 1960 Willey, 1955a and Corbett, 1954 and McGimsey, 1954 Zevallos Menéndez and Holm, 1960

14. The Problem of Transpacific Influences in Mesoamerica

ROBERT

traits but with archaeological comTplexes delimited in area and period, HIS ARTICLE1

deals not with isolated

in the Old World as well as in the New. Into this chronological frame nonmaterial traits may eventually be fitted. Only those apparently interconnected transpacific movements which were first begun by the coastal peoples of China and later continued by Indian and southeast Asiatic mariners will be considered. Metallurgy, for instance, although certainly of Old World origin (Heine-Geldern, 1954), will not be discussed because it came to the Maya and Mexicans indirectly by way of South America. MARBLE VASES FROM THE ULUA VALLEY AND THE T A J I N STYLE OF MEXICO

The comparisons here should be viewed against the background of what seems to have happened in the western Pacific between approximately 700 and 200 B.C. This 1 This article is based on the results of research sponsored by the Wenner-Gren Foundation for Anthropological Research.

HEINE-GELDERN

must have been a time of great maritime activity on the part of the peoples of eastern Indochina and of southern and central China. Influences of the bearers of the Dongson culture of Tonkin and Annam spread over vast regions of Indonesia and along the northern coast of New Guinea as far as the Solomon Islands. Although it is generally recognized that the art styles of most of the areas concerned perpetuate in modified forms that of the Dongson culture, those of central Flores and of a large part of Borneo are doubtlessly derived from China. This applies also to the Trobriand Islands and even more so to New Zealand. Not only the ornamental designs but also the stoneworking techniques and the stone clubs of the Maori are unknown elsewhere in Polynesia, but are typically East Asiatic and more specially Chinese (Heine-Geldern, 1937, pp. 200-04; Ling, 1956). Whether these traits were introduced directly into New Zealand in the Late Chou period, prior to 200 B.C., or later by way of some intermediate area, does not concern us here. The important point is that whatever 277

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

FIG. 1—MAYA AND CHINESE DESIGNS, a, Design on a marble vase from the Ulua Valley, Honduras (Stone, 1938). b, Stone tablets from Lo-yang, China (White, 1934). c, Jade ornament, China (Salmony, 1938).

transpacific voyages may have been undertaken by the peoples of coastal China, of the ancient kingdoms of Wu and Yueh, they were not isolated events but part of a vast maritime expansion which extended over the western Pacific no less than toward the east. 278

The "Chinese" aspect of the designs on the marble vases from the Ulua Valley has been noticed and commented on by several scholars. The stylized animal heads on the vases are built up of spirals, the upper and lower jaws, cheek, back of the head, and tongue being formed by one spiral each (fig. 1,α). They correspond, almost element by element, to similar animal heads on small stone tablets and on a jade ornament of Late Chou China (fig. l,b,c). The scale bands on the upper rims of Ulua vases are identical in form with those on bronze vessels of the Middle Chou period (fig. 2). The Chinese character of the designs of the Ulua vases is so conspicuous that, had these been found not in America but somewhere in Asia, no one would doubt that they represent a colonial version of Chinese art. Even more important in this respect are the ornamental designs of the Tajin style of Mexico. The scroll patterns on several palmas (fig. 3,a) resemble the Late Chou pattern of "teeming hooks and scrolls" (fig. 3,b) to such an extent that, were they illustrated alone, without the objects on which they are found, it would be difficult to recognize that they are not Chinese. The scrolls show one of the most characteristic traits of Tajin ornamental design, the double contour. This, too, is found in China, as well as in those art styles of Indonesia (central Borneo, central Flores) in which elements of Late Chou art still survive. The comparison of the interlaced ribbons characteristic of the Tajin style with the interlaced dragons of Late Chou art might seem arbitrary, did not the winglike designs on a frieze at Tajin correspond exactly, in shape and position, to the dragon's wings on a Chinese bronze vessel (fig. 4). This sheds light on the origin of numerous Tajin designs consisting of interlaced bands with hooklike projections—remnants of the wings —which are usually interpreted as stylized representations of mythical serpents. In China, too, there was a tendency for the

TRANSPACIFIC INFLUENCES IN MESOAMERICA

FIG. 3—SCROLL PATTERNS, a, Palma (Spinden, 1933). b, Bronze fragment, China, Late Chou period (Umehara, 1936).

FIG. 2—SCALE BANDS, a, Marble vase from the Ulua Valley, Honduras (Peabody Museum, Harvard University), b, Bronze tripod, China, Middle Chou style.

motif of interlaced dragons to disintegrate into mere ribbons, with only the sickleshaped wings and jaws of the dragons surviving and indicating the original design (fig. 5,a). The comparable design on the back of a pyrite mirror from Kaminaljuyu, with its interlaced bands and hooks, would seem due to its being only one step further along in stylistic disintegration (fig. 5,b). The winged dragon itself was not completely unknown in Mexico. A palma in the American Museum of Natural History (fig. 5,d) shows two interlaced serpents with hook-shaped appendices projecting from their bodies. Even though the Mexican artist obviously misunderstood and garbled the original design, its derivation from the interlaced dragons with sickle-shaped wings of Chinese art (fig. 5,c) is unmistakable.

FIG. 4—CHINESE AND MEXICAN DESIGNS. a, Bronze vase, China, Late Chou period, b, Stone frieze from Tajin, Mexico. (Covarrubias, 1947.)

One of the most characteristic motifs of a certain group of Late Chou bronzes consists in dragons with ornamental spiral and double-spiral designs on their bodies (fig. 6,a). The same motif occurs on a Mexican palma (fig. 6,b). Significantly, it occurs also in New Zealand, the art of which was derived from that of Late Chou China (fig. 6,c). In the New Zealand carving even the original Chinese dragon's sickle-shaped wing is still clearly recognizable. The scrolls on an Olmec-style stone mask in the Peabody Museum of Harvard University obviously represent tattooing (fig. 7,a). The spirals on forehead, cheeks, nose, and chin are closely matched by those of 279

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

FIG. 5—STYLIZED W I N G E D DRAGON, a, Bronze object, China, Late Chou period (Umehara, 1936). b, Back of a pyrite mirror, Kaminaljuyu, Guatemala (Kidder, Jennings, and Shook, 1946). c, Chinese bronze vase, Late Chou period (Vessberg, 1937). d, Palma (American Museum of Natural History, drawing by M. Covarrubias).

Maori tattoo designs (fig. 7,b). According to a Chinese source, certain ancient tribes of southern China used to tattoo their foreheads with "circles," which probably means spirals. Several Late Chou bronze figures of demons, inlaid with silver, show spiral designs, no doubt representing tattooing, on forehead, cheeks, and body (fig. 7,c). We thus have here another instance of the same motif having spread from China to New 280

Zealand on the one hand and to Mesoamerica on the other. The disc-shaped pyrite mirrors of Mexico and Guatemala suggest, of course, some connection with the disc-shaped bronze mirrors of China. The Chinese character of the design on the back of a mirror from Kaminaljuyu (fig. 5,b) has been pointed out above. A Mexican mirror with a Tajin-style design on its back has a circle of triangular perforations running parallel to its rim (fig. 8,a). A similar circle of triangular perforations is found on a bronze mirror from China or from the Chinese-Mongolian borderland (fig. 8,b). The correspondences listed here, and other, no less striking ones which must be omitted for lack of space, are too numerous and too specific to be due to mere chance. It is no exaggeration to say that the bulk of purely ornamental Tajin designs seems to be of Chinese origin. The chronological problem must, of course, be given due attention. The scale band, which occurs not only on the Ulua vases but also on several palmas (figs. 2,a; 9,a) is a motif of the Middle Chou style of China (9th to 8th centuries B.C.), although it survived to some extent into the Late Chou period. It indicates a relatively early beginning of Chinese influence in Mesoamerica, possibly as early as around 700 B.C. This is significant since in Peru, too, Chinese influence appeared around this date (Heine-Geldern, 1954, pp. 383-86; 1959c). Most of the correspondences, however, suggest relations with China between approximately 600 and 200 B.C. In the past, the main objection raised against the assumption of Chinese influence on Mexican art was based on the alleged time gap between the end of the Late Chou style of China, about 200 B.C, and the appearance of the Tajin style. This objection is no longer valid. What counts is not the date of the temples of Tajin nor that of the yokes and palmas, but that of the oldest work of art in the Tajin style. This is a

FIG. 6—DRAGON SPIRAL DESIGNS, a, Chinese bronze, Late Chou period (Umehara, 1936). b, Palma, Mexico (Fewkes, 1907). c, Woodcarving, New Zealand (Hamilton, 1901).

FIG. 7—SPIRAL TATTOOING, a, Stone mask, Mexico (Peabody Museum, Harvard University), b, Wooden head, New Zealand (Rautenstrauch-Joest-Museum fiir Volkerkunde, Cologne), c, Bronze figure with silver inlay, China, Late Chou period (Minneapolis Institute of Fine Art).

FIG. 8—MIRRORS W I T H RIM PERFORATIONS, a, Pyrite mirror, Queretaro (American Museum of Natural History), b, Bronze mirror, China or Inner Mongolia (Andersson, 1932).

FIG. 9—TAJIN STYLE. Palma, Mexico (Fewkes, 1907). b, Wall painting, Teotihuacan II (Gamio, 1922).

281

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

FIG. 10—POTTERY VESSELS, α-c, Chinese, Han period, northern Vietnam (Janse, 1947). d-ί, Teotihuacan III style, Kaminaljuyu, Guatemala (Kidder, Jennings, and Shook, 1946). j'-l, Funeral pottery, China, Han period (Laufer, 1909).

wall painting on the ground floor of a Teotihuacan II building, the upper part of which was pulled down during the city's third period (fig. 9,b). Two discordant runs of carbon-14 counts for material from Atetelco which contained Teotihuacan II pottery yielded the dates of 294 B.C ± 180 years and 72 A.D. ± 200 years (Libby, 1955, p. 129; Stuiver, Deevey, and Gralenski, 1960, p. 57). We may discard the first of these counts as unreliable. If we take the margin of error into account, the second date indicates the possibility that Teotihuacan II may have been in existence as early as the latter half of the 2nd century B.C. The fact that at Atetelco Teotihuacan III ceramics were found at all levels, whereas Teotihuacan II pottery predominates only in the lowest level (Séjourné, 195657), proves that the date for Atetelco refers only to the very end of the Teotihuacan II period. The supposed chronological gap between the Late Chou and Tajin styles is thus practically closed. The appearance of a Tajin-style paint282

ing at Teotihuacan at a date long before the mural decorations of the temples of Tajin and long before the palmas and yokes calls for an explanation. We know the ornamental art of the Late Chou period of China mainly through luxury objects of metal and jade. It must have been primarily an art of woodcarvers and painters, particularly as far as the decoration of buildings was concerned. It was in woodcarving and painting that the Late Chou style was introduced into Borneo, Flores, the Trobriand Islands, and New Zealand. The same must have been true in Mesoamerica. The numerous wooden objects of the Aztec period, as well as the wooden lintels and reliefs from Tikal and Chichen Itza, show how important a role woodcarving played in ancient Mesoamerican art. It is, of course, not surprising that no woodwork of the older periods has survived. All the Tajinoid traits, such as scrolls and double contours, found in Mexico (Teotihuacan III, Monte Alban III, Xochicalco) and in practically the whole Maya area suggest that what, for the lack

TRANSPACIFIC INFLUENCES IN MESOAMERICA

FIG. 11—LOTUS MOTIF. a,e, Amaravati, India, 2nd century A.D. (Coomaraswamy, 1928-31). b-d, Chichen Itza, Yucatan (Maudslay, 1889-1902). f , g , Details of frieze in lower inner temple, Pyramid of the Magician, Uxmal, Yucatan (Uxmal, Official Guide of the Instituto Nacional de Antropología e Historia).

of a better term, we call the Tajin style was originally an early horizon style, at first used only in woodcarving and painting; that it once covered the whole area from Teotihuacan and Veracruz to Guatemala and western Honduras; and that later its designs were reproduced in stone, mainly but by no means exclusively, in the Veracruz region. POTTERY OF THE TEOTIHUACAN AND HAN PERIODS

The comparisons shown here in figure 10 need hardly any comment (Heine-Geldern, 1959b). The correspondences between

Teotihuacan III and Chinese pottery of the Han period are by no means restricted to mere forms. In Teotihuacan III new ceramic techniques and principles appear which are precisely characteristic of Han pottery (Ekholm, 1964a). The carbon-14 date for Atetelco, 72 A.D. ± 200 years, marks not only the approximate date of the end of Teotihuacan II, but at the same time that of the beginning of Teotihuacan III. A carbon-14 date from Tlamimilolpan, 236 A.D. ± 65 years (Linné, 1956) brings Teotihuacan III down to the end of the Han dynasty. All this cannot be due to mere chance. Apparently the transpacific voyages 283

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

from China which had started around the beginning of the Late Chou period were continued into Han times, 202 B.C. to A.D. 220. Later than this there seems to be no trace of Chinese influence in Mesoamerica. The end of the Chinese transoceanic voyages may have been brought about by the long period of political troubles and internal wars which eventually resulted in the downfall of the Han dynasty. MESOAMERICA AND THE HINDU-BUDDHIST CULTURES OF INDIA AND SOUTHEAST ASIA

Around the middle of the 1st century A.D. Indian merchants were sailing to lower Burma and to Sumatra. This was the beginning of a truly explosive movement. In the 2nd century the coasts of Further India were already studded with Indian colonial cities, and Hindu kingdoms had been founded as far east as Cambodia and what now is Vietnam. It seems highly improbable that the Indians should have independently rediscovered America. In Indonesia and on the coasts of Indochina they inevitably came into contact with Chinese sailors and merchants. This was the time of the later Han dynasty, when Chinese transpacific voyages either had not yet been discontinued or must at least have still been a living memory. It was no doubt from the Chinese that the Indians heard of the land on the other side of the ocean and of the route thither. Therefore it is probable that Indian sailors and adventurers began to cross the Pacific in the 2nd century A.D., when the initial vigour of the eastward movement had not yet spent itself. At a later date these voyages must have been continued by the Hinduized peoples of southeast Asia, particularly the Khmer and the Cham. Archaeological indications of these contacts abound, but only a limited number can be discussed here. Ever since the 2nd century B.C. the lotus motif played an important role in the art 284

of India. As applied on architraves and in border designs in the early period, 2nd century B.C. to 2nd century A.D., it shows not only the flowers and leaves but the whole plant, including the rhizome, which in nature is buried in the mud at the bottom of ponds and therefore is normally not seen. Although the flowers and leaves are frequently represented in a naturalistic manner, the rhizome is always transformed into a purely decorative undulating creeper (fig. 11,α). In Maya art the lotus motif, if we may use this term for the sake of convenience, is found at Palenque and at various other sites (Rands, 1953). But it is in the art of the Toltec period at Chichen Itza, in the Temple of the Jaguars, that it corresponds most closely to Indian designs, particularly to those of Amaravati in south India (2nd century A.D.) (fig. 11,b,c). In the early Buddhist art of India the lotus is occasionally interspersed with human figures. The same is true at Chichen Itza. The similarity is particularly striking in figures shown in reclining position and holding on to the stem of the rhizome, a part of the plant which usually is inaccessible (fig. ll,c,d,e). By no stretch of imagination could we assume that the south Indian and Maya artists, independently from one another, had reproduced scenes they had actually seen. It has been objected that the difference in date between Amaravati and the Temple of the Jaguars, almost a millennium, makes it appear exceedingly improbable that there might have been any connection between them. However, Alberto Ruz Lhuillier discovered reliefs with the motif of the reclining person holding the stem of the undulating lotus rhizome also in the Lower Inner Temple of the Pyramid of the Magician at Uxmal, where it dates from about the 7th century A.D. (fig. 11). This cuts the time gap between Amaravati and Chichen Itza in two. Obviously it is only due to the vagaries of archaeological discovery and to

TRANSPACIFIC INFLUENCES IN MESOAMERICA

FIG. 12—MONSTER HEAD, a, Lintel, Lolei, Cambodia, 9th century A.D. (P. Stern, 1934). b, Monster head combined with two makara heads, Mison, Champa, 10th century A.D. (Naudou, 1961). c, Combination of three heads, Chichen Itza, Yucatan (Maudslay, 1889-1902).

the role woodcarving must have played in Maya art that so far only this one intermediate link has been found. In view of what has been said above concerning the probable date of the first Indian transpacific voyages, it is quite conceivable that the lotus motif was introduced into America directly from India as early as the 2nd century A.D. On the other hand, we know practically nothing of the art of the first half-millennium of Indian colonial cultures in southeast Asia. Its works, probably predominantly of wood, have perished. We cannot, therefore, discount the possibility that the Amaravati version of the lotus motif was first introduced into southeast Asia and from there reached America at some later date. Of course, its appearance at Uxmal provides a terminus ante quern, about A.D. 600. In the Temple of the Jaguars at Chichen Itza some lotus rhizomes emerge from the

corners of the mouth of a monster mask without lower jaw (fig. 11,b). This, too, is a motif of Hindu-Buddhist art. The monster mask of India was originally meant to represent a lion's face. It appeared first in the 5th century A.D. and soon gained in popularity. In Java it was one of the most important motifs, beginning with the oldest surviving monuments, of the 8th century A.D. On Cambodian lintels of the 9th century the garland of baroque ornamental leaves, into which by that time the original lotus design had developed, emerges from the corners of the monster's mouth (fig. 12,a). The motif is so purely conventional that it would be difficult to imagine its having been invented independently twice. In the Old World we can trace its gradual development (Gangoly, 1920; Vogler, 1949), but not in America. It is unlikely, however, that the motif came to America prior to the 9th century. The garland must have been replaced there 285

FIG. 13—MAKARA. a, Java, 9th century A.D. b, Makara-like monster on wooden lintel, Chichen Itza, Yucatan (Maudslay, 1889-1902). c, Lintel, Sambor Prei Kuk, Cambodia, 7th century A.D. (Marchai, 1951). d, Sarnath, India (Vogler, 1949). e, Champa (Vogler, 1949). f, Head of makara-like monster, Copan, Honduras (Maudslay, 1889-1902).

FIG. 14—ELEPHANT HEAD, a, Maya rain god, Codex Madrid, b, Statue of the God Ganesa, Mison, Champa, 9th to 10th century A.D. (Parmentier, 1922). c, Statue, San Salvador (Balfour Gourlay, 1940). d, Relief, Oaxaca, probably Monte Alban III period (Museum für Võlkerkunde, Berlin; courtesy, W. Krickeberg).

TRANSPACIFIC INFLUENCES IN MESOAMERICA

FIG. 15—COSMIC TREE, a, Relief in the Temple of the Foliated Cross, Palenque, Chiapas (Maudslay, 1889-1902). b, Wayang figure, Java (Museum für Volkerkunde, Vienna).

by the locally familiar original lotus design, which apparently had survived in Maya artistic tradition since the 6th or, possibly, even the 2nd century A.D. As Jean Naudou (1961) pointed out, the lotus frieze of Chichen Itza contains still a third motif of Hindu-Buddhist origin. In the medieval art of the Dekkan and in that of the Cham of Indochina the monster mask without lower jaw, seen en face, is frequently combined with two makara heads in profile (fig. 12,b). A similar design occurs at Chichen Itza. A monster mask without lower jaw is flanked by two animal heads in profile, which with their open mouths, the teeth in their upper jaws, and their forward-curved trunks correspond closely to the makaras (fig. 12,c). The makara, a fantastic sea monster combining in variable proportions traits of the crocodile, fish, elephant, and Hellenistic

dolphin, is one of the main ornamental and symbolic motifs of Hindu-Buddhist art. It was widely used in architectural decoration and in that of thrones, weapons, and personal ornaments. The history of the development of its numerous variants is fairly well known (Vogel, 1929-30; Vogler, 1949, pp. 68-96). In Maya art a certain class of socalled serpents, easily distinguishable from the representations of either real or mythical serpents, shares with the makara the wideopen mouth, the powerful teeth protruding from the upper jaw, and a trunk. The head of such a makara-like monster from a wooden lintel at Chichen Itza, for instance, may be compared with a Javanese makara head of the 9th century; its body and tail, which certainly are not serpentine, resemble most those of Cambodian makaras of the 7th century (fig. 13,a,b,c). In India, as well as in Indonesia, Cambodia, and 287

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

FIG. 16—THRONE SCENE, a, Relief from the Palace at Palenque, Chiapas (Museo Arqueológico Nacional, Madrid). h, Bronze figure of a Bodhisattva, Bihar, India, probably 8th century A.D. (Jayaswal, 1933).

Champa, a human figure is frequently shown emerging from the makara's mouth (fig. 13,d,e). Thus it is in Maya art (fig. 13,f) (Maudslay, 1889-1902, vol. 1, pls. 23, 24, 53, 116). Is it too bold to speak of a Mesoamerican version of the makara? Some of the representations of the Maya rain god in the codices share with the elephant not only the trunk but the very characteristic shape of the head with the depression over the eye (compare figs. 14,a and 14,b). This latter trait survives even in those versions in which the god's trunk has been replaced by an exaggeratedly long but still human nose. A stone statue from San Salvador shows a standing person with an elephant's head (fig. 14,c). The same motif, a human with an elephant's head but in sit288

ting position, occurs in a statue from the state of Veracruz (Nomland, 1932) and in a Zapotec relief from the state of Oaxaca (fig. 14,d). Figures of human persons with elephant heads in a country where at this time the elephant was unknown, and in surroundings in which indications of Hindu influence are fairly numerous, inevitably suggest that they are derived from the Indian god Ganesa (fig. 14,b). The fact that in India representations of Ganesa appear not prior to the 5th century A.D. provides a terminus post quern for the introduction of the motif into Mesoamerica. Ekholm (1953, pp. 76-77) was right in comparing a unique relief from Palenque, a man sitting on a throne and holding a lotus flower in his left hand (fig. 16,a), with

TRANSPACIFIC INFLUENCES I N MESOAMERICA

Bodhisattva images of the Pala period of northeastern India (fig. 16,b). The awkward manner in which the bent right leg is represented shows that the Maya artist was faced with an unusual problem with which he was not familiar. The relief suggests some kind of relation in the 8th century A.D.

The trees represented in the reliefs of the Temple of the Cross at Palenque resemble stylistically the cosmic tree as shown in the Javanese shadow play (fig. 15,a,b). Moreover, in both cases a demoniac face (in Java probably a sun symbol) is placed in the tree's branches. The fact that a highly schematized version of the motif occurs in the reliefs of Angkor Vat in Cambodia, 12th century A.D., shows that it must be of considerable antiquity in southeast Asia. Clay animal figures on wheels have been popular toys in India ever since the 3rd millennium B.C. (fig. 17,a). In Mexico wheeled toys of this kind have been found in different cultural surroundings, dating from different periods (fig. 17,b-e) (Ekholm, 1946b; Caso and others, 1946; Linné, 1951). Most writers on the subject have ascribed them to independent local invention. In view of all the other correspondences listed here it seems far more likely that the principle of wheeled toys was introduced from India. The parallels between Mesoamerican and Hindu-Buddhist atlantean figures and serpent bannisters deserve at least passing notice. Of far greater significance is the decoration of wall panels with rows of colonnettes in the Puuc style of Yucatan (fig. 18,a). They resemble the colonnettes found as window gratings in Khmer temples (fig. 18,b,c). In Cambodia these colonnettes appeared in the 9th century and were widely used during the Angkor period. They obviously were imitations of wooden ones turned on the lathe. A wooden colonnette of the same shape as those of stone was actually excavated at Angkor (Commaille, 1913, p. 14). In Laos, where the

FIG. 1 7 — W H E E L E D TOY ANIMALS O F CLAY. a, Taxila, India, 2nd century B.C. (Annual Report, Archaeological Survey of India, 1919-20). b-e, Mexico (Ekholm, 1946b).

tradition of wooden temple architecture survived until recent times, wooden colonnettes of the same shape and employed in the same manner as those of the classic Khmer monuments can still be seen (Parmentier, 1954, pp. 74-77). The Cambodian colonnettes had a real function, but those of Yucatan are mere wall decorations. However, a link is furnished by those instances in Cambodia in which the colonnettes were placed as grating within the frame of a false window, thus against the background of a wall (fig. 18,c). From this 289

FIG. 18—STONE C O L O N N E T T E S . α, Palace, Labna, Yucatan (ProskouriakoiF, 1946). b, As window grating, Prah Ko, Cambodia, 9th century A.D. (courtesy, Jean L a u r ) . c, In false window, Bakong, Cambodia, 9th century A.D. (courtesy, Jean L a u r ) .

TRANSPACIFIC INFLUENCES IN MESOAMERICA

there was only one step to the way the colonnettes are used in Puuc-style buildings. Furthermore, not only do we know that the Cambodian stone colonnettes were preceded by similar ones of wood, but they, too, were turned on the lathe. Otherwise their complicated forms and their perfect finish could never have been achieved. The lathe was not known in America. Yet, the shapes of most of the Puuc-style colonnettes, although simpler than the Cambodian ones, still allow us to recognize that they must be derived from lathe-turned prototypes. Therefore we are confronted here with what probably constitutes the most decisive proof of the Old World origin of an American architectural motif. Several authors have thought that the Mesoamerican stepped pyramid with a temple on its platform might be derived from more or less similar Cambodian structures. Chronologically this is impossible, for in Cambodia the stepped pyramid appeared only as late as the 9th century. Should there be any real relation in this case, the direction of influence must have been reversed. It is, of course, theoretically conceivable that Cambodians who had visited America and seen stepped pyramids introduced this architectural form into their homeland. With this brief survey, it should be emphasized that the examples listed do not by any means exhaust the similarities between Maya and Mexican art motifs on the one hand and those of India and Hindu-Buddhist southeast Asia on the other. Further comparisons may be found in Heine-Geldern and Ekholm, 1951, and in Ekholm, 1953. FURTHER ASIATIC TRAITS IN THE CULTURES OF MESOAMERICA

In an interesting paper Dennis Wing-sou Lou (1957) discussed a large number of parallels between Chinese and Mesoamerican religious concepts and rituals, some of them truly striking. His comparison of the role of the dragon in China with that of the

serpent in Mesoamerica will gain in significance if viewed in conjunction with what has been said above concerning the derivation of certain Tajin-style designs from Chinese representations of the winged dragon. In China the four thunder and rain gods were associated with the four quarters of the world, and special colors were assigned to them. This was exactly the same among the Maya, a duplication that can hardly be due to mere accident. Various Mesoamerican culture traits suggest Indian origin, for instance the game of patolli with its parallels to the Indian pachisi (Tylor, 1896). Of greater importance are the correspondences in cosmological ideas, particularly those concerning the world ages. According to Buddhist doctrine, the worlds, following one after the other, are annihilated by either fire, water, or wind (Pallegoix, 1854, pp. 421-31). In Mexican belief the successive "suns" or worlds were destroyed by fire, water, wind, jaguars, or earthquakes (W. Lehmann, 1938, pp. 61-63; León-Portilla, 1956, pp. 106-21). In numerous instances the decision between Chinese or Indian derivation is not easy. All ancient Asiatic civilizations were to some extent interconnected. Despite their differences they shared a large number of common traits. This makes it often difficult, even in southeast Asia, to determine whether a given trait had come from India or from China. This is true in Mesoamerica. Moreover, we must consider the possibility that different variants of the same trait may have been introduced at different times from both China and India. The calendar is a case in point. In view of the close correspondences between the highly artificial calendrical systems of Mesoamerica and Asia (India, China, southeast Asia)—correspondences in basic principles, as well as in very specific details—the Asiatic origin of the former can hardly be doubted (Humboldt, 1816, 1:38492, 2:1-99; Graebner, 1920-21; Rock, 1922; 291

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

Kirchhoff, 1964). Since calendrical dates appear as early as in the glyphs of Monte Alban I (Caso, 1947), the first stimuli in this respect can have come only from China. However, many traits indicate later modifications due to Indian or southeast Asiatic influence. The principle according to which special colors were attributed to the main directions of the world has been thoroughly studied by Nowotny (n.d.) He traced it from the ancient Near East to India, China, southeast Asia, and America and came to the conclusion that it reached America in several waves. Although the Mexican cosmic colors correspond in general to those of India, there are indications and remnants of older systems. Apparently in this case, too, Chinese influence preceded the Indian. Irrespective of the question of Chinese or Indian origin, any scholar familiar with the ancient civilizations of the Old World, when studying those of Mesoamerica, will necessarily be struck by the fact that their cosmological ideas, the system of lucky and unlucky days, the divination from the calendar, the institutions of the state, the hierarchy of offices, kingship, the royal courts and ceremonials, the insignia of royalty and rank (parasol, fan, litter), the schools, the administration of justice, the markets, all correspond to a very large extent to those of Asia. Here we have stressed the comparison of archaeological materials in order to establish that chronological basis which alone will enable us to put the various contacts between Asia and Mesoamerica into a plausible historical sequence. But future research will probably indicate that Asiatic influences of nonmaterial character were far more important than those in art and architecture. It must have been they which changed the whole structure of native society and transformed the ancient tribal cultures into civilizations more or less comparable to those of the Old World. 292

PROBLEMS OF DISCOVERY AND OF TRANSPACIFIC SAILING

According to an old and incomplete survey, within the one century from 1775 to 1875 twenty Japanese junks (thus an average of one every five years) were involuntarily driven by storms and currents to the American coast, which they reached at various points from the Aleutian Islands to Mexico (Brooks, 1875). It is easy to imagine the number of Asiatic ships which must have suffered the same fate in ancient times. If only one of them succeeded in returning to its homeland, this sufficed to make known the existence of land on the other side of the Pacific. There can be little doubt that the ancient Asiatic sailors followed in their eastward voyages the same route with which they probably had first become acquainted by mere accidental drifting. This means that they used the westerly winds and currents of the northern Pacific in order to reach the coast of California, which they then followed southward. It stands to reason that they availed themselves of the southeastern trade winds for their return voyages. This was the same route the Spanish galleons took through 250 years in their voyages from Manila to Mexico and back. We know little about ancient Chinese craft. But the fact that in the 6th or early 5th century B.C. the kings of Wu, around the mouth of the Yangtse, undertook expeditions against distant lands and returned with thousands of prisoners of war, indicates that they must have had sea-going vessels of considerable size (Eberhard, 1942, pp. 332-35, 338, 345). We are better informed about the ships of ancient India. In the 1st century A.D. the Periplus of the Erythrean Sea mentions the large vessels which from south India sailed to southeast Asia. Chinese sources of the 3rd century A.D. speak of south Asiatic—apparently Indian—ships which were 160 feet long and

TRANSPACIFIC INFLUENCES IN MESOAMERICA

able to carry 600-700 persons and 1000 metric tons of merchandise. Some of them had four or even seven sails. Their crews consisted largely of people from the lands and islands of southeast Asia (Pelliot, 1925). The ship on which the Buddhist pilgrim Fa-hien crossed the Indian Ocean from Ceylon to Java in A.D. 414 carried 200 persons (Giles, 1923, pp. 76-77). Ships larger than those of Columbus and Magellan and able to cross the Indian Ocean were certainly able to cross the Pacific as well. CONCLUSIONS

We do not know what instigated the first voluntary transpacific voyages of the coastal peoples of China. As far as trade was concerned, only luxury goods could have made these voyages profitable. It is easy to understand that eventually jade and feathers, both highly valued in ancient China, may have exercised a considerable attraction. The drug trade, too, so important in the Far East at all periods, may have played a role. In the 3rd century B.C. the Emperor Ch'inshe-huangti is said to have sent an expedition into the Pacific in order to search for the island of the immortals and the herb of immortality. When toward the end of the Han period Indians and Indianized southeast Asiatics stepped into the tracks of the Chinese, this may, in the beginning at least, have been mainly a continuation of that powerful eastward movement, as a result of which colonies of Indian merchants, artisans, adventurers, and missionaries spread over vast regions of Indochina and Indonesia. To judge by what we know about the history of Indian expansion, it seems unlikely that Brahmans and Buddhist missionaries would have missed the opportunity to follow sailors and merchants to a new country. Their participation would explain the Indian traits in Mesoamerican cosmology, divination, political organization. The fact that neither Hinduism nor Buddhism gained

a permanent foothold is not surprising. In southeast Asia, too, we know instances of incomplete conversion and subsequent disappearance of these religions. The wide range of dates indicated by the various Asiatic traits which must have been introduced in the course of, first, Chinese and, later, Hindu-Buddhist contacts, shows that ever since the beginning of transpacific voyages, probably around 700 B.C, the sequence was never really interrupted until the 9th or, perhaps, the 10th century A.D. Why it finally ended, we do not know. The fact that the correspondences between Asiatic and Mesoamerican traits, when taken as a whole, conform to a plausible and consistent chronological and historical pattern, may perhaps be claimed as the most significant result of the present investigation. There is no evidence of large-scale Asiatic sea-borne migrations to Mesoamerica. However, the native cultures could not have been so deeply affected as they actually were, if considerable numbers of single persons and small groups had not settled among them more or less permanently. The whole process may be compared to that which led to the formation of Indian colonial cultures in southeast Asia: the implantation of a foreign civilization on more primitive indigenous cultures by small groups of immigrants, soon absorbed by the local population, and, in consequence, the emerging of new civilizations which, despite all their own new creations, despite their original and unique characters, still allow us to recognize many of the features of both the foreign and native parent cultures from which they were derived. ARGUMENTS FOR INDEPENDENT DEVELOPMENT

The present treatise would be incomplete and one-sided without taking account of some, at least, of the arguments offered in order to disprove Old World influence in 293

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

the development of American Indian civilizations. Most of them are based on the absence of various Asiatic culture traits in America. They are, of course, wrong in principle since they do not consider the well-established fact that cultures are never borrowed wholesale by one people from another, but that invariably some kind of selection takes place. In numerous instances it is even quite easy to explain why a given trait was not adopted and even could not have been adopted in America. This concerns, for instance, the wheel and the plow. Of what use could they have been in an area where no draft animals were available? It has been objected that the wheel could have been used for handcarts or wheelbarrows. But how about habit and tradition? Even today one sees Indians with loads on their backs trudging along modern Mexican highways, but it would be difficult to discover an Indian pulling a handcart or pushing a wheelbarrow. Furthermore, we must not forget that there is no evidence of whole populations having moved across the Pacific. Therefore the presence or absence of any Old World technique depended on whether or not members of the respective profession had come to America. No special training was necessary in order to be able to fashion the wheeled toy animals, but only a professional wheelwright could have built a cart. The same applies to the potter's wheel. It could have been introduced only by professional potters. Even if a few of these had chanced to come to America, would the new technique have been able to prevail against established tradition? After four centuries of Spanish contacts many Mesoamerican Indians still fashion pots by hand without the use of the wheel (Lothrop, 1927b; R. H. Thompson, 1958b). It has often been said that if Asiatics really had come to America, they certainly would have introduced the true vault. In China the vault was used to a considerable 294

extent in brick-built tombs during the Han period. But was there any special inducement for Chinese bricklayers to emigrate to America? Whom could they have served there? In India the vault was unknown in pre-Mohammedan times, nor was it ever used by the Javanese, Khmer, and Cham. And yet, the Cham kingdom bordered on a Chinese province in which vaulted tombs abounded since the 2nd century A.D. The lack of Old World grain crops in the New World is still widely considered as the most powerful argument against the assumption of transpacific contacts. The argument, of course, is based on the erroneous belief that Asiatic relations with America must have been of the same character as those with the later Europeans. There is no evidence, however, of Asiatic conquests or of large-scale immigration of Asiatic farmers, comparable to what happened in America after Columbus. Without the massive support of large numbers of Asiatic agriculturists any attempt to introduce Old World crops, if it ever was made, was bound to fail. It could not have prevailed against the firmly established tradition of maize planting and the powerful religious significance of maize (HeineGeldern, 1958). Another argument often heard is based on the absence of any reference to transpacific voyages in Asiatic literature. But the original literature of the ancient coastal states of Chou times, Wu and Yueh, is lost, as are many works of the Han and even later periods. Moreover, it is questionable if in these early times private undertakings were ever recorded in writing. We have two Han reports on voyages to Indonesia, but significantly they both concern court envoys; the Chinese colonies which, as archaeological evidence indicates, existed in Sumatra, and perhaps also in Java and Borneo, left no trace in literature. Nothing of the Cham, Khmer, Malay, and Javanese literatures of the period in question has survived, and

TRANSPACIFIC INFLUENCES I N MESOAMERICA

whenever Indian legends speak of voyages to distant countries they are so vague that it is mostly impossible to recognize to what they refer. Attempts have been made to disprove Old World influence by explaining away single facts. This has been tried with regard to the conformities between the pachisi and patolli games and to those between the Indian and Maya lotus motifs (Erasmus, 1950; Rands, 1953). Even if the respective papers were not open to criticism from the

point of view of methods, they could not invalidate all the other correspondences. Any attempt to disprove the role of Asiatic influences in the development of American Indian civilizations would have to take into account the whole complex. It would have to give a plausible explanation for the fact that similar styles and motifs appear in the same chronological order on both sides of the Pacific. So far, no such attempt has been made.

REFERENCES Andersson, 1932 Balfour Gourlay, 1940 Brooks, 1875 Caso, 1947 and others, 1946 Commaille, 1913 Coomaraswamy, 1928-31 Covarrubias, 1947 De Terra, 1951 Eberhard, 1942 Ekholm, 1946b, 1953, 1964a, 1964b Erasmus, 1950 Fewkes, 1907 Gamio, 1922 Gangoly, 1920 Giles, 1923 Graebner, 1920-21 Hamilton, Α., 1901 Heine-Geldern, 1937, 1954, 1958, 1959a, 1959b, 1959c, 1964a, 1964b and Ekholm, 1951 Humboldt, 1816 Janse, 1947 Jayaswal, 1933 Kidder, 1947 , Jennings, and Shook, 1946 Kirchhoff, 1964 Laufer, 1909 Lehmann, W., 1938 León-Portilla, 1956

Libby, 1955 Ling, 1956 Linné, 1951, 1956 Lothrop, 1927b Lou, 1957 Marchai, 1951 Maudslay, 1889-1902 Naudou, 1961 Nomland, 1932 Nowotny, n.d. Pallegoix, 1854 Parmentier, 1922, 1954 Pelliot, 1925 Proskouriakoff, 1946 Rands, 1953 Rock, 1922 Salmony, 1938 Séjourné, 1956-57 Spinden, 1933 Stern, P., 1934 Stone, 1938 Stuiver, Deevey, and Gralenski, 1960 Thompson, R. H., 1958b Tylor, 1896 Umehara, 1936 Vessberg, 1937 Vogel, 1929-30 Vogler, 1949 White, 1934

295

15. The Role of Transpacific Contacts in the Development of New World Pre-Columbian Civilizations

PHILIP

w

ITH A FEW notable exceptions writings on the subject of preColumbian contacts across the Pacific take the form of comparisons of specific elements shared by Asiatic and American cultures without reference to geographical, chronological, or cultural contexts. The rationale is that if parallels are sufficiently close, the only possible explanation is transmission across the Pacific by means that do not require to be specified. The writers are generally referred to as "diffusionists," a not altogether appropriate designation. All anthropologists are diffusionists whether they believe in transpacific contacts or not. The usage goes back to a time when Americanists were divided by a controversy over the roles of diffusion and invention in the building of cultures. Diffusion as now understood covers the transmission of cultural ideas by any sort of means; but anthropologists increasingly tend to think of it in terms of small-scale interaction requiring nothing more than short-range 296

PHILLIPS

movements of individuals or small groups of people. In this sense the proponents of transpacific contact are not really diffusionists at all. They have been obliged by various considerations (that we cannot go into here) to abandon the idea of island-toisland diffusion across the Pacific in favor of direct sea voyages over vast distances. This has forced them back into the kind of thinking that explained everything by migrations. Voyages are not migrations, to be sure, but from the point of view of anthropological theory they are the same kind of phenomena. They are events, not processes. As such they give us the right to demand not only evidence that they occurred but where, when, and how. If the proofs of transpacific diffusion have fallen short, the implications drawn from them have been of a most far-reaching nature. It has occurred to me that a useful approach to this complicated subject might be to consider first the implications and then to proceed to an examination of the evi-

TRANSPACIFIC CONTACTS IN P R E - C O L U M B I A N CIVILIZATIONS

dence that would be required to support them. As a point of departure I take a recent paper by Wilhelm Koppers (1957) on the concept of an universal world history.1 This all-important project, he holds, is impossible of fulfillment without demonstration of the worldwide unity of culture—this being the special task of cultural anthropologists and prehistorians. Professor Koppers states that on the levels of high culture this task has already been achieved by Heine-Geldern and others, the crux of the demonstration being the proof of the unity of the civilizations of the Americas with those of the Far East, and ultimately of course the Middle East, the cradle of all civilization. He refers to this "breakthrough" as though there could be no two opinions about it. All that remains is to demonstrate a similar unity at lower levels of cultural development. Not many American archaeologists will accept this claim, but with the proposition that the problem must be dealt with in a deeper context of development there can be no disagreement. Gordon Ekholm, one of the few American scholars who has not ignored the transpacific question, has expressed a similar view (1953, p. 72). In slightly different terms, the question is not whether there have been contacts across the Pacific but whether such contacts could have been determinant at critical moments in the development of American culture. For example, it signifies little in terms of development that a lotus design of presumed Asiatic origin appears in certain Maya centers in the 8th century A.D. long 1 This paper makes clear in a most interesting fashion why continental European scholars have taken the lead in transpacific studies. Koppers holds that the concept of an universal world history, the product of 19th-century German historians, is the "main line" of historical research. Anglo-Americans, under the influence of empirical philosophy and evolutionary natural history, have been drawn off into sidetracks: materialism, determinism, parallelism, and evolutionism, conceded to be valuable in their own way but not contributing to the main task of history.

after Maya civilization achieved its classic expression. It is necessary to reject the assumption that inheres in both isolationist and diffusionist positions, to wit, that acknowledgment of the transpacific provenience of one, several, or even many, particular elements settles the question against the integrity of American civilizations once and for all. The isolationist appears to believe that if he lets anything in, the trickle will irresistibly become a flood. The diffusionist cannot be blamed for exploiting this convenient situation; all he has to do is prove the existence of a trickle. In my opinion this is the most he can so far claim to have done. Before any assertions can be made about the unity of New and Old World civilizations, it will be necessary not only to produce evidence of historical contacts, with some degree of precision as to time, place, and means of transport, but also to show that the role of such contacts was decisive in the development of Nuclear American civilizations in their formative stages, that without such contacts the level of civilization would not have been attained. The views expressed here involve two familiar assumptions about the nature of culture change. The first is that it does not proceed at a constant unvarying rate; but rather by a succession of sharp advances separated by intervals of less rapid change, stagnation, or even recession. The causes of such advances may be manifold, not to say mysterious, but always to be reckoned with is the possibility of new ideas from an external source. The second assumption is that the appearance of new features tends to coincide with other changes not necessarily related in a direct and ascertainable way. The culture seems to move forward on a broad front. In a recent seminar on culture contact situations this kind of diffusion was somewhat ponderously described as "trait unit intrusion, type B4: fusion with emergence of new traits which have no obvious antecedents in the trait units or the receiving culture" (Willey et al, 1956, pp. 22-23). 297

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

It was the view of the conferees that this is the usual way one civilization influences another. With these assumptions in mind it would seem reasonable to approach the problem of transpacific diffusion by locating the important breaks in the development of Nuclear American civilizations and then to scan the western horizon for signs of contact at these times and places. By focusing on the archaeological features that mark the breaks, they may be thought of as "events" and thereby rendered intelligible in a culture-historical context. The events that appear to have special significance in this context are: (1) the domestication of plants; (2) the introduction (or invention) of pottery; (3) the establishment of the village farming socio-economic pattern; (4) the addition of minor ceremonial centers to the village pattern; (5) the emergence of major ceremonial ("great temple") centers. In this article we shall attempt to see to what extent these events can be related to the kind of events envisioned by the diffusionists, i.e. contacts of culture-bringing voyagers from Asia. The main emphasis will be on Middle America, but none of these developments can be dealt with adequately without reference to North and South American data. BEGINNINGS OF PLANT DOMESTICATION

If we seem to be starting the argument about transpacific diffusion at a stage well along in the development of New World culture, it is not because nothing of consequence to that development happened before, but because there is no serious argument about what did happen. It is almost universally held by isolationists and diffusionists alike that the original seepage of peoples into the New World was by way of Bering Strait, and that by a time conveniently estimated at 10,000 B.C. both continents were fully albeit sparsely occupied. As to what sort of cultural impedimenta the earlier migrants brought with them, there 298

is less agreement. A commonly accepted theory is that the Arctic environment was an effective screen against all traits not compatible with a hunting-fishing-collecting mode of life. The isolationist could grant the Asiatic origin of all the technologies of the pre-agricultural levels in the New World without seriously affecting his position. This, however, would be giving away too much. We know that a considerable amount of indigenous development did in fact take place on these levels.2 The first real brush between isolationist and diffusionist comes with the stage that is marked by the beginnings of plant domestication. It is now believed that the "agricultural revolution" in the Near East was preceded by a long period of incipient cultivation, in which the chief reliance continued to be on hunting and collecting but experiments leading to the cultivation of cereals were taking place (Braidwood, 1958, 1960). The creation of an agricultural system is certainly one of the greatest achievements in human history, and Neolithic culture (in the sense of village agricultural subsistence) was made possible by its spread from the Near Eastern hearth to most parts of the Old World where cultivation of cereals was possible.3 Finds of 2 If recent opinions of Americanists who also know something about Siberian archaeology are correct, the bifacial projectile point tradition which was the mainstay of the early American hunters has temporal priority as well as technical superiority over comparable Siberian developments (Tolstoy, 1958b; Chard, 1958b; Griffin, 1960a). This might be taken as an earnest of American inventiveness at later stages more important from the standpoint of civilization, but we need not insist on it here. 3 This of course is not the whole story of agriculture in the Old World. In the area of southern and southeastern Asia an agriculture based on an entirely different assemblage of plants came into being (Sauer, 1952), and recently the possibility of a third center in Africa has been suggested (Anderson, 1960). The priority and relationships of these Old World centers of agricultural dispersal are subject to dispute, but some connection between them is probable. Here the diffusionist does not have to reckon with the possibility of wholly separate and independent "in-

TRANSPACIFIC CONTACTS I N P R E - C O L U M B I A N CIVILIZATIONS

domesticated plant remains in early New World contexts now afford the view of a similar period of incipient farming, not as early in its inception, but lasting as long if not longer. This is not the growth pattern one would expect if the knowledge of plant domestication came to the New World fully developed. Botanical arguments for transpacific diffusion are usually in the form of evidence that specific plants were known on both sides of the Pacific in pre-Columbian times, the assumption being that the transfer must have been by the agency of man. A further implication not always expressed is that the entire development of American agriculture originated in such a transaction. Such views were considerably shaken by the early dating of domesticated plant remains in the Huaca Prieta midden on the north coast of Peru (Bird, 1951) and will be even more so by recent developments in the archaeology of Tamaulipas, Mexico (MacNeish, 1958; Mangelsdorf and Reeves, 1959). Without insisting on the somewhat doubtful evidence of domesticated plant remains in the Infiernillo phase (7000-5000 B.C.) in Tamaulipas, it appears that by the Ocampo phase (5000-3000 B.C.) cultivation of gourds, squash, peppers, and beans is fairly certain. No one has produced acceptable evidence, or even an hypothesis, of sea-borne contacts across the Pacific at this remote period. Leaving aside the question of domestication in general to focus on the plant that furnished the economic basis for Nuclear American civilization, we find the temporal vention" of plant domestication. It is interesting but not germane to the present issue that in lowland South America there may have been a separate center of agricultural dispersal based on root crops, as in southeast Asia, and here also the relationship with the Nuclear American seed-plant center is in question (Sauer, 1952; Rouse, 1960b). 4 The discovery of fossil maize pollen in drillcores from 200 feet below Mexico City (Barghoorn and others, 1954) appeared to settle any doubts about the American origin of maize; but possibility of misidentification of the pollen has recently been suggested (Kurtz, Tucker, and Liverman, 1960).

context is only slightly later. The most primitive maize so far discovered is from Bat Cave, New Mexico, with a series of radiocarbon dates on associated charcoal running back to about 4000 B.C. (Libby, 1951; Dick, 1952). Precise data on the association of the corn and charcoal have not been published, so this early date may be justly regarded with suspicion. A safer date perhaps for the considerably less primitive maize of the La Perra phase in Tamaulipas is 2500 B.C. (Libby, 1952). These finds of early maize have taken much of the steam from the hotly debated question of Asiatic origin of maize (Stonor and Anderson, 1949; Carter, 1950; Mangelsdorf and Oliver, 1951; Ho, 1955). 4 The emphasis now has shifted to the possibility of pre-Columbian diffusion of maize from America to Asia. Less important from our standpoint is the argument about the grain amaranths, anciently cultivated in the New World and also known from certain regions of southern and central Asia. So far as I am aware it has not been stated with confidence that the amaranths were cultivated in Asia before they were in America, so the outcome of this particular question has little bearing on the origins of American agriculture. Ever since the pioneer papers of Hutchinson and Stephens (Hutchinson, Silow, and Stephens, 1947; Stephens, 1947) suggesting relationships of Old and New World cottons, the interested layman has watched from the sidelines a botanical dispute of terrifying complexity. Discovery of cotton and cotton cloth in the pre-maize levels of Huaca Prieta added some nonbotanical complications. Hutchinson's theory requires the crossing of an American wild cotton with a cultivated Asiatic cotton, the latter having been carried across the Pacific by man, and the Huaca Prieta dates require that this event be pushed back to some time before 2500 B.C. According to Heine-Geldern's findings, cotton was introduced into southeast Asia from India some time after the beginning of the Christian Era and 299

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

reached China even later. It was cultivated at Mohenjo Daro in the Indus Valley as early as the latter half of the third millennium B.C., perhaps earlier still in the Near East, but even to Heine-Geldern the notion of a direct transfer by sea from India or Arabia to South America as early as 2500 B.C. "sounds fantastic" (1952, p. 347). The only comfort he can find is in the possibility that Huaca Prieta cotton was not like other American cottons, that the crossing hypothesized by Hutchinson and his colleagues actually took place at a later date, but in a more recent paper (1958, p. 397, note 101) he points out that this solution involves some difficult botanical problems. His final conclusion, that we must wait for more evidence from the botanists, is one with which no one can fail to agree. My impression is that the same conclusion can be extended to all botanical aspects of the transpacific question, but I cannot refrain from commenting further on the paper just cited because it exhibits a remarkable shift of strategy. Heine-Geldern is now prepared to accept, even to welcome, the American origins of corn and the grain amaranths. Their introduction into Asia, he contends, was by the return voyages of his south Chinese and Dongson adventurers of the first millennium B.C. Thus in the special case adverse findings are turned to positive account. But the general implication seems to be that it is a matter of slight concern whether the original impulse to cultivation in America came from within or without, 5 a position not heretofore taken by proponents of transpacific diffusion. This goes far beyond the question of the independence 5

"Beim gegenwartigen Stand der Forschung lâsst sich die Frage nicht entscheiden. Gleichgültig aber, ob der Pflanzenbau in Amerika selbständig ist oder auf Anregungen zurückgeht, die aus der Alten Welt gekommen waren: Zur Zeit als die ersten starken Einfliisse asiatischer Hochkulturen sich gelten machten, wahrscheinlich im 8 Jahrhundert v. Chr., hatte er in Mexico, Zentralamerika und Peru beriets sein charakteristisehes Geprage mit dem Mais als Hauptfrucht angenommen" (1958, p. 382).

300

of American agriculture. If the first and most decisive step toward civilization was taken without stimulus from cultures further along that road, it becomes more difficult to reject the possibility of other steps of similar independent nature. In a more specific sense, if the characteristics of Nuclear American maize economy—and I think Heine-Geldern would agree that associated religious and social features would have to be included—were already set before the first transpacific influences could have been brought to bear, it raises a serious question as to how decisive such influences (if proved) could have been at any subsequent stage of development. FIRST APPEARANCE OF POTTERY

It cannot be said that the earliest pottery in Nuclear America marks a turning point save in the sense that from here on the archaeologist is provided with an analytical tool dear to his heart. Nevertheless, pottery is relevant to the present inquiry as one element that we have some hope of tracing to its ultimate source, or sources, and this might tell us something about diffusion in general. In Nuclear America it had been customary to equate the earliest pottery with the beginnings of the village farming pattern in an "Archaic" "Formative" or "Preclassic" period, but more recently archaeologists have been finding pottery in coastal shell middens under conditions similar to those of North America, where the equation of pottery with agriculture has long since broken down. However, the temporal frontiers of agriculture have been pushed back also. The general concensus is that the first appearance of pottery was long after the beginnings of plant domestication but before the consolidation of the village farming pattern (Willey, 1960). The threshold of pottery has been pushed back, but this has not brought the problem of origins any nearer solution. The most influential theory at the moment is that Nuclear American

TRANSPACIFIC CONTACTS IN PRE-COLUMBIAN CIVILIZATIONS

pottery is the result of a fusion of two separate developments: the Woodland tradition of North America, with its ultimate origins in Asia; and an indigenous Nuclear American tradition the dispersal center of which has not yet been located. Another theory would derive all Nuclear American pottery from the Woodland tradition. Thus both explanations involve the possibility of ultimate Asiatic origin in whole or in part. At present (1961) certain data must be accommodated in any hypothesis of origin: (1) The earliest dated potteries in South America and lower Central America are in the third millennium B.C.: Puerto Hormiga, Columbia, ca. 2900-2500 B.C. (Reichel-Dolmatoff, 1961); Valdivia, Ecuador, with three dates on shell ranging from ca. 2500-2100 B.C. (Evans, Meggers, and Estrada, 1959); Rancho Peludo, Venezuela, ca. 2500 B.C. (Cruxent and Rouse, 195859); Monagrillo, Panama, ca. 2100 B.C. (Willey, 1958, 1960). Of these potteries Monagrillo is perhaps the most primitive, but none are what might be expected of the initial stage of pottery technology. (2) The earliest pottery in Middle America, from Tehuacan, Puebla, ca. 2300-1300 B.C, has not yet been described. It is sufficiently crude to have prompted a hypothesis of local origin from stone prototypes of the underlying preceramic stage (MacNeish, 1961). The earliest pottery in the Valley of Mexico, Early Zacatenco, is considerably later (ca. 1360 B.C.) and more advanced technologically (Vaillant, 1930; Johnson, 1951). (3) The oldest pottery in North America is the plain fiber-tempered phase on the Georgia coast with five dates ranging from about 3500 to 1000 B.C. The earliest and latest of these dates appear to be out of line; the three in between are in the order of 2100 to 1700 B.C.6 This fiber-tempered pottery does look technologically primitive. Specialists have suggested that it may have been a local development sparked by stimulus diffusion from the Woodland tradition of the Great Lakes region to the north

(Sears and Griffin, 1950), but the dating now available runs counter to such an explanation. (4) The earliest Woodland pottery, Vinette I, dates from about 1000 B.C.7 Many of the characteristic features of Woodland pottery are believed to have diffused from Asia via Bering Strait (Tolstoy, 1953, 1958b), but the specific type Vinette I has no close counterpart in Asia or the North American Arctic (Griffin, 1960a, p. 811). Oddly enough, the most striking Eurasian parallels to later Woodland pottery (of the Point Peninsula type) seem to be as far away as possible from Bering Strait, in north Russia and Scandinavia (Ridley, 1960), a circumstance that no one has yet been able to explain. Juxtaposition of these findings is not favorable to the theory of Woodland origin of Nuclear American pottery in any specific sense. To leave aside for the moment the fiber-tempered ware, which seems to be a problem all to itself, the chronological data do not support diffusion in a north-south direction. Even as between Middle and South America, the temporal gradient seems to slope the other way. Individual features considered as Woodland, particularly rocker-stamping and zoned rocker-stamping, have been reported from many early phases including Valdivia, but the features in question are not among the earliest or the most fundamental features of Woodland pottery in the home area. So the dating actually favors the theory that these features came into Woodland from the south, as some North American archaeologists have maintained (Griffin, 1952b, p. 126). One pottery that has a slightly more authentic Woodland cast is the recently discovered Ocos phase of the Guatemala coast (M. D. Coe, 6 Only one of these dates has been published (Crane, 1956). The others are Humble Oil Company dates (information by William G. H a a g ) . I understand there are now additional unpublished dates from Florida in the same order of magnitude. 7 Following Griffin (1958, p. 10) in rejection of the date of ca. 2500 B.C. for Vinette I pottery at the Red Lake site in New York State.

301

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

1960b, 1961). This has several of the more fundamental Woodland elements, such as cord- and fabric-marking, but also many other features that have close counterparts in Ecuador, and these have no relationship whatever to Woodland. The phase has not been dated, but on the basis of cross-ties M. D. Coe has given it an estimated time span of 1500-800 B.C. The youngest of these dates is not inconsistent with a North American Woodland connection but, as Coe himself has pointed out, the most specific Woodland traits in Ocos, i.e. those that might be considered as reliable time-markers, such as zoned shell edge and dentate rocker-stamping, are Middle Woodland or "Hopewellian" features. The earliest (Illinois) phase of Hopewellian is abundantly peppered with radiocarbon dates falling within the period from 500 B.C. to A.D. 1. It seems that even Coe's terminal estimate is too early for the theory of Woodland derivation. It might be observed, however, that more thoughtful comparisons are required before any of these Nuclear American features can be called Woodland or Asiatic. As Paul Tolstoy pointed out some time ago (1953, p. 36), before the possibility of Woodland influences in Nuclear America became an issue, progressive attenuation of Asiatic features with increase of distance from their source is to be expected. Perhaps the emphasis on specifically Woodland connections in Nuclear America has been misplaced. Woodland is after all a refocalization of Asiatic traits that came into eastern North America not as a complex but separately and at different times. Some of them, particularly the earlier arrivals, might be unrecognizable as Asiatic today. Fibertempering may be a case in point. It has been reported from both Asia and Alaska (Chard, 1958a). Does this not suggest a similar origin for other ceramic features that are not specifically Woodland? Whatever the ultimate origins, it is be302

coming clear that the earliest diffusion of pottery in North America was in a nonagricultural context, and this may have been true for Nuclear America as well. At some as yet unspecified point in time it was incorporated into an incipient farming culture and there began the well-known American agriculture-pottery association that has been the subject of so much generalization in the past. If this is the correct view, the effect of an admission of remote Asiatic origin of pottery on the integrity of Nuclear American civilizations would be just about nil. ESTABLISHMENT OF VILLAGE FARMING

Tangible evidence for the transition from incipient cultivation to village agriculture is available in only two regions in Nuclear America, Tamaulipas and the north coast of Peru.8 In his long stratigraphic sequence in the Sierra de Tamaulipas, MacNeish (1958) has been able to show a steady increase in percentages of domesticated plant remains (in terms of volume of all food remains) from about 5 per cent in the upper deposits of the Nogales phase (ca. 3000 B.C.) to 40 per cent in the Laguna phase (ca. 500 B . C ) . This figure of 40 per cent apparently represents agricultural stability in this rugged but not forbidding environment, where game and wild plant foods must always have been plentiful. There was a gap, however, in the sequence preceding the Laguna phase, and MacNeish thinks it likely that agricultural stability was reached here before 500 B.C. In other not yet fully reported excavations in the Sierre Madre of southwestern Tamaulipas this gap appears to have been filled. Here, as we have seen, the shift to cultivation may have begun as early as the Infiernillo phase (ca. 7000-5000 B . C ) . 8 MacNeish has also found evidence of this transition in caves in Puebla, central Mexico (personal communication, 1960).

TRANSPACIFIC CONTACTS I N P R E - C O L U M B I A N CIVILIZATIONS

The 40 per cent stability figure was reached in the Mesa de Guaje phase equated with the pre-Laguna gap mentioned above (ca. 1500-500 B . C ) . Food plants in this phase included a number of varieties of maize, several varieties of beans, warty squash, pumpkin, sunflower, cotton, and bottle gourds. Pottery and cotton cloth appeared for the first time in the sequence. There is little information on settlements, but what there is suggests small villages of scattered houses without specialized ceremonial or religious structures (MacNeish, 1958, pp. 168-69). The chief interest of this peripheral evidence to the present inquiry lies in the extent to which it may be assumed to reflect what happened in the more central parts of Mesoamerica, where the earliest agricultural phases so far known are already firmly established on a village farming level. Before we consider these, it will be well to glance briefly at the situation in Peru. Almost all the relevant Peruvian information is from the north coast where, at the mouths of the many rivers coming down from the Andes, small areas of water-holding soil made some cultivation possible. Combination of this and the more dependable resources from the sea was sufficient to permit the growth of fairly large and stable villages. In terms of our nomenclature, under these rather special conditions, village settlement patterns developed on an incipient agricultural base. About 1400 B.C (the date is not securely fixed) maize was incorporated into this economy, and shortly thereafter, pottery. Soon there began a shift of population up into the valleys away from the sea, where more land was available for cultivation. It is an arid country, even in the valleys. Whether this shift could have taken place without the aid of irrigation is a point that has not been settled. In any case, this is the threshold of village farming that we are seeking. The earliest chronological datum for this new economic pattern is the

Cupisnique phase in the Chicama Valley, with a radiocarbon date of ca. 700 B.C. But Cupisnique is a coastal phase of the Chavin culture of the highlands, the type site of which Chavin de Huantar is a cult center in no uncertain terms. It is now believed that other elaborate centers farther down the coast, such as Las Aldas, may date from a still earlier pre-maize period (Engel, 1957b). Thus the courses of development in Peru and Mesoamerica are clearly not the same. By the time the threshold of village farming as herein defined (on the basis of Mesoamerican data) has been reached in Peru, there are already indications that some cultures at least have passed beyond it. The explanation of this apparent anomaly may lie in the fact that the first appearance of maize in Peru cannot be dated earlier than about 1400 B.C. (Collier, 1959). The corresponding date in Middle America (backed up by the United States Southwest) is more than a thousand years earlier. Technical analyses of early Peruvian maize have not been published, but it has recently been suggested by the most qualified experts that its great diversity resulted from the hybridization of Peruvian and Mexican popcorns (Mangelsdorf and Reeves, 1959, p. 422). If our assumption is correct that it was maize that made possible the shift to a village farming economy in Mesoamerica, it might be argued that a similar shift in Peru had to wait for the Mexican component of the hybridization postulated by Mangelsdorf and Reeves, and by the time this happened advances toward more complex social and religious organization had already taken place. We do not have to assume that these other developments also diffused from Mesoamerica, although it is a possibility. This of course is a flagrant oversimplification, taking no account of the Peruvian highlands, where data on early agricultural developments are not yet available. In any case on present evidence it is 303

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

clear that priority, so far as this particular turning point is concerned, rests with Mesoamerica. This is the ground on which the case for transpacific diffusion, if there is a case, must be argued. The earliest dated occurrence of the village farming pattern in Mesoamerica is Vaillant's Early Zacatenco phase in the Valley of Mexico (Vaillant, 1930) with a radiocarbon date of 1360 B.C. ± 250 (Johnson, 1951 ). 9 The earliest phase of the closely related El Arbolillo site (Vaillant, 1935) is thought to be slightly older. This is of no great importance here except as a further indication that a date of 1360 B.C. for the beginning of the village farming era is probably on the conservative side. What is important is that these cultures show no signs of fumbling beginnings or sudden development under the stimulus of new ideas. One has a strong impression that this is not the threshold of village farming but merely its first appearance in the archaeology of a particular region. It is in fact held that the emergence of fully settled village farming took place somewhere in the southern part of Mesoamerica. The reason for this supposition is that the earliest known Preclassic cultures of the highlands and Pacific coast of Guatemala, though not as yet securely dated, seem to have been further along in social and economic development as evidenced by the presence of fairly large temple mounds. At the present time, however, the El Arbolillo I-Early Zacatenco culture seems to be as close as we can come to the threshold

9 A slightly earlier date of 1456 B.C. ± 250 from the famous burial site of Tlatilco (Johnson, 1951) has not been confirmed by recent dates run on more carefully selected samples, but is perhaps not entirely worthless. According to Μ. Ν. Porter (1953), there were Zacatenco sherds scattered through the soil of the burial area and in bottleshaped pits nearby, so it is altogether possible that this date relates to an Early Zacatenco component at the site.

304

of village farming in Nuclear America while still remaining on firm ground so far as documentation and chronology are concerned. It remains now to see what are the salient features of this culture and what the chances are that any of them can be attributed to influence from across the Pacific. For this purpose the barest outline will suffice. Settlements were small lakeside villages, that can certainly be characterized as "permanent," to judge from the depths of the refuse. Total area of the Zacatenco site was about 3.5 acres, that of El Arbolillo somewhat larger. Vaillant gave no population estimates. A more recent authority suggests 200 inhabitants as a probable village limit (Piña Chan, 1955, p. 25). Houses were small, rectangular, constructed on a framework of poles with walls of wattlework daubed with mud and roofs of thatch, a type of dwelling that remained in vogue throughout all periods of Middle American prehistory and is still in use today. There is no evidence of any other kind of structures, religious or secular, no defensive works, and no special repository for the dead, who were simply placed in shallow graves scooped into the soft and pungent refuse that accumulated around the dwellings. Burials were not oriented consistently and no rigid canons with respect to position of the body were in evidence, though tight flexure was the most common form. Some burials were covered with red ochre, a custom that may have a remote Asiatic origin. Its early occurrence in North America, however, is not favorable to a transpacific route. Grave offerings were not lavish; at Zacatenco they were very often lacking altogether. Technological aspects of El Arbolillo I Early Zacatenco culture were equally undistinguished. It was plainly an era of household arts, with few signs of specialized craftsmanship. Implements of ground or chipped stone, the latter mainly of obsidian,

TRANSPACIFIC CONTACTS I N P R E - C O L U M B I A N CIVILIZATIONS

conform to patterns that are practically universal in New World cultures at a comparable level of development. Nevertheless, the presence of a few simple but well-finished ornaments of jade and small flat bits of turquoise reveal the beginnings of crafts that are to show characteristic development in succeeding periods. Pottery, to quote Vaillant, shows a "smug competence uninspired by artistic yearnings." Ninety per cent was a purely utilitarian "bay ware," but there were also more specialized black, white, and painted wares. Shapes were simple bowls and jars with rounded bottoms, but a few vessels had small tripod supports, a feature of interest in the present inquiry. Decoration was by means of incision, sometimes paint-filled, and white-onred painting. Designs of simple linear motifs showed little if any symbolic content. Pottery was the preferred material for many small objects—rattles, whistles, bells, spindle whorls—but the outstanding products from any point of view were the figurines. These fresh and lively little objects of problematical function display something more than the "smug competence" of the pottery, although they were to receive even more intense expression in subsequent periods. Of religious and social aspects of the culture not much can be inferred. A great deal of nonsense has been written about figurines as expressions of an "earth mother" or "fertility" cult. True enough, in El Arbolillo I-Early Zacatenco times they were exclusively female, but it is noteworthy that the interest is centered on the heads, with an endless and charming variety of hair style and headdress, rather than on those portions of the female anatomy devoted to procreation. A sounder inference can be based on the lack of symbolism—this applies to pottery decoration as well—as compared with later figurines (e.g. Tlatilco) which, in addition to being often male, appear as magicians or priests with masks

or god-animal attributes, or as ball-players with helmet, pads, and other ritual paraphernalia. The contrast suggests that there was a lack of complex religious organization in the earlier period. Now what can the transpacific diffusionist make of this? If I am correct in supposing that it is important for him to be able to show that this village farming point was not independently "turned," what specific features of this culture can he point to as trace elements of Asiatic intervention? The basic technologies—agriculture, stone and bone, pottery, weaving—were already present in the incipient farming era. The only features that might be considered as new that have figured in transpacific writings are jade carving, turquoise mosiacs, tripod vessel supports, and figurines. Jade and Turquoise Now that sources of jadeite have been located in the Guatemala highlands (Foshag and Leslie, 1955), there is no further basis for the supposition that the raw material as well as the techniques of jadeworking came from Asia. Nevertheless jade does continue to figure in the transpacific question in a slightly different way. According to Heine-Geldern (1954), his Chinese voyagers of the 8th century B.C. were in search of gold, which was found in Peru but not in Middle America. For their interest in the latter area other lures must be assumed. "Where," asks Krickeberg (1956, p. 567), "did the Olmecs get their advanced techniques of working jade?" The simple little El Arbolillo jades are not to be overlooked in this question. I would merely point out that 1360 B.C. is a half-millennium too early for the Heine-Geldern hypothesis. The technique of inlaying turquoise has been treated in the same offhand fashion (Covarrubias, 1954, p. 54) as though its mere presence on both sides of the Pacific were sufficient evidence of contact. 305

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

Tripods Vessel supports have an immensely long history in China and the New World, in the course of which all possible changes have been rung on this basically simple invention. It would be surprising if, out of countless variations on both sides of the Pacific, some remarkable parallels were not to be found. Covarrubias has done so (1954, fig. 25). His earliest Middle American example is from the Las Charcas phase at Kaminaljuyu. There is every reason to believe that Las Charcas is later than El Arbolillo I. Covarrubias would derive the Las Charcas tripod from China because he has been able to find an example from Kansu that it closely resembles. If tripods were already in existence in Middle America, I would think it more economical to derive the Las Charcas example from them, never mind how close the resemblance. Figurines I would be disinclined to mention figurines and "fertility cults," but the subject comes up in transpacific writings that are not to be ignored (e.g. Ekholm, 1955, p. 103). The unsuitability of El Arbolillo I Early Zacatenco figurines as expressions of fertility has already been noted. The Valdivia figurines (Evans, Meggers, and Estrada, 1959), probably the oldest in the New World, are either unsexy females or androgynous. More significant perhaps is the possibility that Valdivia was not an agricultural phase. On the other hand, the evidence for figurines in known incipient agricultural phases is negative because these had no pottery. In short, the theory that figurines participated in the original spread of agricultural practices in the New World is entirely unsupported by the facts. Perhaps in this sketch of what is currently believed to be the earliest village farming culture in Nuclear America there are other less trivial items of possible Asiatic origin that I have overlooked. But even if 306

the case for such intrusions were clear and irrefutable, it would still remain to show that they could have been crucial in a developmental sense. This, so far as I know, has not even been attempted. We can say therefore with some degree of confidence that the course of development toward civilization in Nuclear America reached the stage of village farming at a time probably not later than 1360 B.C. without serious intervention from across the Pacific. VILLAGE CEREMONIAL CENTER

At the present time it cannot be stated with certainty that village ceremonial centers marked by "temple" mounds of earth were not present in some places in Mesoamerica as early as the threshold of village farming. They are claimed to be associated with the earliest known Preclassic cultures in Guatemala: the Arevalo phase at Kaminaljuyu (Shook, 1951) and the Ocos phase on the coast (M. D. Coe, 1960b, 1961). Although these have not been dated earlier than El Arbolillo I-Early Zacatenco, it cannot be considered an impossibility. In my view there is no serious anomaly in the existence of platform mounds in simple village farming communities, but this does not accord with the view that such mounds imply the existence of societies more complexly organized than is consistent with the concept of simple village farming. If this is correct, the concept of a simple village farming era in Mesoamerica may be a fiction. I have already called attention to the difficulty of establishing such an era in Peru, so the whole proposition is in danger. In my opinion the Guatemala evidence does not support this pregnant interpretation.10 The expression "temple mound" has 10 In neither case has the full evidence for dating the mounds been published. I understand that at Kaminaljuyu dating was by means of sherds on the surface and in the mound fill, at the site near Victoria from surface sherds alone. Both methods of dating are subject to the same possibilities of error, i.e. sherds from an earlier occu-

TRANSPACIFIC CONTACTS I N P R E - C O L U M B I A N CIVILIZATIONS

implications of complex religious institutions. Actually we know nothing about the kind of buildings supported by these primitive earthen platforms. The only inference backed by any concrete evidence is that they were funerary. If we were to call them "burial mounds," the implication of multivillage organization might disappear. There is no clear indication that the mounds in question exceed in size or numbers what might be expected in the public architecture of simple village communities. In the Ohio and Mississippi Valley we find earthen mounds of considerably larger dimensions associated with cultures that are perhaps not even "up to" the village farming level. This is not a matter, however, that any one can afford to be dogmatic about. We certainly must reckon with the possibility that future research in southern Mesoamerica will reflect the existence of complex multivillage patterns at the very beginning of the Preclassic. If so, the diffusionist could with perfect reason seize on this as the turning point in Middle American prehistory, a rags-to-riches leap from the stage of incipient cultivation to something approaching civilization—a change containing the seeds of all subsequent developments in Middle American culture. For such a forward bound what more natural than to look for an external stimulus? A diffusion hypothesis might go something like this: Experiments with plant domesticapation may be contained in soil used in building the mound, and these may subsequently be washed out by erosion and left on the surface. 11 Ekholm (1955) says that "the early pyramid tombs of the Miraflores period in Highland Guatemala have an unusual resemblance to the Shang tombs of China," but gives no particular references. According to information from Kwang-chih Chang, only one small platform "altar" dating from the Shang period has been reported (but not yet published). The great tombs of Anyang (Chu'ü-hsun, 1959) are in deep pits with sloping sides and ramps, very like a Maya pyramid in intaglio. Heine-Geldern (1956b) offers as prototypes the "gigantic grave pyramids of Chinese rulers of the last half of the first millennium B.C." but these are not early enough.

tion began very early in America but led to nothing until a more advanced people appeared, bringing the degree of social and religious organization necessary to establish, on the basis of the food crops made available by said previous experimentation, a stable farming economy. The visible symbol of this revolutionary change is the tomb or temple platform mound of earth. The round date of 1500-1000 B.C. would indicate China as the only possible source of the intrusion. This would be followed by the usual encomiums on Chinese seafaring at this remote period based on evidence of 2000 years later. The arguments that could be deployed against this hypothesis are not despicable. (1) Early Preclassic art does not support a theory of complex religious development. This is not to say that religion played a secondary role at this time, but simply that it gives no indication of having reached the "cult" stage, reflected in archaeology by the dissemination over wide areas of art styles characterized by the dominance of symbolic over purely decorative and naturalistic forms. (2) There is no present evidence of platform mounds in China from which these early Middle American structures could have derived. 11 (3) It has not yet been demonstrated that platform mounds appeared in Middle America fully developed and with dramatic suddenness. A strong case could be made in favor of the proposition that they developed slowly from very small beginnings. In many parts of the world structures of whatever purpose—residential, public, or religious—have tended to be placed on foundation platforms. Given the ever-present tendency for height above ground to be an expression of prestige, such platforms are certain to grow by accretion. So long as their size and number in any one site does not exceed what could be estimated as the capacity for moving dirt and stone of a single village community, implications of further social and religious complexity are gratuitous. When that ca307

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

pacity is plainly exceeded, there are usually other accompanying features to show that more complex institutions have been achieved. EARLY TEMPLE CENTERS

Up to this point we have seen nothing but difficulties for the diffusionist. The agricultural beginnings lie in a period too remote for diffusion by any other than the forbidding Arctic route. Pottery may have come in by this route but its influence in a developmental sense is negligible. The union of agriculture and pottery and the formation of simple village settlements with or without small-scale ceremonial or funerary structures seem to have required no impulse from overseas. We conclude that village farming culture in the New World is neolithic only in a homotaxial sense. When it comes to the point, or points, where civilization begins to rear itself upon this substructure, the question of diffusion becomes acute. The criteria of civilization have been listed elsewhere in this Handbook, with particular reference to the fully developed Classic civilizations of Middle America. The point of view of this article is necessarily not so limited. These critical elements did not appear all at once and in one place. If they had, there would be ample reason to suspect a massive cultural invasion from an external source. I do not think it necessary to spend any time proving that this is not what happened. Civilization is neither a tight association of cultural elements traveling about the world as a "complex" nor a haphazard aggregation of random achievements with no functional relationship between them. If a culture finds the way to one or several such achievements, through its own creativity or by diffusion, or both, others are pretty sure to follow. This permits us to focus on the first appearance of any of the critical elements of civilization instead of waiting for all of them. These first appearances are in the Middle Preclassic period of Mesoamerica 308

and the Early Formative (Chavin horizon) of Peru, coincident with the emergence of "great temple centers," the last turning point I shall consider in this article. It could be argued that the transition from village ceremonial centers to great temple centers is more a matter of growth than change, but this would be to ignore a number of elements, among them some of the criteria of civilization, that seem to have come into existence only with the rise of these great centers. These are: (1) monumental architecture, (2) planned assemblages, (3) sculpture, (4) metallurgy (Peru), (5) hieroglyphic writing (Mesoamerica), (6) great art styles. This is a strictly archaeological list. Reflected in these features are intangibles such as widespread religious movements, complex sociopolitical institutions, craft specialization, and trade. These figure only implicitly in the transpacific question. It will be noted that nothing has been said on the demographic side. The evidence is murky, but so far nothing seems to indicate that the rise of early temple centers was attended by, or dependent on, any far-reaching changes in cultivation and settlement patterns. A new kind of relationship between villages and temple center is obvious, but as yet there are no signs that populations were being drawn to these centers. In short, I am taking the position, without attempting to prove it, that the early great temple centers afford a context for consideration of the beginnings of civilization, and the possibilities of Asiatic influences therein, without involvement in the complex issue of urbanism. Uncertainties of dating make it impossible to focus the discussion on any single site. We cannot even confine it within one or the other of the two nuclear areas. A point that I should think worth making, if I were on the diffusionist side, is that great temple centers appear to have arisen in Mesoamerica and Peru about the same time. Most securely dated is La Venta in Tabasco, the

TRANSPACIFIC CONTACTS I N P R E - C O L U M B I A N CIVILIZATIONS

outstanding site of the Olmec culture, with a series of radiocarbon dates ranging from about 800 to 400 B.C. (Drucker, Heizer, and Squier, 1959, pp. 264-67). Kaminaljuyu in Guatemala and Monte Alban in Oaxaca had almost certainly attained the scale of great temple centers before the end of this interval, perhaps also Uaxactun in the Chicanei phase. Chavin de Huantar in the north Peruvian highlands is cross-dated, through the related Cupisnique on the coast, at about 700 B.C. Thus 700 B.C. may be taken as a round date for the rise of great temple centers in Nuclear America. It could hardly have been much later, but might have been somewhat earlier. It is about this time, Heine-Geldern thinks, that the maritime kingdoms of South China developed the seafaring capability to cross the Pacific. The apparent synchroneity, however, cuts both ways. If these early temple centers developed in response to the same stimulus, one would expect to be impressed more by the similarities than by the differences between them. The most I can do here is to refer briefly to some of the differences under the headings suggested above. Monumental

Architecture

There is an understandable reluctance to consider as "monumental" those structures that do riot involve the use of brick or stone masonry. In the context of this discussion we certainly have to do so. The builders of the great center of Kaminaljuyu used nothing but packed earth, a form of construction so elemental that it is scarcely worthwhile to discuss its origins. At La Venta massive blocks of columnar basalt were used exactly as though they were logs of wood. The similarity to the way stone was handled in Polynesia (at a much later time) is striking. The theory of local development of the Polynesian structures seems eminently reasonable, and the arguments in its favor apply to La Venta with equal force (Emory, 1943). The contrast between these

primitive techniques of construction and the beautiful ashlar masonry of Chavin de Huantar is only equaled by the difference in the structures themselves. One searches Mesoamerica in vain for anything remotely resembling the great Castillo of Chavin with its multistoried rooms and galleries. The relatively modest temple platform known simply as E-VII-sub at Uaxactun is constructed on another system entirely different from any of the foregoing (Ricketson and Ricketson, 1937), a plaster coating over a hearting of earth and rubble, foreshadowing the stone veneer construction of later periods. It may be out of place but I cannot refrain from commenting on the parallel, seldom omitted in transpacific writings, between the Classic period Maya lowland temple-pyramid and similar structures in southeast Asia, especially Cambodia. For example, a specific comparison has been made between the stone-covered Temple II at Tikal and the brick funerary temple of Baksei-Chamkrong in Cambodia (Ekholm, 1950). The first dates from about the 5th century A.D. and the second from the 10th century A.D. It goes without saying that the builders of Tikal Temple II could not have been influenced by Baksei-Chamkrong. What is said, or inferred, is that both derive from some earlier Asiatic prototype. The exact whereabouts and dates of such prototypes are not given. According to students of Cambodian architecture, Baksei-Chamkrong is in an early or transitional phase of classic Cambodian architecture, a period of rapid development and change (Briggs, 1951, p. 126). If there are prototypical structures, in Cambodia or elsewhere, dating say five centuries earlier, is there any likelihood that they would resemble Temple II at Tikal as closely as does BakseiChamkrong? To ask the question is to answer it. The Tikal structure is also in an early classic stage. Behind it lie several centuries of development, going back perhaps to the lowly prototype, Str. E-VII-sub 309

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

at Uaxactun; behind that, according to present dating estimates, the earth and adobe pyramids of the Arevalo and Las Charcas phases at Kaminaljuyu. To my way of thinking the real point brought out by this comparison is quite different from what was intended. It seems that two widely separated and noncontemporaneous architectural traditions have been able to produce, at given stages of their own internal development, buildings that are remarkably similar. If this proves anything, it is that convergence is a factor to be reckoned with in long-range cultural comparisons. Planned Assemblages The plans of the early temple centers under consideration vary as much if not more than the architecture. La Venta in Mesoamerica and Las Aldas in Peru have vaguely similar axial plans, but Chavin de Huantar bears no resemblance to either. We do not know much about the plan of Kaminaljuyu when it first became a temple center, but at its peak it consisted of a number of tightly integrated units, each consisting of several pyramidal structures grouped about a central court. This is the basic Mesoamerican plan. Every list of Asiatic parallels finds room for pyramidal mounds, but no one so far as I know has considered them in terms of total assemblages and possible cosmological inferences. This might be a useful approach. If, for example, they were grounded in the same Asiatic tradition, it would be reasonable to expect in the planning of the Maya some reflection of the cosmic symbolism implicit in the planning say of the Khmer, i.e. the concept of a celestial mountain, center of the universe, consisting of a pyramid or quincunx of tower sanctuaries with surrounding wall, moats, gates, and avenues leading to the central feature (Coedès, 1947, 1948; Heine-Geldern, 1956a; M. D. Coe, 1957a). One sees no sign of this in the typical Maya plan. It might be argued that nothing is left but the pyramid itself, the 310

sacred mountain. 12 But Maya pyramids are not usually centrally placed. If there is a central feature, it is the plaza with pyramids grouped tangentially around it. More often there is no consistent organization of plan that would invite speculation in terms of cosmology. Enclosing walls are rare and never placed in reference to a single mound; gates and avenues equally so.13 These are profound differences, quite outranking any similarities that may be seen in the individual structures. Is this a case of divergence from concepts of planning that were formerly similar? On the Asiatic side the central principle is fundamental, I believe, going back, it is said, to the ziggurat of Mesopotamia. On the American side the tangential principle seems also to be rooted in the past. The plans of La Venta and Chavin de Huantar differ markedly from the classic Maya assemblage but not in the direction of the centralized temple-mountain concept of Asia. In both cases the largest structure— pyramid at La Venta, castillo at Chavin— dominates the composition by size alone, not by a central position in relation to other structures of the assemblage. Sculpture The Olmec and Chavin cultures also appear to have priority in the production of monumental sculpture. 14 Again there is little similarity in the forms or the techniques of carving. Relationships of style present intriguing problems. Many Classic and Postclassic Maya forms have their prototypes in La Venta and other Olmec sites, 12 Krickeberg (1950) states with assurance that Maya pyramids symbolize the celestial mountain, but in the same paper he has already pointed out that symbols diffuse but their meanings change. 13 The saches of Yucatan and their "gates" seem to have no relationship to any specific pyramidal structures. 14 A statement made possible by the lack of chronological information about San Agustín and other sites with monumental sculpture in Colombia.

TRANSPACIFIC CONTACTS I N P R E - C O L U M B I A N CIVILIZATIONS

including some that have figured in transpacific writings. For example, Ekholm compared atlantean figures from Chichen Itza and the great stupa of Sanci, India, admitting that the case was somewhat complicated by the existence of an Olmec example from San Lorenzo Tenochtitlan (1953, p. 81, fig. 16). What he failed to point out was the uncanny resemblance between the little potbellied old men of San Lorenzo and Sanci, a good deal closer than between the latter and those of Chichen. The west gate of Sanĉi, where the figures are located, is said to date from the early first century A.D. (Zimmer, 1955, pl. 20, caption), almost certainly too late to have inspired the San Lorenzo sculpture. The theoretical implications of this parallel are not insignificant. The Olmec culture is held by diffusionists to be the result of Chinese contacts in a period necessarily antedating Indian expansion into southeast Asia. The San Lorenzo-Sanci parallel must be fortuitous or it makes nonsense of this hypothesis. Metallurgy The exquisite gold ornaments in the Chavin style that have been found from time to time (Lothrop, 1941, 1951) have been overlooked in the transpacific controversy. The artistic quality and design of these possibly earliest examples of goldworking in the New World, do not seem to have been surpassed in any subsequent periods, and this is what one might expect if the techniques, if not the actual artisans, were Asiatic. The objects, however, do not specifically figure in Heine-Geldern's monograph (1954), from which I conclude that he was unable to find close Chinese parallels. A greater difficulty for the diffusionist position, it seems to me, is the complete lack of metals in the early temple centers to the north. With one or two trifling exceptions metallurgy comes into Mesoamerica about the end of the Classic period. For some regions lack of raw materials may have been a factor, but this is small comfort

to the diffusionist who is trying to show that the impetus to civilization was in the form of direct and repeated contacts with metalusing cultures of the Far East. If, on the other hand, he is content to show that these contacts operated only indirectly on Mesoamerica via South America (a position that is beset by other difficulties that we cannot go into here), it is not less incredible that metallurgy, or at least objects of metal, were not relayed along with other elements of civilization of alleged Asiatic provenience. Hieroglyphic Writing, and the Calendar

Numeration,

One might expect that a decisive impact of one civilization on another would be most clearly revealed in the domain of intellectual culture. The outstanding archaeological expressions in this domain are, of course, the hieroglyphic inscriptions of Mesoamerica and what they reveal of an extraordinary preoccupation with the passage of time—a limited but highly characteristic body of mathematical and astronomical knowledge. The complete lack of a comparable development in Peru is one of the most striking differences between the two nuclear areas. If this body of knowledge derived from an Asiatic source it ought to be possible to prove it. Yet it seems that research in this field has been so far remarkably barren of results. However, I must point out that the diffusionist looks at this question in an entirely different way. For him the possibility of anything so complex as a system of writing being invented more than once simply does not exist. For example, in a recent paper on the perennial question of the origin of the Easter Island script, Heine-Geldern asks (freely translated): "Can we admit for a moment that Mexican and Maya writing were invented in complete independence without any suggestion from the systems of writing in the Old World? . . . What proofs do we have of this?" (1956-57, p. 19). I 311

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

ask, What sort of proofs does he demand? In a historical question how does one prove that something did not happen otherwise than by pointing to the lack of evidence that it did? It is not clear to me that HeineGeldern accepts this burden of proof, but he does indicate what sort of evidence he would bring to bear. The earliest examples of hieroglyphic writing, he holds, are the glyphs associated with the danzantes of Monte Alban I. The style of these remarkable bas-reliefs, he promises to show in a forthcoming paper, is Asiatic. Their date, possibly as early as the 8th century B.C., indicates as source the maritime states of south China, whose archaeology is so little known that direct comparisons cannot be made. Since the Monte Alban script cannot have been indigenous, it is not necessary to show its derivation from any specific Chinese script, but merely to indicate by a general stylistic comparison where that source would have to lie. It seems to me that from the standpoint of method this leaves something to be desired. Even if the stylistic affinity of the danzantes was beyond question, it would not preclude the possibility that the glyphs had a totally different history. Consider for a moment that it is by no means certain that these are the earliest glyphs in Mesoamerica. Pedantic as it may appear, I submit that the way to prove the relationship of one script to another is to compare the scripts. Great Art Styles A considerable share of transpacific "evidence" rests on comparisons in the domain of style. Close stylistic parallels are triumphantly proclaimed by the diffusionists as clinchers and greeted by the isolationists in embarrassed silence. Few attempts are made to consider the theoretical implications involved in such comparisons. It may make things clearer to begin with an extreme example, although it has nothing to do with early temple centers. Ekholm (1950) presents a very telling comparison 312

between a Marajoara pot from the mouth of the Amazon dating about A.D. 1200-1500 (Meggers and Evans, 1957, fig. 149) and a bronze drum of the Shang period which must date some time before 1028 B.C. (Grousset, 1959, p. 315). Here we have two objects decorated in a remarkably similar way and separated by a prodigious gap of time and space and we are asked to believe that they "hint at a common artistic tradition." Ekholm does not furnish any theory to account for the supposed relationship. For this we may turn to others. Thus Covarrubias (1954, p. 32): "The time difference is really no obstacle to the soundness of the theory. On the one hand the true age of American civilization is still undetermined. On the other there is a great probability that the cultural waves that started from eastern Asia between 1200 and 200 B.C. spread gradually across the Pacific by successive infiltration of styles from one place to another from various sources and at various times, a process that must have taken very long. What traveled, consequently, was the styles acquired second-hand by different peoples, not the actual objects from Asia." One of the way-stations Covarrubias had in mind was the Marquesas where, according to Heine-Geldern (1937), elements of the Shang style were preserved down to the 19th century.15 But style is a complex phenomenon like culture itself. One style can be influenced by another with a result that is unlike either, but that a style can be taken over by an alien people without serious modification is a doubtful proposition. And that it can be preserved in essential purity for thousands of years is another. How, it may be asked, does a style diffuse? Through movement of objects or of people capable of making them, or of showing 15

Heine-Geldern has more recently argued that the Chinese contacts with South America were by way of the North Pacific, bypassing Polynesia, where there could have been no way-stations to speak of.

TRANSPACIFIC CONTACTS IN P R E - C O L U M B I A N CIVILIZATIONS

other people how to make them? Covarrubias does not favor the first alternative because objects of proven Asiatic provenience have not been found in pre-Columbian America,16 but it is at least conceivable. If Shang bronzes found their way to and across South America to the mouth of the Amazon, they could have been imitated by Marajoara potters. Highly improbable but not impossible. But if they had to get the idea from people, who in turn got it from other people and so on for thousands of years, it is simply inconceivable that the end product would still be recognizable as Shang. Here is a problem that cannot be laid to rest by formulae such as "successive 16 There are numerous references to finds of Chinese objects in post-Columbian contexts, which shows that things do get transported once contacts have been established. See Meighan, 1950; Mason, 1951; Meighan and Heizer, 1952; Heizer, 1953; Camman, 1952; Borhegyi, 1955; Kamer, 1955; H. G. Smith, 1955. 17 We may take as a means of estimating the time gap the two pairs of examples figured by Ekholm (1950). I have no information about the precise date of the Chou bronzes but assume that they were made before 221 B.C., the end of the period of the Warring States (Grousset, 1959, p. 315). The Tajin examples are Yoke 9 and Palma 11 in Proskouriakoff's numbering (1954, figs. 1, 6 ) . Yoke 9 is classified in her Yoke Group A, which has features that she finds reminiscent of Olmec jades (p. 7 4 ) . She also calls attention to similar treatment on a carved mirror-back found in an Esperanza phase tomb at Kaminaljuyu crossdating with Teotihuacan HI, Monte Alban III, and Late Tzakol, in other words the end of the Early Classic period about A.D. 350 or 600, depending on the correlation used. Palma 11 is in the Tajin style of Proskouriakoff, the treatment "particularly suggestive of an advanced date" (p. 8 2 ) . The human sacrificial scene on the reverse, she says, is similar to that of the late ball-court sculpture at Tajin. Parenthetically, if Ekholm had illustrated both sides of this palma, the Chou resemblance would have been rather overshadowed by this gruesome presentation. Proskouriakoff says that so far there is no evidence that palmas were made before the Postclassic period. So the time gap between these objects and the Chou bronzes is 550 or 800 years, depending on the Maya-Christian calendrical correlation used. 18 A suggestion that the Tajin style is a persistence of an earlier style which may have been confined to woodcarving has the inestimable advantage that it cannot be confirmed or denied.

infiltration of styles" or "lost intermediaries." Such is the theoretical background of a comparison more pertinent to the present discussion, one that invariably figures in diffusionist writings, i.e. the similarity of ornamentation in the Classic Veracruz or Tajin style to that of ritual bronzes of the late Chou period in China. The time gap (some 500-800 years) 17 is not so great as in the Shang-Marajoara parallel. In this case, however, the interest is not primarily in the time gap, but in the indications that the Tajin style is grounded in the Olmec style, which is approximately contemporaneous with Middle Chou. If there were contacts with China at this time, either direct or by way of South America, they should be reflected in the Olmec style, not in Classic or Postclassic outgrowths from it. This is a problem for the diffusionist. I do not say that no such reflection can be seen, only that so far all suggested parallels have related to the later post-Olmec styles. The point that interests me is this: if there was a connection between the art of Chou and Preclassic Olmec, why did the style get to look more Chou-like with the passage of time? "Convergence" is a forbidden word in diffusionist literature, but I cannot think of any other explanation.18 Chavin, Olmec, and the Hypothesis of Early Chinese Contacts Most of the facets of the transpacific question commented on in this study, and others that might have been, are involved with the Chavin and Olmec cultures of the Early Formative or Middle Preclassic periods. These are proper testing grounds for theories of early Chinese contacts with America. In the most formidable monograph on the transpacific question yet to appear Heine-Geldern proposed an hypothesis of direct voyages to South America by merchant adventurers from the maritime kingdoms of southern China in the period from the end of the 8th century B.C. to some time in the 4th century B.C. (Heine-Geldern, 313

ARCHAEOLOGICAL FRONTIERS & EXTERNAL CONNECTIONS

RECAPITULATION

evidence for direct Asiatic influence is practically nonexistent. Except for pottery, I have not considered possibilities of direct diffusion via Bering Strait and North America, which may have been not insignificant. (2) Items having to do with the emergence of civilization, which I have equated with the rise of great temple centers in the Peruvian Early Formative and Mesoamerican Middle Preclassic periods. Here is a formidable transpacific hypothesis based on stylistic comparisons and conjectures about Chinese seafaring capabilities in the first millennium B.C. Whether this is "evidence" is a matter of opinion. I have done my best to indicate that something more substantial is required. In any case the importance of this hypothesis is not to be minimized. If I were a diffusionist, this is where I would concentrate my efforts. (3) Items, by far the largest number, that have to do with the fully developed Classic period civilizations of South and Middle America or with their less advanced contemporaries in between. In this category are included most of the famous parallels of transpacific literature, from Tylor's patolli-pachesi (Tylor, 1879; Erasmus, 1950) down to the recently reported Ecuadorian finds (Estrada, 1960). These I have not considered at all. They are fascinating problems of culture history, but from the point of view of this article they are of minor interest. We could concede them all without affecting the integrity of Nuclear American civilization in a strictly developmental sense.

For purposes of summary the issues that have figured in the controversy over transpacific diffusion may be grouped into three categories: (1) Items having to do with the Preformative and Early Preclassic cultures of Nuclear America in the long period that saw the beginnings of agriculture and settled village life. In this category the

My final conclusion is that Professor Koppers, in the passage cited at the beginning of this article, has greatly overstated the position. The diffusionists have not proved the unity of Far Eastern and American civilizations, and as for the lower levels of cultural development, they have not even made a case.

1954, pp. 383-87). The object of the voyages was the search for gold, the result was the Chavin culture. Krickeberg, I believe, was the first to point out in print that this hypothesis could be extended to the Olmec culture (1956, p. 575). Chavin and Olmec occupy similar developmental niches in their respective areas. Connections between them have been postulated (Porter, 1953) but not thoroughly explored. Chronological relationships are not very firm, but offer no serious difficulty. If I understand it correctly, the priority of Chavin over Olmec would be assumed in the expanded hypothesis because of the importance of gold as motivation for the voyages and the knowledge of goldworking in Chavin as evidence that they occurred. Influences from Chavin would have been responsible for a second florescent culture in Mesoamerica, the Olmec. Subsequent development on the foundation provided by these two affiliated cultures would have been fostered by continued contacts from China, and later from Southeast Asia, resulting in the classic civilizations of Peru and Mesoamerica. It is a grand hypothesis. Merely to state it gives one a vicarious thrill. But there are enormous difficulties, as I have attempted to show. Not the least is that Chavin and Olmec are not similar enough to have derived, one from the other, or both from the same Asiatic source.

314

TRANSPACIFIC CONTACTS I N PRE-COLUMBIAN

CIVILIZATIONS

REFERENCES Anderson, 1960 Barghoorn, Wolfe, and Clisby, 1954 Bird, 1951 Borhegyi, 1955 Braidwood, 1958, 1960 Briggs, 1951 Camman, 1952 Carter, 1950 Chard, 1958a, 1958b Chu'ü-hsun, 1959 Coe, M. D., 1957a, 1960b, 1961 Coedés, 1947, 1948 Collier, 1959 Covarrubias, 1954 Crane, 1956 Cruxent and Rouse, 1958-59 Dick, 1952 Drucker, Heizer, and Squier, 1959 Ekholm, 1950, 1953, 1955 Emory, 1943 Engel, 1957b Erasmus, 1950 Estrada, 1960 Evans, Meggers, and Estrada, 1959 Foshag and Leslie, 1955 Griffin, 1952b, 1958, 1960a Grousset, 1959 Heine-Geldern, 1937, 1952, 1954, 1956a, 1956b, 1956-57, 1958 Heizer, 1953 Ho, 1955 Hutchinson, Silow, and Stephens, 1947 Johnson, F., 1951

Kamer, 1955 Koppers, 1957 Krickeberg, 1950, 1956 Kurtz, Tucker, and Liverman, 1960 Libby, 1951, 1952 Lothrop, 1941, 1951 MacNeish, 1958, 1961 Mangelsdorf and Oliver, 1951 and Reeves, 1959 Mason, J. Α., 1951 Meggers and Evans, 1957 Meighan, 1950 and Heizer, 1952 Piña Chan, 1955 Porter, Μ. N., 1953 Proskouriakoff, 1954 Reichel-Dolmatoff, 1961 Ricketson and Ricketson, 1937 Ridley, 1960 Rouse, 1960b Sauer, 1952 Sears and Griffin, 1950 Shook, 1951 Smith, H. G., 1955 Stephens, S. G., 1947 Stonor and Anderson, 1949 Tolstoy, 1953, 1958b Tylor, 1879 Vaillant, 1930, 1935 Willey, 1958, 1960 et al., 1956 Zimmer, 1955

315

THIS PAGE INTENTIONALLY LEFT BLANK

R E F E R E N C E S AND I N D E X

THIS PAGE INTENTIONALLY LEFT BLANK

REFERENCES

ARLEGUI, J.

ACOSTA, J. R.

1957

Resumen de los informes de las exploraciones arqueológicas en Tula, Hgo., durante las IX y X temporadas, 1953-1954. An. Inst. Nac. Antr. Hist., vol. 9.

AGUILAR P., C.

1953

ALEGRÍA, R.

1951

E.

The ball game played by the aborigines of the Antilles. Amer. Antiquity, 16: 348-52.

ALESSIO ROBLES, V.

1927 1938

Bibliografía de Coahuila. Mexico. Coahuila y Texas en la época colonial. Mexico.

ALMADA, F.

1945

R.

Geografía del estado de Chihuahua. Chihuahua.

ALVARADO, P . DE

1924

An account of the conquest of Guatemala in 1524. Ed. and tr. by S. J. Mackie. Cortés Soe., no. 3. New York.

ALVAREZ RUBIANO, P.

1944

AMSDEN,

1928

A.

The Pinto Basin artifacts. In Campbell and Campbell, 1935, pp. 33-51.

1916 1960

The evolution of domestication. Tax, 1960, 2: 67-84.

ANDERSSON, J.

1932

AVELEYRA ARROYO DE ANDA, L.

1951

1956

1962

G.

ARCHAEOLOGICAL SALVAGE PROGRAM F I E L D OFFICE

1958

Appraisal of the archaeological resources of Diablo Reservoir, Val Verde County, Texas. Austin.

H.

1952a A fluted point from central Baja California. Amer. Antiquity, 17: 2 6 2 63. 1952b A primitive food preparation technique in Baja California. SW. Jour. Anthr., 8: 36-39. 1959 The central desert of Baja California: demography and ecology. IberoAmer., no. 42.

In

Hunting magic in the animal style. Mus. Far Eastern Antiquities, Bull. 4, pp. 221-317.

Waterfall Cave, southern Chihuahua, Mexico. Amer. Antiquity, 26: 2 7 0 74.

ASCHMANN,

ANDERSON, E.

1960

Relation of the journey . . . Viscaino (1602). In Bolton, 1916.

ASCHER, R., AND F . J. C L U N E , J R .

Μ.

Archaeological reconnaissance in Sonora. SW. Mus. Papers, no. 1.

Α.

Late Pleistocene and Early Recent changes in land forms, climate, and archaeology in central Baja California. Univ. California Pub. Geog., 10: 201-318.

ASCENSIÓN, A. DE LA

Pedrarias Dávila. Inst. Gonzalo Fernández de Oviedo. Madrid.

AMSDEN, C.

1935

1957

Η.

Historia de la Compañía de Jesus en Nueva España. 3 vols. Mexico.

Crónica de la Provincia de N.S.R.S. Francisco de Zacatecas. Mexico.

ARNOLD, Β.

Retes, un depósito arqueológico en las faldas del Irazu. San José.

ALEGRE, F. J.

1841

1737

Reconocimiento arqueológico en la zona de la Presa Internacional Falcón, Tamaulipas y Texas. Rev. Μex. Estud. Antr., 12: 31-59. The second mammoth and associated artifacts at Santa Isabel Iztapan, Mexico. Amer. Antiquity., 22: 1 2 28. Antigüedad del hombre en Mexico y Centroamérica: catálogo razonado de localidades y bibliografía selecta (1867-1961). Cuad. Inst. Hist., Ser. Antr., no. 14.

, M. MALDONADO-KOERDELL, AND P. MARTÍNEZ DEL RÍO

1956

Cueva de la Candelaria. Nac. Antr. Hist., no. 5.

Mem. Inst.

BAEGERT, J. J.

1942

Noticias de la peninsula americana de California. Tr. by Pedro Hendrichs. Mexico. 319

REFERENCES BALFOUR GOURLAY,

1940 BALSER,

1953 1955

C.

El jade precolombino de Costa Rica. Mus. Nac. Costa Rica. A fertility vase from the old line, Costa Rica. Amer. Antiquity, 20: 384-87.

BANCROFT, Η.

1882 1886

W.

Man and elephant in Central America. Man, 40: 86-88.

BANDELIER, A.

F.

1954

BEAL, C.

1948 BEALS, R.

Κ. W O L F E , AND Κ.

Η.

Fossil maize from the Valley of Mexico. Bot. Mus. Leafl., Harvard Univ., 16: 229-40. Conociendo a Coahuila, los laguneros o irritilas su cultura y sus restos. Univ. Coahuila. Saltillo.

BARTLETT, J. R.

1854

Personal narrative of explorations and incidents in Texas, New Mexico, California, Sonora and Chihuahua, connected with the United States and Mexican Boundary Commission, during the years 1850, '51, '52, and '53. New York.

BATRES, L.

1903

Visita a los monumentos arqueológicos de la Quemada, Zacatecas, Mexico. Mexico.

BAUDEZ, C.

1959 1962

320

L.

The comparative ethnology of northern Mexico before 1750. IberoAmer., no. 2. 1942 Shell mounds and other sites in Sonora and northern Sinaloa. Notebook Soc. Amer. Archaeol., 2: 38-40. 1944a Northern Mexico and the Southwest. In El Norte de Mexico, pp. 191-98. 1944b Relations between Mesoamerica and the Southwest. Ibid., pp. 245-51. BELDING, L.

1885 1960

The Pericue Indians. Sci., vol. 1, no. 4.

F.

Nuevos aspectos de la escultura lítica en el territorio Chorotega. 33rd Int. Cong. Amer. (San Jose), 2: 286-95. Recherches archéologiques dans la

West

Amer.

E.

Evidence of a fluted point tradition in Ecuador. Amer. Antiquity, 26: 10206.

BENAVIDES

1954

See Forrestal, 1954.

BENNETT, W.

1946

BARRAGÁN, Ε., C. CÁRDENAS, AND C. VALDÉS

1960

H.

Reconnaissance of the geology and oil possibilities of Baja California, Mexico. Geol. Soc. Amer., Mem. 31.

BELL, R.

The mission frontier in Sonora, 16201687. U. S. Catholic Hist. Soc., Monogr. Ser., 26.

BARGHOORN, E. S., Μ. CLISRY

Archaeological sequences in northwestern Costa Rica. Acts 34th Int. Cong. Amer. (Vienna), pp. 366-73.

1932

F.

1880-85 Portfolio. Drawings without copy. Vatican Library, Vat. Lat. 14112-16. Microfilm. 1890a The ruins of Casas Grandes. The Nation, Aug. 28 and Sept. 4. 1890b Final report of investigations among the Indians of the southwestern United States, Part 1. Papers Archaeol. Inst. Amer., Amer. Ser., 3. 1892 Final report of investigations among the Indians of the southwestern United States, Part 2. Ibid., 4. 1955

AND M . D . C O E

1962

Η.

The native races. Vol. 1. History of the north Mexican states and Texas. Vol. 1, 1531-1800. In The works of Hubert Howe Bancroft, vol. 15.

BANNON, J.

1963

vallée du Tempisque, Guanacaste, Costa Rica. 34th Int. Cong. Amer. (Vienna). Cultural development in lower Central America. In Meggers and Evans, 1963, pp. 45-54.

1948

C.

The archeology of Colombia. In Handbook of South American Indians, 2: 823-50. A reappraisal of Peruvian archaeology. Soc. Amer. Archaeol., Mem. 4.

BENZONI, G.

1857 BERLIN,

1940 BIRD, J.

1948 1951

History of the New World. Ed. and tr. by W. H. Smyth. Hakluyt Soc. London. H.

Relaciones precolombinas entre Cuba y Yucatan. Rev. Μex. Estud. Antr., 4: 141-60. B.

Preceramic cultures in Chicama and Viru. In Bennett, 1948, pp. 21-29. South American radiocarbon dates. In F. Johnson, 1951, pp. 37-49.

BLACKISTON, A.

1905

H.

Cliff dwellers of northern Mexico. Records of the Past, 4: 355-61. 1906a Casas Grandian outposts. Ibid., vol. 5.

REFERENCES

1906b Ruins on the Cerro de Montezuma. Amer. Anthr., 8: 256-61. 1909 Recently discovered cliff dwellers of the Sierra Madres. Records of the Past, 8: 20-32. 1910 Recent discoveries in Honduras. Amer. Anthr., 12: 536-41.

1955 1959

BLOM, F.

1934 BOGGS, S.

Commerce, trade and monetary units of the Maya. Smithsonian Ann. Rept. 1934, pp. 423-40. H.

1943a Notas sobre las excavaciones en la hacienda "San Andres," Departamento de La Libertad. Tzunpame, 3: 104-26. 1943b Observaciones respecto a la importancia de "Tazumal" en la prehistoria salvadoreña. Ibid., 3: 127-33. 1944a A human-effigy pottery figure from Chalchuapa, El Salvador. Carnegie Amer. Inst. Wash., Notes Middle Archaeol. Ethnol., no. 31. 1944b A preconquest tomb on the Cerro del Zapote, El Salvador. Ibid., no. 32. 1944c Excavations in central and western El Salvador. In Longyear, 1944, pp. 53-72. 1945a Informe sobre la tercera temporada de excavaciones en las ruinas de "Tazumal." Tzunpame, 5: 33-45. 1945b Archaeological material from the Club Internacional, El Salvador. Carnegie Inst. Wash., Notes Middle Amer. Archaeol. Ethnol., no. 60. 1949 Tlaloc incensarios in the Baratta collection, El Salvador. Ibid., no. 94. 1950 "Olmec" pictographs in the Las Victorias group, Chalchuapa archaeological zone, El Salvador. Ibid., no. 99. BOLTON, Η.

1916 1919 1936

Ε.

Spanish explorations in the Southwest, 1542-1706. New York. Kino's historical memoir of Pimeria Alta. 2 vols. Cleveland. Rim of Christendom. New York.

BORBOLLA, S. A. R. DE LA, AND L. AVELE YRA ARROYO DE ANDA

1953

A Plainview point from northern Tamaulipas. Amer. Antiquity, 18: 392-93.

BORHEGYI, S.

1960

BOSQUE, F. DEL

1675

Diary.

BOURKE, J.

1886

In Bolton, 1916.

G.

An Apache campaign in the Sierra Madre. New York.

BOVALLIUS, C.

1886

Nicaraguan antiquities.

BOYD, W.

1950 1959

Stockholm.

C.

Genetics and the races of man. Boston. A possible example of the action of selection in human blood groups. Jour. Medical Education, 34: 398-99.

AND L. G. BOYD

1937

Blood grouping tests on 300 mummies. Jour. Immunology, 32: 3 0 7 19.

BRADFIELD,

1923 1931

W.

Preliminary report on excavating at Cameron Creek site. El Palacio, vol. 15, no. 5. Cameron Creek Village: a site in the Mimbres area in Grant County, New Mexico. School Amer. Research, Monogr. 1.

BRAIDWOOD, R. J.

1958 1960

Near Eastern prehistory. Science, 127: 1419-30. The agricultural revolution. Sci. Monthly, 203: 131-48.

AND G . R . W l L L E Y

1962

Courses toward urban life. Fund Pub. Anthr., no. 32.

BRAINERD, G.

1953 1932 1935 1936

Viking

W.

A cylindrical stamp from Ecuador. Masterkey, 27: 14-27.

BRAND, D.

F.

1951a A study of three-pronged incense burners from Guatemala and adjacent areas. Carnegie Inst. Wash., Notes Middle Amer. Archaeol. Ethnol., no. 101. 1951b Further notes on three-pronged in-

cense burners and rim-head vessels in Guatemala. Ibid., no. 105. Chinese figurines in Mesoamerica. Amer. Antiquity, 20: 286-88. Pre-Columbian cultural connections between Mesoamerica and Ecuador. Tulane Univ., Middle Amer. Research Records, vol. 2, no. 6. Pre-Columbian cultural connections between Mesoamerica and Eucador: addenda. Ibid., vol. 2, no. 7.

D.

Historical geography of northwest Chihuahua. Doctoral dissertation, Univ. California. The distribution of pottery types in northwestern Mexico. Amer. Anthr., 37: 287-305. Notes to accompany a vegetation map of northwest Mexico. Univ. New Mexico Bull., Biol. Ser., vol. 4, no. 4. 321

REFERENCES

1937 1938 1939 1943 1944

The natural landscape of northwestern Chihuahua. Univ. New Mexico Bull, Geol. Ser., vol. 5, no. 2. Aboriginal trade routes for sea shells in the Southwest. Yearbook Assoc. Pacific Coast Geog., 4: 3-10. Notes on the geography and archaeology of Zape, Durango. In Brand and Harvey, 1939, pp. 75-105. The Chihuahua culture area. Univ. New Mexico Quar., 6: 115-58. Archaeological relations between northern Mexico and the Southwest. In El Norte de Mexico, pp. 199-202.

AND F. E. HARVEY,

1939

1881 1882

1959

1944 1946

O.

Mexican influence upon the Indian cultures of the Southwestern United States in the sixteenth and seventeenth centuries. In The Maya and their Neighbors, pp. 341-48. On the Pueblo IV and on the Katchina-Tlaloc relations. In El Norte de Mexico, pp. 241-44. Archaeology of Alkali Ridge, southeastern Utah. Papers Peabody Mus., Harvard Univ., vol. 21.

BRIGGS, L.

1951

1960

322

1871

Plant materials from a cave on the Amer. Río Zape, Durango, Mexico. Antiquity, 27: 356-69. L.

Races of maize in the West Indies.

Ν.

The narrative of Alvar Núñez Cabeza de Vaca. Tr. by T. Buckingham Smith. Washington.

CAMMAN, S. V.

1952

R.

A chínese soapstone carving from Yucatan. Amer. Antiquity, 18: 6 8 69.

CAMPBELL, E. W. C , AND W. H. CAMPBELL

1935

The Pinto Basin site. Mus. Papers, no. 9.

CAMPBELL, Τ.

1958

Southwest

Ν.

Origin of the mescal bean cult. Amer. Anthr., 60: 156-60. Archaeology of the central and southern sections of the Texas coast: a review of Texas archaeology, Part 1. Bull. Texas Archaeol. Soc, 29: 1 4 5 75.

1960

CANBY, J. S.

1949 1951

Excavations at Yarumela, Spanish H o n d u r a s . Doctoral dissertation, Harvard Univ. Possible chronological implications of the long ceramic sequence recovered at Yarumela, Spanish Honduras. In Tax, 1951, pp. 79-85.

CANDOLLE, A. DE

1959

Origin York.

of cultivated plants.

New

CANO, F.

1873

W.

Reports of Japanese vessels wrecked in the north Pacific, from the earliest records to the present time. Proc. California Acad. Sci., 6: 50-66.

BROWN, W.

Preliminary report on an archaeological site, district of Chame, province of Panama. Panama ArchaeoL, vol. 2, no. 1.

CABEZA DE VACA, Α.

P.

BROOKS, R. H., L. KAPLAN, H. C. CUTLER, AND T. W. WHITAKER

1962

Η.

1959

The ancient Khmer empire. Trans. Amer. Phil. Soc, vol. 41, pt. 1.

BROOKS, C.

1875

A.

Excavations at Nantack Village, Point of Pines, Arizona. Univ. Arizona, Anthr. Papers, no. 1.

BREW, J.

1940

F.

Archaeological researches in Nicaragua. Smithsonian Contrib. Knowledge, no. 25. Report on explorations in Central America in 1881. Smithsonian Ann. Rept. 1882.

BRETERNITZ, D.

BULL, Τ.

eds.

So live the works of men. Seventieth anniversary volume honoring Edgar Lee Hewett. Univ. New Mexico and School Amer. Research.

BRANSFORD, J.

Nat. Acad. Sci., Nat. Research Council, Pub. 792.

Testimonio del descubrimiento y posesión de la laguna del Nuevo Mexico, hecho por Francisco Cano, teniente de alcalde mayor de las minas de Mascipil en la Nueva Galicia, 1568. In Col. Doc. Inéditos del Archivo de las Indias, by Pacheco y Cárdenas. Mexico.

CAREY, H.

1931 1951

1954

A.

An analysis of the northwestern Chihuahua culture. Amer. Anthr., 33: 325-74. Review of The Pendleton ruin, Hidalgo County, New Mexico, by A. V. Kidder, H. S. and C. B. Cosgrove. Amer. Antiquity, 17: 156-57. Grant No. 1597 (1953), $750,000. The ancient Indian culture centering in the Casas Grandes Valley, north-

REFERENCES

1955

western Chihuahua, Mexico. Year Book Amer. Phü. Soc, pp. 313-16. Grant No. 1779, $750,000. The Casas Grandes culture, Chihuahua, Mexico. Ibid., pp. 314-16.

CARTER, G.

1950 1957

1947

CLAVIGERO, F.

1937

F.

Plant evidence for early contacts with America. SW. Jour. Anthr., 6: 1 6 1 82. Pleistocene man at San Diego. Baltimore.

CASO, A.

1927

the Indians. Tr. from original Italian by C. Cullen. 2 vols. London.

Las ruinas de Tizatlan, Tlaxcala. Rev. Μex. Estud. Hist., 1 ( 4 ) : 39-72. Calendario y escritura de las antiguas culturas de Monte Alban. Mexico.

CLUNE,

1960

Urnas de Oaxaca. Antr. Hist., no. 2.

Mem. Inst.

Nac.

1779 M. 1957a

COE,

1957b

AND OTHERS

1946

¿Conocieron la rueda los indígenas mesoamericanos? Cuad. Amer., 25: 1-15.

CASON, J. F.

1952

Report on archaeological salvage in Falcon reservoir, season of 1952. Bull. Texas Archaeol. and Paleontol. Soc, 23: 218-59.

1960a 1960b

1961

CERECEDA, A. DE

1926

Relación. 29.

CHAMPE, J.

1946

1962a

L.

CHAPMAN, Κ.

1923

In Lothrop, 1926, 1: 2 8 -

Ash Hollow cave, a study of stratigraphic sequence in the central Great Plains. Univ. Nebraska Studies, n.s., no. 1.

CHARD, C.

Art

and

1942

G.

1787

W. 1955

COE,

1957

The royal cemetery of the Yin dynasty at Anyang. Nat. Taiwan Univ., Dept. Archaeol. and Anthr., Bull. 13, 14. Tapei.

1959

CLAVIGERO, D. F.

S.

The history of Mexico collected from Spanish and Mexican historians from manuscripts, and ancient paintings of

Textiles and matting from Waterfall Cave, Chihuahua. Amer. Antiquity, 26: 274-77. The unfortunate Englishmen. London. D. The Khmer settlement pattern: a possible analogy with that of the Maya. Amer. Antiquity, 22: 409-10. Preclassic cultures in Mesoamerica: a comparative survey. Papers Kroeber Anthr. Soc, 17: 7-37. A fluted point from highland Guatemala. Amer. Antiquity, 25: 412-13. Archaeological linkages with North and South America at La Victoria, Guatemala. Amer. Anthr., 62: 3 6 3 93. La Victoria: an early site on the Pacific coast of Guatemala. Papers Peabody Mus., Harvard Univ., Vol. 53. Preliminary report on archaeological investigations in coastal Guanacaste, Costa Rica. Acts 34th Int. Cong. Amer. (Vienna), pp. 358-65. An Olmec design on an early Peruvian vessel. Amer. Antiquity, 27: 579-80. Olmec and Chavin: rejoinder to Lanning. Ibid., 29: 101-104.

AND C . F . BAUDEZ

1961

Penguin

What happened in history. Books.

CHU'Ü-HSUN, KAO

1959

1963

S.

1958a Organic tempering in northeast Asia and Alaska. Amer. Antiquity, 24: 193-94. 1958b An outline of the prehistory of Siberia. Part 1: The premetal period. SW. Jour. Anthr., 14: 1-33. CHILDE, V.

1962b

Μ.

Casas Grandes pottery. Archaeol, 16: 25-34.

D.

COCKBURN, J .

AND I. BERNAL

1952

J.

The history of (Lower) California. Tr. by S. E. Lake and A. A. Gray. Stanford.

The zoned bichrome period in northwestern Costa Rica. Ibid., 26: 5 0 5 15. R. Excavations in El Salvador. Bull. Univ. Pennsylvania Mus., 19: 15-21. A distinctive artifact common to Haiti and Central America. Amer. Antiquity, 22: 280-82. Piedras Negras archaeology: artifacts, caches, and burials. Mus. Monogr., Univ. Pennsylvania, no. 4.

COEDÉS, G.

1947 1948

Pour mieux comprendre Angkor. Libraire d'Amerique et d'Orient. Paris. Les états hindouisés dTndochine et

323

REFERENCES

d'Indonésie. In Histoire du Monde, vol. 8, ed. by M. E. Cavaignac. Paris. COLLIER,

1959

D.

Agriculture and civilization on the coast of Peru. Paper presented to annual meeting, Amer. Anthr. Assoc, Mexico City. Mimeographed. Chicago.

Papers Peabody Mus., Harvard Univ., vol. 24, no. 2. COSGROVE, H. S., AND C. B. COSGROVE

1932

COSTA RICA - PANAMA ARBITRATION

1913

COLLINSON, J.

1870

The Indians of the Mosquito Territory. Mem. Anthr. Soc. London, 3: 148-56. Vida del Almirante Don Cristobal Colón. Fondo de Cultura Económica. Mexico.

COLUMBUS, CHRISTOPHER

1930

The voyages of Christopher Columbus, being the journals of his first and last voyages, to which is now added the account of his second voyage written by Andrés Bernáldez, now newly translated and edited, with introduction and notes, by Cecil Jane. London.

COMMAILLE, J.

1913

Notes sur la décoration cambodgienne. Bull. I'Ecole Française d'Extrème-Orient, 13 ( 3 ) : 1-38. Hanoi.

CONZEMIUS, E .

1928 1932

Los Indios Payas de Honduras. Estud. Geog., Hist., Etnog., y Ling. Paris. Ethnographical survey of the Miskito and Sumu Indians of Honduras and Nicaragua. Smithsonian Inst., Bur. Amer. Ethnol, Bull. 106.

1956

The extent and significance of disease among the Indians of Baja California. Ibero-Amer., no. 12.

COOK DE LEONARD, C.

1954

Dos extraordinarias vasijas del museo de Villa Hermosa (Tabasco). Yan, 3: 83-104.

COOMARASWAMY, Α .

1928-31 Yakshas. Coll, no. 80.

Smithsonian

Quito.

1947

1954 1957

1954 1955

Diferentes tipos de tumbas prehispánicas en Nayarit. Yan, 3: 46-50. Tumba de el Arenal, Etzatlan, Jal. Inst. Nac. Antr. Hist., Informe 3.

COSGROVE, C. B.

1947

324

Caves of the Upper Gila and Hueco areas, in New Mexico and Texas.

Mexico south: the isthmus of Tehuantepec. New York. The eagle, the jaguar and the serpent: Indian art of the Americas. New York. Indian art of Mexico and Central America. New York.

CRANE, H.

1956

R.

University of Michigan radiocarbon dates, I. Science, 124: 664-72.

AND J. B. GRIFFIN

1958a University of Michigan radiocarbon dates, II. Ibid., 127: 1098-1105. 1958b University of Michigan radiocarbon dates, III. Ibid., 128: 1117-23. 1960 University of Michigan radiocarbon dates, V. Amer. Jour. Sci., Radiocarbon Suppl., 2: 31-48. 1961 University of Michigan radiocarbon dates, VI. Ibid., 3: 105-25. CROOK, W. W., JR., AND R. K. HARRIS

1957

Hearths and artifacts of early man near Lewisville, Texas, and associated faunal materials. Bull. Texas Archaeol. Soc, 38: 7-97.

CRUXENT, J. M., AND I. ROUSE

1958-59 An archeological chronology of Venezuela. 2 vols. Pan Amer. Union, Social Sci. Monogr., no. 6. CURILLOS, J. C.

1955

Misc.

CORONA NUÑEZ, J.

Trabajas del Instituto Ecuatoriano de Antropología y Geografía. Bol. Informaciones Cien. Nac, 8: 582-608.

COVARRUBIAS, M .

COOK, S. F.

1937

Documents annexed to the argument of Costa Rica. Vol. 1.

COSTALES SAMANIEGO, A.

COLÓN, Η.

1947

The Swartz ruin: a typical Mimbres site in southwestern New Mexico. Ibid., vol. 15, no. 1.

Tumaco (notas a r q u e o l ó g i c a s ) . Ministerio E d u c , Dept. Extensión Cultural. Bogota.

CUTLER, H.

1960

AND

1961

C.

Cultivated plant remains from Waterfall Cave, Chihuahua. Amer. Antiquity, 26: 277-79. T.

W.

WHITAKER

History and distribution of the cultivated cucurbits in the Americas. Ibid., 26: 469-85.

REFERENCES DADE, P. L.

1959

D I C E , L.

Tomb burials in southeastern Veraguas. Panama Archaeol., vol. 2, no. 1.

DAHLGREN, B., AND J. ROMERO

1951

La prehistórica Baja-California, redescubrimiento de pinturas rupestres. Cuad. Amer., 4: 3-28.

DAMPIER,

1699

W.

A new voyage around the world. (4th ed.) 2 vols. London.

DANSON, E.

1941

1959

T.

Further notes on clay human figurines in the western United States, no. 71. Univ. California Archaeol. Survey, Rept. 48, pp. 16-31.

DECORME, G.

1941

La obra de los Jesuítas mexicanos durante la época colonial 1572-1767. Vol. 2: Las misiones. Antigua Librería. Mexico.

DELLA SANTA, E.

1959

Les Cupisniques et I'origine des Olmèques. Rev. Univ. Bruxelles, 5: 1-24.

1952

Comments on radiocarbon dates from Mexico. Mem. Soe. Amer. Archaeol, 8: 33-36.

DEUEL, T.,

1952

1899

1905

Aboriginal Indian remains in Jamaica. Jour. Inst. Jamaica, 2 ( 4 ) : 1-52. Kingston.

1941

Groningen radiocarbon Science, 127: 123-37.

D'HARCOURT,

1942

1950 1951 1955 1956

DIAMOND, S.,

1960

R.

ed.

Culture in history: essays in honor of Paul Radin. New York.

DÍAZ DEL CASTILLO, B.

1927

The discovery and conquest of Mexico. 2 vols. New York.

Wash.,

C.

Painted stone slabs of Point of Pines, Arizona. Amer. Antiquity, 16: 5 7 65. The Babocomari village site on the Babocomari River, southeastern Arizona. Amerind Found., no. 5. Two Cerro Guamas Clovis fluted points from Sonora, Mexico. Kiva, 21 ( 1 , 2 ) : 13-15. The Upper Pima of San Cayetano del Tumacacori. Amerind Found., no. 7.

AND H. C. CUTLER

1958

The Reeve ruin of southeastern Arizona: a study of a prehistoric western Pueblo migration into the middle San Pedro valley. Ibid., no. 8.

DIXON, K.

1959

dates.

Archéologie de la province d'Esmeraldas, Equateur. Jour. Soc. Amer. Paris, 34: 61-200.

El Salvador. Carnegie Inst. Year Book 40.

DIPESO, C.

DEVRIES, H.

1958

Rapport sur une mission scientifique dans la Basse-Californie. Nouvelles Archives des Archives Scientifiques, 9: 1-53. Paris. Anciennes sepultures indigenes de la Basse-Californie meridionale. Jour. Soc. Amer. Paris, n.s., vol. 2.

DlMICK, J.

DEURDEN, J. E.

1897

The Claypool site: a Cody complex site in northeastern Colorado. Amer. Antiquity, 26: 223-35.

DIGUET, L.

ed.

Hopewellian communities in Illinois. Illinois State Mus. Sci. Papers, vol. 5, no. 3.

W.

Evidences of early man in Bat Cave and on the plains of San Augustin, New Mexico. In Tax, 1952, pp. 158-63.

AND B. MOUNTAIN

1960

D E TERRA, H.

1951

R.

The biotic provinces of North America. Univ. Michigan Press.

DICK, H.

B.

An archaeological survey of the Santa Cruz River valley from the headwaters to the town of Tubac in Arizona. MS report on file at Arizona State Mus.

DAVIS, J.

1943

A.

Ceramics from two preclassic periods at Chiapa de Corzo, Chiapas, Mexico. Papers New World Archaeol. Found., no. 5.

DOCKSTADER, F. J.

1961

A figurine cache from Kino Bay, Sonora. In Lothrop and others, 1961, pp. 182-91.

DOCUMENTOS PARA LA HISTORIA DE MEXICO

1853-57

Mexico.

DRIVER, H. E., AND W. C. MASSEY

1957

Comparative studies of North American Indians. Trans. Amer. Phil. Soc, 47 (pt. 2 ) : 165-456. 325

REFERENCES

1955

DRUCKER, P.

1943a Ceramic sequences at Tres Zapotes, Veracruz, Mexico. Smithsonian Inst., Bur. Amer. Ethnol., Bull. 140. 1943b Ceramic stratigraphy at Cerro de las Mesas, Veracruz, Mexico. Ibid., Bull. 141. 1952 La Venta, Tabasco: a study of Olmec ceramics and art. Ibid., Bull. 153. , R. F. HEIZER, AND R. J. SQUIER

1959

Excavations at La Venta, Tabasco, 1955. Ibid., Bull. 170.

DUCRUE,

1765

N.

Description of Crouley, n.d.

DUNNE, P.

1940 1944 1948

M.

W.,

D.

KRIEGER,

AND J.

B.

Kultur und Siedlung der Randvolker Chinas. Leyden. rates.

1944

F.

El Norte de Mexico y el Sur de Estados Unidos. Tercera reunión de mesa redonda sobre problemas antropológicos de Mexico y Centro America. Mexico.

EMORY, Κ.

1943

Current

Results of an archaeological survey of Sonora and northern Sinaloa. Rev. Mex. Estud. Antr., 3: 7-10. 1940 The archaeology of northern and western Mexico. In The Maya and their Neighbors, pp. 320-30. 1942 Excavations at Guasave, Sinaloa, Mexico. Amer. Mus. Natural Hist., Anthr. Papers, vol. 38, pt. 2. 1944 Excavations at Tampico and Panuco in the Huasteca, Mexico. Ibid., vol. 38, pt. 5. 1946a The probable use of stone yokes. Amer. Anthr., 48: 593-606. 1946b Wheeled toys in Mexico. Amer. Antiquity, 11: 223-28. 1950 Is American Indian culture Asiatic? Natural Hist., 59: 344-51, 382. 1953 A possible focus of Asiatic influence in the late classic cultures of Mesoamerica. Mem. Soc. Amer. Archaeol., 9: 72-89.

326

EL NORTE DE MEXICO

S.

Neolithic diffusion Anthr., 2: 71-102.

EKHOLM, G.

1939

A.

W.

EDMONSON, M.

1961

O'-

The archaeological zone of Buena Vista, Huaxcama, San Luis Potosí, Mexico. Amer. Antiquity, 13: 1 5 33.

EBERHARD,

1942

In

Pioneer Black Robes on the west coast. Univ. California Press. Pioneer Jesuits in northern Mexico. Ibid. Early Jesuit missions in Tarahumara. Ibid.

Du SOLIER, GRIFFIN

1947

California.

The new orientation toward problems of Asiatic-American relationships. In Meggers and Evans, 1955, pp. 9 5 109. 1958 Regional sequences in Mesoamerica and their relationships. In Willey, Vogt, and Palerm, eds., Middle American Anthropology, pp. 15-24. Pan Amer. Union, Social Sci. Monogr., no. 5. 1961 Puerto Rican stone "collars" as ballgame belts. In Lothrop and others, 1961, pp. 356-71. 1964a Possible Chinese origin of Teotihuacan cylindrical tripod pottery and certain related traits. Acts 35th Int. Cong. Amer. (Mexico), 1: 3 1 45. 1964b Transpacific contacts. In Jennings and Norbeck, eds., Prehistoric man in the New World, pp. 489-510.

ENGEL,

Ρ.

Polynesian stone remains. In Coon and Andrews, Studies in the anthropology of Oceania and Asia, pp. 9 21. Papers Peabody Mus., Harvard Univ., vol. 20. F.

1956

Curayacu, a Chavinoid site. Archaeology, 9: 98-105. 1957a Early sites on the Peruvian coast. SW. Jour. Anthr., 13: 54-68. 1957b Sites et établissements sans céramique de la cote Péruvienne. Jour. Soc. Amer. Paris, 46: 65-155. EPSTEIN, J.

F.

1957

Late ceramic horizons in northeastern Honduras. Doctoral dissertation, Univ. Pennsylvania. 1959 Dating the Ulua polychrome complex. Amer. Antiquity, 25: 125-29. 1960a Centipede and Damp caves: excavations in Val Verde County, Texas, 1958. Austin. 1960b Burins from Texas. Amer. Antiquity, 26: 93-97. ERASMUS, C. J.

1950

Patolli, pachisi, and the limitation of possibilities. SW. Jour. Anthr., 6: 369-87.

REFERENCES ESCUDERO, J.

1834

Noticias estadísticas del estado de Chihuahua. Mexico.

EXQUEMELIN, A.

1686

ESPINOSA, GASPAR DE

1519

1892

Relación e proceso quel Licenciado Gaspar de Espinosa, . . . ciudad de Panama a las provincias de París e Nata. . . . Col. Doc. Ined. Amer. y Oceania, 1864-84, 20: 5-119. Relación hecha por Gaspar de Espinosa, alcalde mayor de Castilla de Oro, dada á Pedranas de Avila, lugar-teniente general de aquellas provincias, de todo lo que sucedió en la entrada. In Doc. Ined. Colombia, vol. 2.

ESPINOSA, ISIDRO FELIX DE

1742

1746

El peregrino septentional Atlante delineado en la exemplarissima vida del . . . Antonio Margil de Jesus. Valencia. Crónica apostólica y seráfica de todos los colegios de propaganda fide de esta Nueva España. Mexico.

ESTRADA, E.

1956

Valdivia: un sitio arqueológico formativo en la costa de la provincia del Guayas, Eucador. Mus. Victor Emilio Estrada, Pub. 1. Guayaquil. 1957a Ultimas civilizaciones pre-históricas de la cuenca del Rio Guayas. Ibid., Pub. 2. 1957b Los Huancavilcas, ultimas civilizaciones pre-históricas de la costa del Guayas. Ibid., Pub. 3. 1957c Prehistoria de Manabi. Ibid., Pub. 4. 1958 Las cultural pre-clásicas, formativas o arcaicas del Ecuador. Ibid., Pub. 5. 1960 Newspaper article. Guayaquil. 1962 Arqueología de Manabi central. Mus. Victor Emilio Estrada, Pub. 7. AND C. EVANS

1963

Cultural development in Ecuador. In Meggers and Evans, 1963, pp. 77-88.

ESTUDIOS ANTROPOLÓGICOS

1956

Estudios antropológicos publicados en homenaje al doctor Manuel Gamio. Mexico.

EVANS, C ,

1957 1958 ,

1959

AND B. J. MEGGERS

Formative period cultures in the Guayas basin, coastal Ecuador. Amer. Antiquity, 22: 235-47. Valdivia, an early formative culture. Archaeology, 11:3. , AND E. ESTRADA

Cultura Valdivia. Mus. Victor Emilio Estrada, Pub. 6.

O.

Histoire des aventuriers qui se sont signalez dans les Indes. 2 vols. Paris. Bucaniers of America. London.

1893 EZELL, P.

H.

1954

An archaeological survey of northwestern Papagueria. Kiva, 19 ( 2 4 ) : 1-26. FAY, G. E. 1953 The archaeological cultures of the southern half of Sonora, Mexico. Year Book Amer. Phil. Soc, pp. 2 6 6 69. 1955 Prepottery, lithic complex from Sonora, Mexico. Science, 121: 777-78. 1956a Peralta complex, a Sonoran variant of the Cochise culture. Science, 124: 1029. 1956b A Seri fertility figurine from Bahia Kino, Sonora. Kiva, 21 (3, 4 ) : 1 1 12. 1957a A prepottery, lithic complex from Sonora, Mexico. Man, 57: 98-99. 1957b Peralta complex, a Sonoran variant of the Cochise culture. New World Antiquity, 4 ( 3 ) : 41-44. 1958 The Peralta complex, a Sonoran variant of the Cochise culture. Proc. 32d Int. Cong. Amer. (Copenhagen), pp. 491-93. 1959 Peralta complex, a Sonoran variant of the Cochise culture: new data, 1958. El Palacio, 66: 21-24. FERDON, E.

1953 1955 FERIZ,

1959

N.

Tonalá, Mexico, an archaeological survey. School Amer. Research, Monogr. 16. A trial survey of Mexican-Southwestern architectural parallels. Ibid., Monogr. 21. H.

Zwischen Peru und Mexiko. 2 vols. Kónongluk Inst. Tropen. Amsterdam.

FERNÁNDEZ,

L.

1881-1907 Colección de documentos para la historia de Costa Rica. 10 vols. San Jose, París, Barcelona. FERNÁNDEZ DE NAVARRETE,

1945

M.

Colección de los viajes y descubrimientos que hicieron los Españoles desde fines del siglo XV. 5 vols. Buenos Aires.

FERNÁNDEZ DE OVIEDO Y VALDES, G.

1851-55 Historia natural y general de las Indias. Madrid.

327

REFERENCES

1959a Historia general y natural de las Indias. Bib. Autores Españoles, vols. 67-71. Madrid. 1959b Natural history of the West Indies. Tr. and ed. by S. A. Stoudemire. Univ. North Carolina Studies Romance Lang, and Lit., no. 32. FERNÁNDEZ GUARDIA,

1908

FEWKES, J.

1893

1898

1907 1912 1922

FLINT,

n.d.

R.

Cartas de Juan Vázquez de Coronado. Barcelona. W.

A Central American ceremony which suggests the snake dance of the Tusayan villages. Amer. Anthr., o.s., 6: 284-306. Archaeological expedition to Arizona in 1895. Smithsonian Inst., Bur. Amer. Ethnol., 17th ann. rept., pp. 519-752. Certain antiquities of eastern Mexico. Ibid., 25th ann. rept., pp. 221-96. Casa Grande, Arizona. Ibid., 28th ann. rept., pp. 25-180. A prehistoric island culture area of America. Ibid., 34th ann. rept., pp. 35-281.

1960

Letters to Prof. F. W. Putnam, Peabody Museum, Harvard University. MS. The Janos, Jacomes, Mansos and Sumas Indians. New Mexico Hist. Rev., 32: 319-34. Apache, Navaho and Spaniard. Univ. Oklahoma Press.

FORD, J. A.

1951

Greenhouse: a Troyville—Coles Creek period site in Avoyelles Parish, Louisiana. Amer. Mus. Natural Hist., Anthr. Papers, vol. 44, no. 1.

AND G. I. QUIMBY

1945

1954

The Tchefuncte culture, an early occupation of the lower Mississippi Valley. Mem. Soc. Amer. Archaeol., no. 2. Poverty Point, a late archaic site in Louisiana. Amer. Mus. Natural Hist., Anthr. Papers, vol. 46, pt. 1.

1955

328

Jadeite from Manzanal, Guatemala. Amer. Antiquity, 2 1 : 81-83.

FUENTES Y GUZMAN, F.

Crooks site, a Marksville period burial mound in La Salle Parish, Louisiana. Dept. Conservation, Louisiana Geol. Survey, Anthr. Study, no. 3. New Orleans.

A.

1882-93 Historia de Guatemala o recordación florida escrita el siglo XVII. 2 vols. Madrid. GABLE, Ν.

1950 GAMIO,

1922

Ε.

The skeletal remains of Ventana Cave. In Haury, 1950, pp. 473-520. Μ.

La población del valle de Teotihuacan. Mexico.

GANGOLY, O.

1920

C.

A note on Kirtimukha: being the lifehistory of an Indian architectural ornament. Rupam, 1: 11-19. Calcutta.

GARCÍA CONDE, P.

1842 1849 1860

Ensayo estadístico sobre el estado de Chihuahua. Chihuahua. El album mexicana. Mexico. D.

Carta dirijida al rey de España, año 1576. In Squier, 1860.

GARCÍA PAYÓN, J.

1950

Restos de una cultura prehistórica encontrados en la región de Zempoala, Veracruz. Uni-Ver, 2 ( 1 5 ) : 90-130.

GARCÍA PELÁEZ, F. DE P.

1851-52 Memorias para la historia del antiguo reino de Guatemala. 3 vols. Guatemala. GARCÍA TORRES, V.

1856 GERALD,

1957

Documentos para la historia de Mexico. Mexico. R.

A historic house excavation near Janos, northwest Chihuahua, Mexico. M.A. thesis, Univ. Pennsylvania.

GERHARD, P.

n.d.

Aboriginal population of Baja California. MS.

GIFFORD, E.

1946

AND G . R . W l L L E Y

1940

Pub. Doc.

FOSHAG, W . F . , AND R . LESLIE

AND C. H. W E B B

1956

tr.

Benavides' Memorial of 1630. Acad. Amer. Franciscan Hist., Ser., vol. 2.

GARCÍA DE PALACIO,

E.

FORRES, J. D.

1957

FORRESTAL, P. R.,

W.

Archaeology in the Punta Peñasco region, Sonora. Amer. Antiquity, 11: 215-21.

AND R . H . L O W I E

1928

Notes on the Akwa'ala Indians of Lower California. Univ. California Pub. Amer. Ethnol. Archaeol., 3 1 : 257-334.

REFERENCES GILES, Η.

1923

A.

The travels of Fa-hsien (399-414 A.D.) or record of the Buddhistic kingdoms. Cambridge.

GILMORE, R.

1947

M.

Report on a collection of mammal bones from archaeologic cave-sites in Coahuila, Mexico. Jour. Mammalogy, 28: 147-65.

GLADWIN, H.

1937 1957

HAURY, E.

AND H. S.

N.

Ε.

Obras completas.

1956

1958

B.

C.

D.

The northern and southern tions of Antillean culture. Amer. Anthr. Assoc., no. 35.

GRAEBNER,

affiliaMem.

1960b

1960c 1961a

F.

1920-21 Alt- und neuweltliche Kalender. Zeit. für Ethnol, 5 2 / 5 3 : 6-37. GREENGO, R.

1960

1955

1960a

Monterrey.

1898a Researches in the Uloa valley. Mem. Peabody Mus., Harvard Univ., vol. 1, no. 4. 1898b Caverns of Copan. Ibid., vol. 1, no. 5. 1927

1953

Biological investigations in Mexico. Smithsonian Misc. Coll., vol. 115.

GORDON, G.

GOWER,

1952c

GLADWIN

A.

GONZÁLEZ, J.

1885

SAYLES, AND

The red-on-buff culture of the Papagueria. Ibid., no. 4.

GOLDMAN, E.

1951

B.

Excavations at Snaketown: material culture. Medallion Papers, no. 25.

GLADWIN, W.,

1929

1952b

S.

Excavations at Snaketown, II: Comparisons and theories. Medallion Papers, no. 26. A history of the ancient Southwest. Portland, Me.

, E. W. GLADWIN

1937

1952a

1961b

E.

Rocker-stamped pottery in the Old and New World. In Wallace, 1960, pp. 553-65.

1962

GRIFFIN, JAMES B.

1944

1945 1946

1949

Archaeological horizons in the southeast and their connections with the Mexican area. In El Norte de Mexico, pp. 283-86. The ceramic affiliations of the Ohio Valley Adena culture. In Webb and Snow, 1945, pp. 220-46. Culture change and continuity in eastern United States archaeology. Papers Peabody Found. Archaeol., 3: 37-96. Meso-America and the southeast: a

commentary. In John W. Griffin, 1949, pp. 77-99. [ed.] Archaeology of eastern United States. Univ. Chicago Press. Some early and middle Woodland pottery types in Illinois. In Deuel, 1952, pp. 93-129. An interpretation of the place of Spiro in southeastern archaeology. Missouri Archaeol., 14: 89-106. Prehistoric chronology estimates in the eastern United States and central Mexico, 1940-50. In Huastecos, Totonacos y sus Vecinos, pp. 485-96. Notes on the grooved axe in North America. Pennsylvania Archaeol., 25: 32-44. Prehistoric settlement patterns in the northern Mississippi Valley and the upper Great Lakes. In Willey, 1956, pp. 63-71. The chronological position of the Hopewellian culture in the eastern United States. Univ. Michigan, Mus. Anthr., Anthr. Papers, no. 12. Some prehistoric connections between Siberia and America. Science, 131: 801-12. Climatic change: a contributory cause of growth and decline of northern Hopewellian culture. Wisconsin Archaeol, 4 1 : 21-33. A hypothesis for the prehistory of Winnebago. In Diamond, 1960, pp. 809-65. Commentary on "Neolithic diffusion rates," by Munro S. Edmonson. Current Anthr., 2: 92-93. Some correlations of climatic and cultural change in eastern North American prehistory. Ann. New York Acad. Sci., 95: 710-17. A discussion of prehistoric similarities and connections between the Arctic and Temperate zones of North America. In J. M. Campbell, ed., Prehistoric cultural relations between the Arctic and Temperate zones of North America. Arctic Inst. North Amer., Tech. Papers, no. 11, pp. 154-63.

— AND A. ESPEJO

1947

La alfarería correspondiente al último período de occupación Nahua del Valle de Mexico, I: Tlatelolco a través de los tiempos. Mem. Acad. Mex. Hist., 6: 10-26. 329

REFERENCES

1950

La alfarería correspondiente al último período de occupación Nahua del Valle de Mexico, II. Ibid., 9: 3-54.

AND A. D. KRIEGER

1947

Notes on some ceramic techniques and intrusions in central Mexico. Amer. Antiquity, 12: 156-69.

, W. C. M C K E R N , AND P. F. TITTERINGTON

1945

Painted pottery figurines from Illinois. Ibid., 10: 295-302.

AND R. G. MORGAN,

1941

GRIFFIN, JOHN W.,

1949

eds.

Contributions to the archaeology of the Illinois River valley. Trans. Amer. Phil. Soc., 32: 1-208. ed.

The Florida Indian and his neighbors. Winter Park.

GROBMAN, Α., W. SALHUANA, AND R. SEVILLA, IN COLLABORATION WITH P. C. MANGELSDORF

1961

Races of maize in Peru: their origins, evolution and classification. Nat. Acad. Sci., Pub. 915.

GROUSSET,

1959

R.

Chinese art and culture.

GUIGNABAUDET,

1953

Nuevos descubrimientos arqueológicos en las tolas de Huaraqui. Bol. Informaciones Cien. Nac, 6: 168-86. Quito.

GUILLEMIN-TARAYRE,

1867

1869

London.

P.

The sculptures of Santa Lucia CosuSmithmalwhuapa in Guatemala. sonian Contrib. Knowledge, vol. 22.

HABERLAND,

1955

W.

Preliminary report on the Aguas Buenas culture, Costa Rica. Ethnos, 20: 224-30. 1957a Excavations in Costa Rica and Panama. Archaeology, 10: 258-63. 1957b Black-on-red painted ware and associated features in intermediate area. Ethnos, 22: 148-61. 1958 A pre-classic complex of western El Salvador, C. A. Proc. 32d Int. Cong. Amer. (Copenhagen), pp. 485-90. 1959a Archäologische Untersuchungen in Siidost - Costa Rica. Acta Humboldtiana, Ser. Geog. Ethnog., no. 1. Wiesbaden.

330

AND W. H. GREBE

1957

Prehistoric footprints from El Salvador. Amer. Antiquity, 22: 282-85.

HACKETT, C. W.,

ed.

1923-37 Historical documents relating to New Mexico, Nueva Vizcaya, and approaches thereto, to 1773. 3 vols. Carnegie Inst. Wash., Pub. 330. H A H N , P.

G.

1961

A relative chronology of the Cuban nonceramic tradition. Doctoral dissertation, Yale Univ.

E.

Notes archéologiques et ethnographiques: vestiges laissés par les migrations Américaines dans le nord du Mexique. Comm. Scien. Mexique, 3: 341-470. Paris. Exploration mineralogique des regions Mexicaines. París.

HABEL, S.

1878

1959b Chiriquian pottery types. Panama Archaeol, vol. 2, no. 1. 1959c A re-appraisal of Chiriquian pottery types. 33d Int. Cong. Amer. (San Jose), 2: 339-46. 1960a Cien años de arqueología en Panama. Pub. Rev. "Loteria," no. 12. 1960b Die Steinfiguren von Barriles in Panama. Die Umschau in Wissenschaft und Technik. Heft 23, 1 Dizember. Frankfurt am Main. 1960c Villalba, pt. I. Panama Archaeol, 3: 9 - 2 1 . in press a Observaciones en la península de Osa. Informes Inst. Geog. Costa Rica. in press b The scarified ware and the early cultures of Chiriqui, Panama. 34th Int. Cong. Amer. (Vienna).

HAMILTON,

1901

A.

Maori art.

HAMILTON, H.

1952

Wellington.

W.

The Spiro mound. ol., 14: 17-276.

Missouri Archae-

HAMMOND, G. P., AND A. REY

1928

Obregón's history of the 16th century explorations in western America. Los Angeles.

HANDBOOK OF SOUTH AMERICAN INDIANS

1946-59 J. H. Steward, ed. Smithsonian Inst., Bur. Amer. Ethnol, Bull. 143. 7 vols. HARCUM, C.

1923

HARDY, R. W.

1829

G.

Indian pottery from the Casas Grandes region, Chihuahua, Mexico. Bull Royal Ontario Mus. Archaeol. (December), pp. 4 - 1 1 . H.

Travels in the interior of Mexico. London.

HARRINGTON, M.

1933

HARTE, Ε.

1958

R.

Gypsum Cave, Nevada. Papers, no. 8.

SW.

Mus.

Μ.

Guacamaya Indian culture. ma, February.

Pana-

REFERENCES HAYDEN, J.

HARTE, Ν. Α.

1958 1960

A Madden Lake cave. Panama Archaeol., vol. 1, no. 1. Preliminary report on petroglyphs of the Republic of Panama.

HARTMAN, C.

1901 1907

Archaeological researches in Costa Rica. Stockholm. Archaeological researches on the Pacific coast of Costa Rica. Mem. Carnegie Mus., vol. 3, no. 1. Colonia Juarez, an intimate account of a Mormon village. Salt Lake City.

HAURY, E.

W.

1936a Some southwestern pottery types, series IV. Medallion Papers, no. 19. 1936b The Mogollon culture of southwestern New Mexico. Ibid., no. 20. 1944 Mexico and the southwestern United States. In El Norte de Mexico, pp. 203-05. 1945a The excavation of Los Muertos and neighboring ruins in the Salt River valley, southern Arizona. Papers Peabody Mus., Harvard Univ., vol. 24, no. 1. 1945b The problem of contacts between the southwestern United States and Mexico. SW. Jour. Anthr., 1:55-74. 1950a [ed.] The stratigraphy and archaeology of Ventana Cave, Arizona. Univ. New Mexico Press and Univ. Arizona Press. 1950b Summary of the archaeology of Papagueria. In his 1950a. 1960 Association of fossil fauna and artifacts of the Sulphur Spring stage, Cochise culture. Amer. Antiquity, 25: 609-10. 1962 The greater American Southwest. In Braidwood and Willey, 1962, pp. 106-31. AND OTHERS

1959

American Antiquity, vol. 24, no. 3. (Complete issue devoted to study of North American archaic cultures.)

, E. B. SAYLES, AND W. W.

1959

1936

WASLEY

The Lehner mammoth site, southeastern Arizona. Amer. Antiquity, 25: 2-30.

HAWLEY, F.

D.

Notes on the archaeology of the central coast of Sonora, Mexico. Kiva, 21 ( 3 , 4 ) : 19-23.

HEINE-GELDERN,

R.

1937

V.

HATCH, N. S.

1954

1956

M.

Field manual of prehistoric Southwestern pottery types. Univ. New Mexico Bull, Anthr. Ser., vol. 1, no. 4. (Rev. ed. 1950.)

L'art prébouddique de la Chine et de I'Asie du Sud-Est et son influence en Océanie. Rev. Arts Asiatiques, 11: 177-206. 1952 Some problems of migration in the Pacific. In Koppers, Heine-Geldern, and Haekel, eds., Kultur und Sprache. Wiener Beitrage zur Kulturgeschichte und Linguistik, 9: 3 1 3 62. 1954 Die asiatische Herkunft der südamerikanischen Metalltechnik. Paideuma, 5: 347-423. 1956a Conceptions of state and kingship in southeast Asia. Southeast Asia Program, Dept. Far Eastern Studies, Cornell Univ., Data Paper 18. 1956b Herkunft und Ausbreitung der Hochkulturen. Osterreichische Akad. Wissenschaften, Almanach (1955), Jahrgang 105, pp. 252-67. 1956—57 La escritura de la Isla de Pascua y sus relaciones con otras escrituras (observaciones al articulo del Dr. Thomas Barthel). Runa, 8: 5-27. 1958 Kulturpflanzengeographie und das Problem vorkolumbischer Kulturbeziehungen zwischen Alter und Neuer Welt. Anthropos, 53: 361-402. 1959a Chinese influences in Mexico and Central America: the Tajin style of Mexico and the marble vases from Honduras. Acts 33d Int. Cong. Amer. (San Jose), 1: 195-206. 1959b Chinese influence in the pottery of Mexico, Central America, and Colombia. Ibid., 1: 207-10. 1959c Representations of the Asiatic tiger in the art of the Chavin culture: a proof of early contacts between China and Peru. Ibid., 1: 321-26. 1964a Indonesian cultures. Encyclopedia World Art, vol. 8, columns 41-59. 1964b Traces of Indian and southeast Asiatic Hindu-Buddhist influences in Mesoamerica. Acts 35th Int. Cong. Amer. (Mexico), 1: 47-54. AND G. F.

1951

EKHOLM

Significant parallels in the symbolic arts of southern Asia and Middle America. In Tax, 1951, pp. 2 9 9 309. 331

REFERENCES HEISER, C. B., JR.

1965

Cultivated plants and cultural diffusion in nuclear America. Amer. Anthr., 67: 930-49.

HEIZER, R.

1942 1953

F.

Ancient grooved clubs and modern rabbit sticks. Amer. Antiquity, 8: 41-56. Additional notes on Chinese soapstone carvings from Mesoamerica. Ibid., 19: 81.

HOMENAJE CASO

1951

Homenaje al doctor Alfonso Caso. Nuevo Mundo. Mexico.

HOOTON, E.

1930 HOUGH,

W.

1923

Casas Grandes pottery in the National Museum. Art and Archaeol., 16: 34.

HOWARD, A.

1957

Observations on archaeological sites in Topanga Canyon, California. Univ. California Pub. Amer. Archaeol. Ethnol, 44: 237-58.

AND W . C . MASSEY

1953

Aboriginal navigation off the coasts of upper and Baja California. Smithsonian Inst., Bur. Amer. Ethnol., Bull. 151, pp. 285-311.

HENNING, P.

1918

El Xipe del Tazumal de Chalchuapa, Departamento de Santa Ana, República de El Salvador. Disertaciones Cien. Autores Alemanes en Μex., no. 4.

HERRERA, A. DE

1726-30 Historia general de los hechos de los castellanos en las islas i tierra firme del mar océano. 5 vols. Madrid. H E W E T T , E.

1908 1923 1930

HINTON, Τ.

1955 Ho,

1904 1953

1888 1903

1935

The beginnings of Spanish settlement in the El Paso district. Univ. California Pub. Hist., vol. 1, no. 3.

HUGHES, J. T.

1947

An archaeological reconnaissance in Tamaulipas, Mexico. Amer. Antiquity, 13: 33-39.

HUMROLDT, A . DE

1816 HUNT,

1960

Vues des cordillères et monuments des peuples indigènes de I'Amérique. Paris.

H.

Ancient art of the province of Chiriqui, Colombia. Smithsonian Inst., Bur. Amer. Ethnol., 6th ann. rept. Aboriginal pottery of the eastern United States. Ibid., 20th ann. rept. Artifacts from Estero de Tastiota, Sonora. Kiva, 21 (3, 4 ) : 12-19.

A.

Archeology of the Death Valley salt pan, California. Univ. Utah, Anthr. Papers, no. 47.

HUTCHINSON, STEPHENS

1947

J.

B.,

R.

A.

SILOW, AND S.

G.

The evolution of Gossypium and the differentiation of the cultivated cottons. London.

INFORME DE LOS PADRES

1697

The introduction of American food plants into China. Amer. Anthr., 57: 191-201.

HOLZKAMPER, F . M .

1956

Huastecos, Totonacos y sus vecinos. Ed. by Bernal and Dávalos Hurtado. Rev. Μex. Estud. Antr., vol. 13, nos. 2 and 3.

HUGHES, A. E.

Β.

A survey of archaeological sites in the Altar valley, Sonora. Kiva, 21 (1, 2 ) : 1-12.

HOLMES, W.

Notes on the Indians of Sonora, Mexico. Amer. Anthr., 6: 51-89.

HUASTECOS, TOTONACOS Y sus VECINOS

PING-TI

1955

332

HRDLIČKA, A.

L.

Les communautés anciennes dans le désert Américain. Librairie Kündig. Geneva. Anahuac and Aztlan: retracing the legendary footsteps of the Aztecs. Art and Archaeol, 16: 35-50. Ancient life in the American Southwest. Indianapolis.

M.

Navacoyan: a preliminary survey. Bull. Texas Archaeol. Soc, 28: 1 8 1 89.

AND E . M . L E M E R T

1947

A.

The Indians of Pecos. Phillips Acad., Dept. Archaeol. New Haven.

IVES, R.

Informe de los padres misioneros Fray Francisco de San Joseph y Fray Pablo de Rebullida. Talamanca. Archivos Nacionales, doc. 5226. San Jose, Costa Rica. L.

1963a The problem of the Sonoral littoral cultures. Kiva, 28 ( 3 ) : 28-32. 1963b The bell of San Marcelo. Kiva, 29 ( 1 ) : 14-22. IZUMI, S., AND T. SONO

1963

Andes 2: excavations at Kotosh, Peru. Univ. Tokyo Exped. to the Andes, 1960. Tokyo.

REFERENCES

JANSE, O. R. T. 1947 Archaeological research china. Vol. 1.

JONES, V. H.,

in

Indo-

1954

JAYASWAL, K. P.

1933

Metal images of Kurkihar monastery. Jour. Indian Soc Oriental Art, vol. 1. Calcutta.

JELKS, E. B., E. M. DAVIS, AND Η. Β. STURGIS

1960

A review of Texas archaeology. Bull. Texas Archaeol. Soc, vol. 29 (for 1958).

JENNINGS, J.

1956 1957

D.

[ed.] The American Southwest: a problem in cultural isolation. In Wauchope, 1956, pp. 59-127. Danger Cave. Mem. Soc. Amer. Archaeol, no. 14.

AND G. NEUMANN

1940

A variation of Southwestern Pueblo culture. Lab. Anthr., Tech. Ser., Bull. 10. Santa Fe.

JOYCE, T.

1916 JUDD, N.

1954

1930 1938 1951

Contribución al conocimiento de los aborígenes de la provincia de Imbabura. Estud. Prehist. Amer., II. Madrid. Una gran marea cultural en el N. O. de Sud America. Jour. Soc. Amer. Paris, 22: 107-97. Sebastian de Benalcazar. Vol. 2. Editorial Ecuatoriana. Quito. Antropología prehispánica del Ecuador, 1945. Quito.

JOHNSON, A. E.

1960

1963

The place of the Trincheras culture of northern Sonora in Southwestern archaeology. M.A. thesis, Univ. Arizona. The Trincheras culture of northern Sonora. Amer. Antiquity, 29: 17486.

JOHNSON, A.

1958

1955

1955

1951

The linguistic map of Mexico and Central America. In The Maya and their Neighbors, pp. 88-114. [ed.] Radiocarbon dating. Mem. Soc. Amer. Archaeol., no. 8.

JOHNSON, J. B.

1950

The Opata: an inland tribe of Sonora. Univ. New Mexico Pub. Anthr., no. 6.

1947 1952 1953

1955 1956 1959

1960

Preliminary pollen analysis of Damp and Centipede rockshelters. In Epstein, 1960a, pp. 167-73.

Late Ming and early Ch'ing porcelain fragments from archaeological sites in Florida. Florida Anthr., 8: 91-110. The atlatl in North America. Indiana Hist. Soc, Prehist. Research Ser., vol. 3, no. 3. C.

The cultural affiliations and chronological position of the Clear Fork focus. Amer. Antiquity, 13: 97-109. Some geographic and cultural factors involved in Mexican-southeastern contacts. In Tax, 1952, pp. 139-44. Reconnaissance and excavation in Durango and southern Chihuahua, Mexico. Year Book Amer. Phil. Soc, pp. 172-76. Juan Sabeata and diffusion in aboriginal Texas. Amer. Anthr., 57: 981-95. Settlement patterns in north central Mexico. In Willey, 1956a, pp. 12839. The desert cultures and the Balcones phase: archaic manifestations in the Southwest and Texas. Amer. Antiquity, 24: 276-88. North Mexico and the correlation of Mesoamerican and Southwestern cultural sequences. In Wallace, 1960, pp. 566-73.

AND W. J. SHACKELFORD

1954

Preliminary notes on the Weicker site, Durango, Mexico. El Palacio, vol. 61, no. 5.

AND H. D. WINTERS

1960

JOHNSON, L.

1960

M.

The material culture of Pueblo Bonito. Smithsonian Misc. Coll., vol. 123.

KELLEY, J.

JOHNSON, F.

1940

A.

Central American and West Indian archaeology. New York.

KELLER, J. H.

S.

Similarities in Hohokam and Chalchihuites artifacts. Amer. Antiquity, 24: 126-30.

FONNER

KAMER, AGA-OGLU

JIJÓN Y CAAMAÑO, J.

1914

AND R. L.

Plant materials from the Durango area. In E. H. Morris and R. F. Burgh, Basket Maker II sites near Durango, Colorado, pp. 93-115. Carnegie Inst. Wash., Pub. 604.

A revision of the archaeological sequence in Sinaloa, Mexico. Amer. Antiquity, 25: 547-61.

KELLY, I. T.

1938

Excavations at Chametla, Ibero-Amer., no. 14.

Sinaloa.

333

REFERENCES

1939

1944 1945

An archaeological reconnaissance of the west coast: Nayarit to Michoacan. Proc. 27th Int. Cong. Amer. (Mexico), 2: 74-77. West Mexico and the Hohokam. In El Norte de Mexico, pp. 206-22. Excavations at Culiacan, Sinaloa. Ibero-Amer., no. 25.

KERR, R.

1814

A general history and collection of voyages and travels. Edinburgh.

KIDDER, A.

1916

1924

1939 1947 1949

1956

KRIEGER, A.

1945

1953

V.

The pottery of the Casas Grandes district, Chihuahua. In Holmes Anniversary Volume, Anthr. Essays, pp. 253-68. An introduction to the study of Southwestern archaeology. Phillips Acad., Dept. Archaeol, Papers SW. Expedition, no. 1. Notes on the archaeology of the Babicora district, Chihuahua. In Brand and Harvey, 1939, pp. 221-32. The artifacts of Uaxactun, Guatemala. Carnegie Inst. Wash., Pub. 576. Certain archaeological specimens from Guatemala. Carnegie Inst. Wash., Notes Middle Amer. Archaeol. Ethnol., no. 92.

1956

The Pendleton ruin, Hidalgo County, New Mexico. Carnegie Inst. Wash., Pub. 585, Contrib. 50.

, J. D. JENNINGS, AND Ε. Μ.

1946

KINNAIRD, L.,

1958

ed.

The frontiers of New Spain: Nicolas de LaFora's description 1766-1768. Quivira Soc. Pub., 13: 98-99.

KIRCHHOFF, P.

1942 1943 1964

Las tribas de la Baja California y el libro de Padre Baegert. In Baegert, 1942. Mesoamérica: sus límites geográficas, composición étnica y caracteres culturales. Acta Amer., 1: 92-107. The diffusion of a great religious system from India to Mexico. Acts 35th Int. Cong. Amer. (Mexico), 1: 73-100.

KOPPERS, W .

1957

Das Problem der Universalgeschichte im Lichte von Ethnologie und Prähistorie. Anthropos, 52: 369-89.

KRICKEBERG,

1950

334

W.

Ostasien-Amerika: Bemerkungen eines Amerikanisten zu zwei Büchern

Archaeological salvage in the Falcon reservoir area: progress report no. 1. Mimeographed. Austin.

KRMPOTIC, M.

1923

D.

The life and works of the Reverend Ferdinand Konščak, S. J. Boston.

KROEBER, A. L.

1925 1932 1939

SHOOK

Excavations at Kaminaljuyu, Guatemala. Ibid., Pub. 561.

D.

An inquiry into supposed Mexican influence on a prehistoric "cult" in the southern United States. Amer. Anthr., 47: 483-515. Recent developments in the problem of relationships between Mexican gulf coast and eastern United States. In Huastecos, Totonacos y sus Vecinos, pp. 497-518. Some Mexican figurine heads in Texas. Bull. Texas Archaeol. Soc, 27: 258-65.

AND J. T. HUGHES

1950

, AND H . S. AND C . B . COSGROVE

1949

Carl Hentzes. Sinologica, 2: 195233. Basel. Altmexikanische Kulturen. Berlin.

Handbook of the Indians of California. Smithsonian Inst., Bur. Amer. Ethnol, Bull. 78. The Patwin and their neighbors. Univ. California Pub. Archaeol. Ethnol, 29: 253-423. The cultural and natural areas of native North America. Ibid., vol. 38.

KURTZ, E. B., H. TUCKER, AND J. L. LIVERMAN

1960

Reliability of identification of fossil pollen as corn. Amer. Antiquity, 25: 605-06.

LADD, J.

1957

A stratigraphic trench at Sitio Conte, Panama. Amer. Antiquity, 22: 2 6 5 71.

LANNING, E.

P.

1960a Cerámica antigua de la costa peruana: nuevos descubrimientos. (2d ed., rev.) Inst. Andean Studies. Berkeley. 1960b Chronological and cultural relationships of early pottery styles in ancient Peru. Doctoral dissertation, Univ. California. 1961 Cerámica pintada pre-Chavin de la costa central del Peru. Rev. Mus. Nac, 30: 78-83. Lima. 1963 Olmec and Chavin: reply to Michael D. Coe. Amer. Antiquity, 29: 9 9 101.

REFERENCES

aeology. In Jelks, Davis, and Sturgis, 1960.

LAS CASAS, Β. DE

n.d. 1951

Historia de Ias índias. Apéndice apologética historia. (Aguilar ed.) Madrid. Historia de Ias índias. (Millares Carlo ed.) Mexico.

LATHRAP, D.

1958 1960 1962 1963 1965

1953 1959

1910 1913 1920 1938

Reisebericht aus San Jose de Costa Rica. Zeit. fiir Ethnol., vol. 40. Berlin. Ergebnisse einer Forschungsreise in Mittelamerika und Mexiko, 1907-09. Ibid., vol. 42, no. 5. Die Archäologie Costa Rica. Abhandlungen der Naturhistorischen Gesellschaft Νumberg, vol. 20. Zentral-Amerika. Pt. 1. Berlin. Die Geschichte der Kõnigreiche von Colhuacan und Mexiko. Quellenwerke zur alten Geschichte Amerikas, vol. 1. Berlin.

LEHMER, D. J.

1948 1949 1960

1956

The Jornada branch of the Mogollon. Bull. Univ. Arizona, vol. 19, no. 2. Archaeological survey of Sonora, Mexico. Chicago Natural Hist. Mus., Bull. 20, no. 12, pp. 4 - 5 . A review of trans-Pecos Texas arch-

La filosofía nahuatl.

LIBBY, W.

1951 1952 1955

Mexico.

F.

Radiocarbon dates, II. Science, 114: 291-96. Radiocarbon dates, III. Ibid., 116: 673-81. Radiocarbon dating. (2d ed.) Chicago.

LINCK, W.

n.d.

Carte. In Doc. para la Historia de Mexico, 1853-57.

LING, SHUN-SHENG

1956

H.

Le personnage couché sur le dos: sujet commun dans I'archéologie du Mexique et de l'Equateur. In Tax, 1951, pp. 291-98. On Noel Morss' "Cradled infant figurines." Amer. Antiquity, 19: 78-80. Les céramiques précolombiennes. Paris.

LEHMANN, W.

1908

Relación y discursos del descubrimiento, población y pacificación de este nuevo reino de Leon [1649]. In Doc. inéditos y muy raros para la historia de Mexico, by Genaro García, vol. 25. Mexico.

LEÓN-PORTILLA, Μ.

Chinese pottery of the Han dynasty. Leyden.

LEHMANN,

1951

1909

W.

The cultural sequence at Yarinacocha, eastern Peru. Amer. Antiquity, 23: 379-88. Review of Cultura Valdivia, by Evans, Meggers, and Estrada. Ibid., 26: 125-27. Yarinacocha, stratigraphic excavations in the Peruvian montaña. Doctoral dissertation, Harvard Univ. Los Andes centrales y la montaña. Rev. Mus. Nac., 32: 197-202. Lima. Origins of central Andean civilization: new evidence. Review of Izumi and Sono, 1963. Science, 148: 796-98.

LAUFER, B.

1909

LEÓN, A. DE

Patu found in Taiwan and other east Asiatic regions and its parallels in Oceania and America. Nat. Taiwan Univ., Bull. Dept. Archaeol. and Anthr., 7: 1-22, 82-104. Taipei.

LlNNE, S.

1929

1934 1942

1951 1956

Darién in the past. Götesborg Kungl. Veterskaps - och Vitterhets Samhülles Handlinar, Femte Följden, Ser. A, 1, 3. Archaeological researches at Teotihuacan, Mexico. Ethnog. Mus. Sweden, n.s., Pub. 1. Stockholm. Mexican highland cultures: archaeological researches at Teotihuacan, Calpulalpan, and Chalchicomula in 1934-35. Ibid., Pub. 7. A wheeled toy from Guerrero, Mexico. Ethnos, 16: 141-52. Radiocarbon dates at Teotihuacan. Ibid., 2 1 : 180-93.

LISTER, R.

H.

1939

A report on the excavations made at Agua Zarca and La Morita in Chihuahua. Univ. New Mexico, Research Graduate School, vol. 3, no. 1. 1946 Survey of archaeological remains in northwestern Chihuahua. SW. Jour. Anthr., 2: 433-54. 1953a The stemmed, indented base point, a possible horizon marker. Amer. Antiquity, 18: 265. 1953b Excavations in Cave valley, Chihuahua, Mexico: a preliminary note. Ibid., 19: 166-69. 335

REFERENCES

1958

Archaeological excavations in the northern Sierra Madre Occidental, Chihuahua and Sonora, Mexico. Univ. Colorado Studies, Anthr. Ser., no. 7.

AND A. M.

1955

1702

Ν.

Arte de la lengua Tequima, vulgarmente llamada Opata. Mexico.

LONGYEAR, J.

1944 1947

1948 1952

M.

Archaeological investigations in El Salvador. Mem. Peabody Mus., Harvard Univ., vol. 9, no. 2. Cultures and peoples of the southeastern Maya frontier. Carnegie Inst. Wash., Theoretical Approaches to Problems, no. 3. A sub-pottery deposit at Copan, Honduras. Amer. Antiquity, 13: 2 4 8 49. Copan ceramics: a study of southeastern Maya pottery. Carnegie Inst. Wash., Pub. 597.

LÓPEZ DE GOMARA,

1941

F.

Historia general de Las Indias. vols. Madrid.

2

LORENZO, J. L.

1959

La colección Leon Diguet de la Baja California en el Museo del Hombre, Paris: el material lítico. Paper presented to annual meeting, Amer. Anthr. Assoc, Mexico City.

LOTHROP, R . W . , AND S. K . LOTHROP

1927

The use of plaster on Porto Rican stone carvings. Amer. Anthr., 29: 728-30.

LOTHROP, S. K.

1919

The discovery of gold in the graves of Chiriqui. Mus. Amer. Indian, Heye Found., Indian Notes and Monogr., vol. 6, no. 2. 1926 Pottery of Costa Rica and Nicaragua. Ibid., Contrib. 8. 1927a Pottery types and their sequence in El Salvador. Ibid., Indian Notes and Monogr., vol. 1, no. 4. 1927b The potters of Guatajiagua, El Salvador. Ibid., Indian Notes, 4: 109-18. 1937-42 Cocle: an archaeological study of central Panama. Mem. Peabody Mus., Harvard Univ., vols. 7, 8. 1939 The southeastern frontier of the Maya. Amer. Anthr., 4 1 : 42-54.

336

Gold ornaments of Chavin style from Chongoyape, Peru. Amer. Antiquity, 6: 250-62. The Sigua: southernmost Aztec outpost. Proc. 8th Amer. Sci. Cong., Anthr. Sci., Native Amer. Cultures, 2: 109-16. Archaeology of southern Veraguas, Panama. Mem. Peabody Mus., Harvard Univ., vol. 9, no. 3. Gold artifacts of Chavin style. Amer. Antiquity, 16: 226-40. Jade and string sawing in northeastern Costa Rica. Ibid., 2 1 : 43-51. Archaeology of the Diquis delta, Costa Rica. Papers Peabody Mus., Harvard Univ., vol. 51.

1942

HOWARD

The Chalchihuites culture of northwestern Mexico. Amer. Antiquity, 21: 122-29.

LOMBARDO,

1941

1950 1951 1955 1963

AND OTHERS

1961 Lou,

D.

1957

Essays in pre-Columbian art and archaeology. Harvard Univ. Press. WING-SOU

Rain-worship among the ancient Chinese and the Nahua-Maya Indians. Acad. Sínica, Bull. Inst. Ethnol., 4: 31-102. Taipei.

LOVEN, S.

1924 1935

Über die Wurzeln der Tainischen Kultur. Gõteborg. Origins of the Tainan culture, West Indies. Gõteborg.

LOWE, G. W., AND C. NAVARRETE.

1959

Research in Chiapas, Mexico. New World Archaeol. Found., Pub. 3.

LUMHOLTZ,

1900 1902 1912

C.

Symbolism of the Huichol Indians. Mem. Amer. Mus. Natural Hist., vol. 3, pt. 2. Unknown Mexico. 2 vols. New York. New trails in Mexico. New York.

MACCURDY, G.

1911

MCGIMSEY, C.

1956 1957

C.

Burial of an early American magician. Proc. Amer. Phil. Soc, 86: 270-98.

MCKUSICK, Μ.

1960

R.

Cerro Mangote: a preceramic site in Panama. Amer. Antiquity, 22: 1 5 1 61. Further data and a date from Cerro Mangote, Panama. Ibid., 23: 4 3 4 35.

MCGREGOR, J.

1943

G.

A study of Chiriquian antiquities. Mem. Connecticut Acad. Arts and Sci., vol. 3.

Β.

Aboriginal canoes in the West Indies. Yale Univ. Pub. Anthr., no. 63.

REFERENCES AND A . T . G l L M A N

1959

MAPS

An acorn grinding site in Baja California. Univ. California Archaeol. Survey, ann. r e p t , pp. 47-58.

MACNEISH, R. S.

1954 1958 1959

1960 1961

1962

An early archaeological site near Panuco, Vera Cruz. Trans. Amer. Phil Soc., 44: 539-641. Preliminary archaeological investigations in the Sierra de Tamaulipas, Mexico. Ibid., 48: 1-209. A speculative framework of northern North American prehistory as of April, 1959. Anthropologica, n.s., 1 ( 1 , 2 ) : 7-23. Ottawa. Rejoinder to Taylor. Amer. Antiquity, 25: 591-93. First annual report of the Tehuacan archaeological-botanical project. Robert S. Peabody Found. Archaeol., Phillips Acad. Andover. Second annual report of the Tehuacan archaeological-botanical project. Ibid.

MACNUTT, F.

1908

1956

1958

1958

C.

1951

1928

1939 1944

1964

1952

Archaeological evidence of the evolution of maize in northwestern Mexico. Bot. Mus. Leafl., Harvard Univ., 17: 151-78.

1954 1961

Archaeological evidence on the diffusion and evolution of maize in northeastern Mexico. Bot. Mus. Leafl., Harvard Univ., 17: 125-50. Domestication of corn. Science, 143 (3606): 538-45.

1962

Whence came maize to Asia? 14: 263-91.

Ibid.,

AND R. G. REEVES

1959

The origin of corn. 440.

1954

P.

La comerca lagunera a fines del siglo XVI y principios del XVII según las fuentes escritas. Inst. Hist., 1st ser., no. 30. Mexico.

MÁRTIR DE ANGLERÍA

See Martyr, P. MARTYR, P.

Ibid., 18: 3 2 9 -

1912

MANJE, J. M.

1954

Mogollon cultural continuity and change: the stratigraphic analysis of Tularosa and Cordova caves. Fieldiana: Anthr., vol. 40. Caves of the Reserve area. Ibid., vol. 42. Mineral Creek site and Hooper Ranch pueblo, eastern Arizona. Ibid., vol. 52. Chapters in the prehistory of eastern Arizona, I. Ibid., vol. 53.

MARTÍNEZ DEL Río,

AND D. L. OLIVER

1951

Estudio arquitectónico comparativo de los monumentos arqueológicos de Mexico. Contrib. of Mexico to 23d Int. Cong. Amer., pp. 1-86. Atlas arqueológico de la República Mexicana. Inst. Panamer. Geog. Hist., Pub. 41. Los monumentos de Mexico y los del suroeste y sureste de Estados Unidos. In El Norte de Mexico, pp. 252-54.

MARTIN, P. S., J. B. RINALDO, AND OTHERS

New evidence on the origin and ancestry of maize. Amer. Antiquity, 19: 409-10.

R. S. MACNEISH, AND W. C. GALINAT

1956

Η.

Le décor et la sculpture Khmers. Etudes d'Art et d'Ethnol. Asiatiques, vol. 3. Paris.

MARQUINA, I.

AND R. H. LISTER

1956

Nuevo mapa oficial de estado de Sonora. Sonora News Co., Nogales, by C . E . Herbert. U.S.A.F. preliminary base. Sec. 4718 (Douglas) Aeronautical Chart Service. Washington. World aeronautical chart. Secs. 405 (Gila River), 406 (Estacado Plains), 471 (Sonora), 470 (Santiago Mts.). U.S. Coast and Geodetic Survey. Estados Unidos Mexicanos. Com. Intersecretarial Coordinadora del Levantamiento de la Carta Geográfica de la República Mexicana. Hermosillo, 12R-IV; Chihuahua, 13R-III; Nogales, 12R-II; Ciudad Juarez, 13R-I. Mexico.

MARCHAL,

R.

Edward Palmer: plant explorer of the American west. Univ. Oklahoma Press.

MANGELSDORF, P.

1954

1948

A.

Letters of Cortés: the five letters of relation from Hernando Cortés to the Emperor Charles V. 2 vols. New York.

MCVAUGH,

1904

Luz de tierra incognita. Pt. 2. by Η. J. Karns. Tucson.

Tr.

1944

De orbe novo: the eight decades of Peter Martyr d'Anghiera. Tr. by F. A. MacNutt. 2 vols. New York. Décadas del nuevo mundo. Buenos Aires.

337

REFERENCES

1796. New Mexico Hist. Rev., 32: 335-56.

MASON, J. A.

1936 1937

1942 1945

1951

Archaeological work in Durango during March, 1936. Bull. Univ. Pennsylvania Mus., 6: 136-37. Late archaeological sites in Durango, Mexico, from Chalchihuites to Zape. In Twenty-fifth Anniversary Studies, Pub. Philadelphia Anthr. Soc, 1: 127-46. New excavations at the Sitio Conte, Cocle, Panama. Proc. 8th Amer. Sci. Cong., vol. 2. Washington. Costa Rican stonework: the Minor C. Keith Collection. Amer. Mus. Natural Hist., Anthr. Papers, vol. 39, no. 3. On two Chinese figurines found in Mesoamerica. In Homenaje Caso, pp. 271-76.

MASSEY, W.

C.

Archaeology in central and southern Baja California: the Castaldi Collection. MS. n.d.,b The historical ethnography of Baja California. MS. 1947 Brief report on archaeological investigations in Baja California. SW. Jour. Anthr., 3: 344-59. 1949 Tribes and languages of Baja California. Ibid., 5: 272-307. 1955 Culture history in the cape region of Baja California. Doctoral dissertation, Univ. California. 1961a The survival of the dart-throwers on the peninsula of Baja California. SW. Jour. Anthr., 17: 81-93. 1961b The cultural distinction of aboriginal Baja California. Soc. Mex. Antr. Mexico.

MAUDSLAY, A.

MAYA AND THEIR NEIGHRORS, T H E

1940

AND C. M. OSBORNE

A burial cave in Lower California: the Palmer Collection, 1887. Univ. California Anthr. Records, vol. 16, no. 8.

AND D . R . TUOHY

n.d.,a n.d.,b n.d.,c

Caves of the Sierra de la Giganta. Petroglyphs of Baja California. Pottery from the northwest coast of Baja California Norte, Mexico.

MATIENZO, J. DE

1910

Obra escrita en el siglo XVI por el Lic. don Juan Matienzo, oidor de la Real Audiencia de Charcos. Facultad Filosofía y Letras, Sec. Hist. Lima.

MATSON, D. S., AND A. H.

1957 338

SCHROEDER

Cordero's description of the Apache,

Ed. by C. L. Hay and others. York.

MAYER-OAKES, W.

1959 1963

New

J.

A stratigraphic excavation at El Risco, Mexico. Proc. Amer. Phil. Soc, vol. 103, no. 3. Early man in the Andes. Sci. Amer., 208: 116-28.

AND R. E.

1960

BELL

Early man site found in highland Ecuador. Science, 131: 1805-06.

MEGGERS, B. J.

1963

n.d.,a

1961

P.

1889-1902 Archaeology. In Biologia Centrali-Americana. 5 vols. London.

Cultural development in Latin America: an interpretative overview. In Meggers and Evans, 1963, pp. 1 3 1 45.

AND C. EVANS

1955 1957 1962 1963

1964

[eds.] New interpretations of aboriginal American culture history. Anthr. Soc. Washington, 75th anniv. vol. Archeological investigations at the mouth of the Amazon. Smithsonian The Machallila culture: an early formative complex on the Ecuadorian coast. Amer. Antiquity, 28: 186-92. [eds.] Aboriginal cultural development in Latin America: an interpretative review. Smithsonian Misc. Coll., vol. 146, no. 1. Especulaciones sobre rutas tempranas de difusión de la cerámica entre Sur y Mesoamérica. Hombre y Cultura, 1 ( 3 ) : 1-5. Panama.

MEIGHAN, C.

1950

1959 1960

W.

Excavations in sixteenth century shell-mounds at Drake's Bay, Marin County. Univ. California Archaeol. Survey, Rept. 9, pp. 27-32. New findings in west Mexican archaeology. Kiva, 25: 1-7. Prehistoric copper objects from westem Mexico. Science, 131: 1534.

AND R. F. HEIZER

1952

Archaeological exploration of sixteenth-century Indian mounds at Drake's Bay. California Hist. Soc. Quar., 3 1 : 98-108.

MEIGS, P.

1935

The Dominican frontier of Lower California. Univ. California Pub. Geog., vol. 7.

REFERENCES

1939

The Kiliwa Indians of Lower California. Ibero-Amer., no. 15.

MOTA Y ESCOBAR, A. DE LA

1940

MENZEL, D., J. H. ROWE, AND L. DAWSON

1964

The Paracas pottery of lea, a study in style and time. Univ. California Pub. Amer. Archaeol. Ethnol., vol. 50.

MÜLLERRIED, F. K. G.

1934

MERWIN, R. E., AND G. C. VAILLANT

1932

The ruins of Holmul, Guatemala. Mem. Peabody Mus., Harvard Univ., vol. 3, no. 2.

MITCHELL, R.

MOLINA SOLIS, J. F.

1896

Historia del descubrimiento y conquista de Yucatan con una reseña de la historia antigua de esta península. Merida.

MONTERDE, J. M.

1842-45 Información sobre las ruinas de Casas Grandes, Chihuahua. MOORE, C. B.

1905

Certain aboriginal remains of the Black Warrior River. Jour. Acad. Natural Sci. Philadelphia, 13: 3 3 7 405.

1950

1600

Los Mayas de Honduras y los indígenas antillanos precolombinos. Tzunpame, año 7, no. 6-7, pp. 9-40. Los Mayas de Honduras y los indígenas antillanos precolombinos. Rev. Arqueol. Ethnol., 2d ep., no. 10-11, pp. 69-100. Havana. Admiral of the ocean sea: a life of Christopher Columbus. 2 vols. Boston.

MORLEY, S.

1946

1961

MORRIS, Ε.

1919

1927

À propos d'un eventuel emprunt de l'art maya aux arts de I'Inde extérieure. Proc. 34th Int. Cong. Amer. (Vienna), pp. 340-47.

NAVARRETE, C.

1960

Archeological explorations in the region of the Frailesca, Chiapas, Mexico. New World Archaeol. Found., Pub. 6.

NAVARRETE, M. F. DE

1825-37 Colección de los viajes y descubrimientos, que hicieron por mar los españoles desde fines del siglo XV. 5 vols. Madrid. NELSON, Ε.

1921

1931

1949

Η.

Η.

The ancient Mímbreños based on investigations at the Mattocks ruin, Mimbres Valley, New Mexico. Logan Mus., Bull. 4. The George C. Davis site, Cherokee County, Texas. Mem. Soc. Amer. Archaeol., no. 5.

NEWMAN, M.

1958

Stanford Univ.

Preliminary account of the antiquities of the region between the Mancos and La Plata rivers in southwestern Colorado. Smithsonian Inst., Bur. Amer. Ethnol., 33d ann. r e p t , pp. 155-206. The beginnings of pottery making in the San Juan area: unfired prototypes and the wares of the earliest ceramic period. Amer. Mus. Natural Hist., Anthr. Papers, vol. 28, pt. 2.

W.

Lower California and its natural resources. Mem. Nat. Acad. Sci., 16: 1-194.

N E W E L L , H. P., AND A. D. KRIEGER

G.

The ancient Maya. Press.

May 15, Doc. 38

NAUDOU, J.

MORISON, S. E.

1942

Guatemala II. (2).

NESBITT, P.

MORALES PATINO, O.

1949

Sobre artefactos de piedra en la porción oriental del estado de Coahuila. An. Mus. Nac. Arqueol. Hist. Ethnol., pp. 205-19.

NATIONAL ARCHIVES OF SEVILLE

H.

1959a Projectile points from Panama. Panama Archaeol., vol. 2, no. 1. 1959b An unreported pottery vessel from Panama. Ibid., vol. 2, no. 1. in press Wooden artifacts from cave urn-burials in Madden Lake. Ethnos.

Descripción geográfica de los reinos de Nueva Galicia, Nueva Viscaya y Nuevo Leon. Mexico.

NICHOLSON, H.

1953

B.

On a supposed Mesoamerican "Thin Orange" vessel from Ecuador. Amer. Antiquity, 19: 164-66.

NICKERSON, N.

1953

T.

A trial formulation presenting evidence from physical anthropology for migrations from Mexico to South America. In R. H. Thompson, 1958a, pp. 33-40.

H.

Variation in cob morphology among certain archaeological and ethnological races of maize. Ann. Missouri Bot. Garden, 40: 79-111.

NOGUERA, Ε.

1926

Ruinas arqueológicas de Casas Grandes. Mexico. 339

REFERENCES

1930

Ruinas arqueológicas del norte de Mexico, Casas Grandes (Chihuahua), La Quemada, Chalchihuites (Zacatecas). Pub. Sec. Educ. Pública, Talleres Gráficos de la Nación, pp. 5 27. Casas Grandes. 1935a Antecedentes y relaciones de la cultura teotihuacana. El Mex. Antiguo, 3 ( 5 - 8 ) : 3-81. 1935b La cerámica de Tenayuca y las excavaciones estratigráficas. Mus. Nac. Arqueol., Hist. Etnog. Mexico. 1939 Exploraciones en "El Opeño," Michoacan. Acts 27th Int. Cong. Amer. (Mexico), vol. 1. 1946 Cultura de El Opeño. Mex. Prehispánico, pp. 150-54. 1954 La cerámica arqueológica de Cholula. Mexico. 1958 Reconocimiento arqueológico en Sonora. Inst. Nac. Antr. Hist., Informe 10. Mexico. 1960 La Quemada Chalchihuites. Guía oficial del Inst. Nac. Antr. Hist. NOMLAND, G.

1932

A.

Proboscis statue from the isthmus of Tehuantepec. Amer. Anthr., 34: 591-93.

NOTES AND NEWS

1946

Notes and news: Middle American area. Amer. Antiquity, 11: 207-08.

NOWOTNY, K.

n.d. 1777

F.

Los Franciscanos en las provincias internas de Sonora y Ostimuri. Mexico.

O'CROULEY, P. A.

n.d.

Ydea compendiosa del reyno de Nueva España. Photostat. MS in Sauer Coll., Bib. Nacional, Madrid.

O ' N E A L E , L.

1948

1935

M.

Textiles of pre-Columbian Chihuahua. Carnegie Inst. Wash., Pub. 574, Contrib. 45.

See Fernández de Oviedo y Valdés. O W E N , R.

1956

Geografía de las lenguas y carta etnografía de Mexico. Mexico.

ORTIZ, F.

1947

340

El huracán, su mitología y sus símbolos. Mexico.

C.

Some clay figurines and Seri dolls from coastal Sonora, Mexico. Kiva, 21 (3, 4 ) : 1-11.

PALLEGOIX

1854

Description du royaume Thai ou Siam. Vol. 1. Paris.

PARMENTIER,

1922 1954 1928 1939 PELLIOT,

1925

1920

C.

Spanish elements in the kachina cult of the Pueblos. Proc. 23d Int. Cong. Amer. (New York), pp. 582-603. Pueblo Indian religion. 2 vols. Univ. Chicago Press. P.

Quelques textes chinois concernant I'Indochine hindouisée. In Etudes asiatiques publiées à I'occasion du vingt-cinquième anniversaire de I'Ecole Française d'Extreme-Orient, 2: 243-63.

PEPPER, G.

1906

H.

Les sculptures chames au Musée de Tourane. Ars Asiática, vol. 4. Paris. L'art du Laos. Vol. 1. Paris and Hanoi.

PARSONS, E.

H.

Human effigy vases from Chaco Canyon, New Mexico. In Boas Anniversary Volume, pp. 320-34. Pueblo Bonito. Amer. Mus. Natural Hist., Anthr. Papers, vol. 27.

PERALTA, Μ. M. DE

1883 1890 1898

Costa-Rica Nicaragua y Panama en el siglo XVI. Madrid. Límites de Costa Rica y Colombia. Madrid. Costa Rica y Costa de Mosquitos. Madrid.

PEREZ DE RIBAS, A.

1645

OROZCO Y BERRA, Μ.

1864

The archaeological problem in Chiriqui. Amer. Anthr., 37: 234-43.

OVIEDO Y VALDES

MS.

Carta edificante histórico-curiosa escrita desde la misión de Sta. María de Baserac en los finos de Sonora. In Noticias de varias misiones, vol. 76. Bib. Nac. de Mexico.

Some archaeological notes from southern Hidalgo County, New Mexico. New Mexico Anthr., 3 ( 3 ) : 21-23.

OSGOOD, C.

A.

OCARANZA,

1933

1938

A.

Farben und Weltrichtungen.

NUNEZ, A.

OSBORNE, D., AND A. HAYES

Historia de los triunfos de nuestra santa fee entre gentes las mas barbaras. Madrid.

PFEFFERKORN,

1949

I.

Sonora, a description of the province. Tr. by Τ. Ε. Treutlein. Univ. New Mexico Press.

REFERENCES PHILLIPS, P.

1940

Middle American influences on the archaeology of the southeastern United States. In The Maya and their Neighbors, pp. 349-67.

PORTILLO, Ε.

1886 1897

, J. A. FORD, AND J. B. GRIFFIN

1951

Archaeological survey in the lower Mississippi alluvial valley, 19401947. Papers Peabody Mus., Harvard Univ., vol. 25.

PICOLO, F.

1919

PROSKOURIAKOFF,

1946 1954

Μ.

Report on the state of the new Christianity of California (1701). In Bolton, 1919, pp. 46-67. Dottings on the roadside in Panama, Nicaragua, and Mosquito. London.

PINA CHAN,

1955 1958 1963

PINART, A.

n.d. PITTIER,

1895

R.

Las culturas preclásicas de la cuenca de Mexico. Fondo de Cultura Económica. Mexico. Tlatilco. Inst. Nac. Antr. Hist., Ser. Invest., nos. 1, 2. Cultural development in central Mesoamerica. In Meggers and Evans, 1963, pp. 17-26.

1927

Relación breve y verdadera. Madrid.

POPENOE, D.

1934 1936

1953 1955 1956

H.

Ν.

Pipas precortesianas. Acta Anthr., vol. 3, no. 2. Tlatilco and the pre-classic cultures of the New World. Viking Fund Pub. Anthr., no. 19. Material preclásico de San Salvador. Inst. Tropical Invest. Cien., 4: 10512. San Salvador. Excavations at Chupicuaro, Guanajuato, Mexico. Trans. Amer. Phil. Soc, vol. 46, pt. 5.

PORTER, W.

1932

2 vols.

Some excavations at Playa de los Muertos, Ulua River, Honduras. Maya Research, 1: 61-81. The ruins of Tenampua, Honduras. Smithsonian Inst., ann. rept. 1935, pp. 559-72.

PORTER, Μ.

1948

1953

1957

W.

The Coahuila piedmont, a physiographic province in northeastern Mexico. Jour. Geol, 40: 338-52.

The story of the American Indian. New York. L.

The water lily in Maya art: a complex of alleged Asiatic origin. Smithsonian Inst., Bur. Amer. Ethnol., Bull. 151, pp. 75-153. Comparative notes on the hand-eye and related motifs. Amer. Antiquity, 22: 247-57.

RAVENEAU DE LUSSAN

1869 REED, E.

1951

H.

Exploración en Talamanca. San Jose.

PONCE, A.

1873

RANDS, R.

L.

Vocabularies of the Hehúe dialect of Opata. MS. Pinart Coll., Bancroft Library, Univ. California.

Τ.

An album of Maya architecture. Carnegie Inst. Wash., Pub. 558. Varieties of classic central Veracruz sculpture. Ibid., Pub. 606, Contrib. 58.

RADIN, P.

P I M , CAPTAIN, AND B. SEEMANN

1869

L.

Apuntes para la historia antigua de Coahuila y Texas. Saltillo. Catecismo geográfico político e histórico del estado de Coahuila de Zaragoza. Saltillo.

Journal du voyage fait à la Mer du Sur, par les flibustiers de I'Amérique en 1684 et années suivantes. Paris. K.

Cultural areas of the pre-Spanish Southwest. New Mexico Quar. Rev., 21: 428-39.

REICHEL-DOLMATOFF,

1955 1961 1965

AND A.

1956

G.

Excavaciones en los Conchales de la costa de Barlovento. Rev. Colombiana Anthr., 4: 249-72. Bogota. Puerto Hormiga: un complejo prehistórico marginal de Colombia. Ibid., 10: 349-54. Excavaciones arqueológicas en Puerto Hormiga (Departamento de Bolivar). Antropología, 2. Bogota. REICHEL-DOLMATOFF

Momil: excavaciones en el Sinu. Rev. Colombiana Antr., 5: 111-333.

RELACIÓN BREVE Y VERDADERA

See Ponce, 1873. RELACIÓN POR ESPINOSA

See Espinosa, Gaspar de, 1892. RICHARDSON, F.

1940 1941 1942

B.

Non-Maya monumental sculpture of Central America. In The Maya and their Neighbors, pp. 395-416. Nicaragua. Carnegie Inst. Wash., Year Book 40, pp. 300-02. Some problems relating to the archaeology of southern Central America. Proc. 8th Amer. Sci. Cong., 2: 93-99. 341

REFERENCES ROGERS,

AND Κ . R U P P E R T

1942

Nicaragua. Carnegie Inst. Year Book 41, pp. 269-71.

Wash.,

A cruising voyage around the world (1712). New York.

ROMERO, J.

RlCKETSON, O . G .

1940

W.

1928

An outline of basic physical factors affecting Middle America. In The Maya and their Neighbors, pp. 1 0 31.

1952

Los patrones de la mutilación dentaria prehispánica. An. Inst. Nac. Antr. Hist., 4: 177-221. Mutilaciones dentarias prehispánicas de Mexico y América en general. Inst. Nac. Antr. Hist., Ser. Investigaciones, no. 3.

1958

AND Ε . Β . RlCKETSON

1937

Uaxactun, Guatemala: Group Ε 1926-1931. Carnegie Inst. Wash., Pub. 477.

RIDLEY, F.

1960

Transatlantic contacts of primitive man: eastern Canada and northwestern Russia. Pennsylvania Archaeol., 30: 46-57.

RIES, M.

1932

Stamping: a mass-production printing method 2000 years old. Tulane Univ., Middle Amer. Research Ser., Pub. 4, pp. 411-83.

RIVERA, P. DE

1736

Diario y derrotero.

Guatemala.

RIVET, P.

1909

Recherches anthropologiques sur la Basse-Californie. Jour. Soc. Amer. Paris, vol. 6.

ROBERTS, F. Η.

1945

Η.

The New World Paleo-Indian. Smithsonian Inst., ann. rept. 1944, pp. 403-34.

ROBERTS, O.

1827

W.

ROBLES, C.

ROCK,

1922

La región arqueológica Grandes. Mexico.

1939

1945 1958

342

de

The West Indies. In Handbook of South American Indians, 4: 495-565. 1951 Areas and periods of culture in the Greater Antilles. SW. Jour. Anthr., 7: 248-65. 1953 The circum-Caribbean theory, an archaeological test. Amer. Anthr., 55: 188-200. 1954 Reply to Stern. Ibid., 56: 107-08. 1960a The entry of man into the West Indies. Yale Univ. Pub. Anthr., no. 61. 1960b Recent developments in American archaeology. In Wallace, 1960, pp. 64-73. 1964 Prehistory of the West Indies. Science, 144: 499-513. AND J. M . CRUXENT

1963 1960 1962

Casas

F.

Kalender, Sternglaube und Weltbilder der Tolteken als Zeugen verschollener Kulturbeziehungen zur Alten Welt. Mitteilungen Anthr. Gesellschaft in Wien, 52: 43-136.

ROGERS, Μ.

1948

ROWE, J.

Narrative of voyages and excursions on the east coast and in the interior of Central America. Constable's Miscellany, vol. 17. Edinburgh.

1929

ROUSE, I.

ROYS, R.

1933

H.

Cultural unity and diversification in Peruvian archaeology. In Wallace, 1960, pp. 627-31. Chavin art, an inquiry into its form and meaning. New York Graphic Soc. Greenwich. L.

The book of Chilam Balam of Chumayel. Carnegie Inst. Wash., Pub. 438.

RUBIN, M., AND C. ALEXANDER

1960

J.

Early lithic industries of the lower basin of the Colorado River and adjacent desert areas. San Diego Mus. Papers, no. 3. An outline of Yuman prehistory. SW. Jour. Anthr., 1: 167-98. San Dieguito implements from the terraces of the Rincon-Pantano and Rillito drainage system. Kiva, 24 ( 1 ) : 1-23.

Venezuelan archaeology. Yale Caribbean Ser., no. 6.

U.S. Geological Survey radiocarbon dates, V. Amer. Jour. Sci., Radiocarbon Suppl., 2: 129-85.

RUBÍN DE LA BORBOLLA, D.

1947

F.

Teotihuacán: ofrendas de los templos de Quetzalcoatl. An. Inst. Nac. Antr. Hist., 2: 61-72.

RuIz, Β. 1884 Relación de los primeros descubrimientos de Francisco Pizarro y Diego de Almagro. In Col. Doc. Inéd. para la Hist. España, 5: 193-201.

REFERENCES SALES, L.

1794

Noticias de la provincia de Californias. En tres cartas de un sacerdote religioso hijo del Real Convento de Predicadores de Valencia a un amigo suyo. 3 vols. Valencia.

1936b An archaeological survey of Chihuahua, Mexico. Ibid., no. 22. AND E. ANTEVS

1941 1952

SALMONY, A.

1938 SANDER,

1959 SAPPER,

1902 1904

Carved jade Berkeley.

1934 1935 1952

1931 1932

1909 1910

1913

1956

K.

Mittelamerikanische Reisen und Studien aus den Jahren 1888 bis 1900. Brunswick. Der gegenwärtige Stand der Ethnographischen Kenntnis von Mittelamerika. Archiv für Anthr., 3 1 : 1— 38. Brunswick.

BRAND

Pueblo sites in southeastern Arizona. Univ. California Pub. Geog., 3: 4 1 5 58. Prehistoric settlements of Sonora with special reference to Cerro de Trincheras. Ibid., 5: 67-148. Aztatlan, prehistoric Mexican frontier on the Pacific coast. Ibero-Amer., no. 1. H.

The antiquities of Manabi, Ecuador: final report. Contrib. South Amer. Archaeol., vol. 1. George B. Heye Exped. New York. Archaeological researches on the coast of Esmeraldas. Proc. 16th Int. Cong. Amer. (Vienna). The antiquities of Manabi, Ecuador: final report. Contrib. South Amer. Archaeol., vol. 2. George B. Heye Exped. New York. Precolumbian decoration of the teeth in Ecuador. Amer. Anthr., 15: 3 7 7 94.

SAYLES, E.

B.

1936a Some Southwestern pottery types, Series 5. Medallion Papers, no. 21.

Ibid., no. 29.

H.

A brief survey of the lower Colorado River from Davis Dam to the international border. Mimeographed. U.S. Dept. Interior, Nat. Park Service. Comments on a trial survey of Mexican-Southwestern architectural parallels. El Palacio, 63 ( 9 - 1 0 ) : 2 9 9 309.

SCHWARTZLOSE, R . A .

1952

The cultural geography of the Mormon settlements in Mexico. MS.

SCHWATKA,

1893 1956

F.

In the land of cave and cliff-dwellings.

SEARS, W.

O.

The road to Cibola. Ibero-Amer., no. 3. The distribution of aboriginal tribes and languages in northwestern Mexico. Ibid., no. 5. Aboriginal population of northwest Mexico. Ibid., no. 10. Agricultural origins and dispersals. Amer. Geog. Soc.

SAVILLE, M.

1907

China.

Fluted points from Madden Lake. Panama Archaeol., vol. 2, no. 1.

AND D. D.

1930

ancient

D.

SAUER, C.

1932

in

The Cochise culture.

SCHROEDER, A.

H.

Excavations at Kolomoki: final report. Univ. Georgia Press.

AND J. B. G R I F F I N

1950

Fiber-tempered pottery of the southeast; fabric-marked pottery in eastern United States. In J. B. Griffin, ed., Prehistoric pottery of the eastern United States. Mus. Anthr., Univ. Michigan.

SEJOURNÉ,

L.

1956-57 Estudio del material arqueológico de Atetelco, Teotihuacán. Rev. Mex. Estud. Antr., 14: 15-23. SELLARDS, E.

1952

H.

Early man in America. Press.

SERRANO Y SANZ, M.,

1908

Univ. Texas

ed.

Relaciones históricos y geográficos de América Central. Vol. 8. Madrid.

SHELVOCKE, G.

1814

A voyage around the world, 17191722. In Kerr, 1814.

SHEPARD, A.

1948

SHETRONE, H.

1926 SHOOK, E.

1951

O.

Plumbate: a Mesoamerican trade ware. Carnegie Inst. Wash., Pub. 573. C.

Exploration of the Hopewell group. In Certain mounds and village sites in Ohio, vol. 4, pt. 4. M.

The present status of research on the pre-classic horizons in Guatemala. In Tax, 1951, pp. 93-100.

AND A. V. KIDDER

1952

Mound E-III-3, Kaminaljuyu, Guatemala. Carnegie Inst. Wash., Pub. 596, Contrib. 53.

343

REFERENCES

1860

SIMPSON, L. B., tr.

1938

Collection of rare and original documents and relations concerning the discovery and conquest of America chiefly from Spanish archives. No. 1. New York.

California in 1792: the expedition of José Longinos Martínez. San Marino.

SKINNER, A.

1926

Notes on Las Mercedes, Costa Rica Farm and Anita Grande. In Lothrop, 1926, app. 4.

SMITH, H.

1955

STADEN, H.

1928

The true history of his captivity, 1557. London.

STEINMAYER, R. A.

G.

Archaeological significance of oriental porcelain in Florida sites. Florida Anthr., 8: 111-16.

1932

A reconnaissance of certain mounds and relics in Spanish Honduras. Tulane Univ., Middle Amer. Research Rec, Pub. 4, pp. 1-22.

SMITH, R. E.

1958

The place of Fine Orange pottery in Mesoamerican archaeology. Amer. Antiquity, 24: 151-60.

SMITH, V. J.

1938

Carved rock shelter. Bull. Texas Archaeol. Paleontol. Soc, 10: 222-33.

SMITH, W.,

1952

1957

E.

Adena portraiture. In Webb Raby, 1957, pp. 47-60.

and

SORENSON, J. L.

1955

A chronological ordering of the Mesoamerican pre-classic. Tulane Univ., Middle Amer. Research Rec., 2: 4 3 68.

SPAULDING, A. C.

1952 1955

The origin of the Adena culture of the Ohio valley. SW. Jour. Anthr., 8: 260-68. Prehistoric cultural development in the eastern United States. In Meggers and Evans, 1955, pp. 12-27.

SPICER, Ε.

1962

Η.

Cycles of conquest. Press.

1947

STERN, P.

1934

Notes on the archaeology of Salvador. Ibid., 17: 446-87. 1917a Ancient civilizations of Mexico and Central America. Amer. Mus. Natural Hist., Handbook Ser., no. 3. 1917b The origin and distribution of agriculture in America. Proc. 19th Int. Cong. Amer. (Washington), pp. 2 6 9 76. SQUIER, E.

1852

344

G.

Nicaragua.

2 vols.

Evolution du linteau Khmer. Arts Asiatiques, 8: 251-56.

New York.

Rev.

STERN, T.

1948 1954

The rubber-ball games of the Americas. Monogr. Amer. Ethnol. Soc, no. 17. A note on Rouse's "The circum-Caribbean theory, an archaeological test." Amer. Anthr., 56: 106-07.

STEWARD, J.

H.

1946-59 [ed.] Handbook of South American Indians. 7 vols. Smithsonian Inst., Bur. Amer. Ethnol., Bull. 143. AND L. C. FARON

1959

Native peoples of South America. New York.

STEWART, T. D., AND P. F. TITTERINGTON

1944

1943

SPINDEN, H. J.

1915

G.

Cytogenetics of Gossypium and the problem of the origin of New World cottons. In Advances in genetics, 1: 431-42.

Filed Indian teeth from Illinois. Jour. Washington Acad. Sci., vol. 34, no. 10.

STIRLING, M.

The place of Tajin in Totonac archaeology. Amer. Anthr., 35: 225-70.

E.

Foreword to Ancient art from Costa Rica. Scripps College.

STEPHENS, S.

Univ. Arizona

SPINDÉN, Ε. S.

1933

1952

AND L. EWING

Kiva mural decorations at Awatovi and Kawaika-a: with a survey of other wall paintings in the Pueblo Southwest. Papers Peabody Mus., Harvard Univ., vol. 37.

SNOW, C.

STENDAHL, A.

W.

Stone monuments of southern Mexico. Smithsonian Inst., Bur. Amer. Ethnol, Bull. 138.

STONE, D. Z.

1938 1941 1943

Masters in marble. Tulane Univ., Middle Amer. Research Ser., Pub. 8,

pt. 1.

Archaeology of the north coast of Honduras. Mem. Peabody Mus., Harvard Univ., vol. 9, no. 1. A preliminary investigation of the flood plain of the Rio Grande de Terraba, Costa Rica. Amer. Antiquity, vol. 9, no. 1.

REFERENCES

The Borucas of Costa Rica. Papers Peabody Mus., Harvard Univ., vol. 26, no. 2. 1951 Orfebrería pre-columbina. Mus. Nac. San Jose. 1956a Breve esbozo etnológico de los pueblos indígenas costarricenses. In Estudios Antropológicos, 503-11. 1956b Date of maize in Talamanca: an hypothesis. Jour. Soc. Amer. Paris, 45: 189-94. 1957 The archaeology of central and southern Honduras. Papers Peabody Mus., Harvard Univ., vol. 49, no. 3. 1958 Introduction to the archaeology of Costa Rica. Mus. Nac. San Jose. 1959 The eastern frontier of Mesoamerica. Mitteilungen aus dem Mus. für Volkerkunde in Hamburg, 25: 118-21. 1962 The Talamancan tribes of Costa Rica. Papers Peabody Mus., Harvard Univ., vol. 43, no. 2.

ern Honduras, 1936. Smithsonian Misc. Coll., vol. 97, no. 1.

1949

AND C. BALSER

1958

The aboriginal metalwork of the isthmian region of America. Mus. Nac. San Jose.

STONOR, C. R., AND E. ANDERSON

1949

Maize among the hill peoples of Assam. Ann. Missouri Bot. Garden, 36: 355-404.

STRANGEWAYS, T.

1822

Sketch of the Mosquito shore, including the territory of Poyais. Edinburgh.

STRONG, W.

1925 1934

1935

1948

D.

The Uhle pottery collections from Ancon. Univ. California Pub. Amer. Archaeol. Ethnol, 2 1 : 135-90. Hunting ancient ruins in northeastern Honduras. In Explorations and field work of the Smithsonian Institution in 1933, pp. 44-48. Archeological investigations in the Bay Islands, Spanish Honduras. Smithsonian Misc. Coll., vol. 92, no. 14. The archeology of Honduras. In Handbook of South American Indians, 4: 71-120.

AND C. EVANS, JR.

1952

Cultural stratigraphy in the Viru valley, northern Peru. Columbia Studies in Archaeol. Ethnol., vol. 4.

A. KIDDER, AND A. J. D. PAUL, JR.

1938

Preliminary report of the Smithsonian Institution - Harvard University archeological expedition to northwest-

STUDLEY, C.

1884

Notes upon human remains from caves in Coahuila. Peabody Mus., Harvard Univ., 16th ann. rept., pp. 233-59.

STUIVER, M., E. S. DEEVEY, AND L. J. GRALENSKI

1960

Yale natural radiocarbon measurements, V. Amer. Jour. Sci., Radiocarbon Suppl., 2: 49-61.

STURTEVANT, W.

1960

C.

The significance of ethnological similarities between southeastern North America and the Antilles. Yale Univ. Pub. Anthr., no. 64.

SUHM, D. Α., AND A. D. KRIEGER

1954

An introductory handbook of Texas archaeology. Bull. Texas Archaeol. Soc, vol. 25.

SULLIVAN, L. R., AND M. HELLMAN

1925

The Punin calvarium. Natural Hist., Anthr. 309-37.

Amer. Mus. Papers, 23:

SWADESH, M.

1959

Indian linguistic groups of Mexico. Escuela Nac. Anthr. Hist. Mexico.

SWAUGER, J. L., AND W . J. M A Y E R - O A K E S

1952

A fluted point from Costa Amer. Antiquity, 17: 264-65.

Rica.

TASSIN, W.

1902

The Casas Grandes meteorite. Nat. Mus., 25: 69-74. Washington.

TATUM, J. L.

1931 S., 1951

TAX,

1952 1960

General geology of northeast Mexico. Amer. Assoc. Petroleum Geol., Bull. 15, pp. 867-93. ed. The civilizations of ancient America. Selected papers, 29th Int. Cong. Amer. (New York). Indian tribes of aboriginal America. Ibid. Evolution after Darwin. Chicago.

TAYLOR, D.

1951

M.

The Black Carib of British Honduras. Viking Fund Pub. Anthr., no. 17.

TAYLOR, H. C , JR.

1948

An archaeological reconnaissance in northern Coahuila. Bull. Texas Archaeol. and Paleontol. Soc, 19*. 7 4 87.

TAYLOR, W. P., W. B. MCDOUGALL, C. PRESNALL, AND K. P. SCHMIDT

n.d.

C.

Preliminary ecological survey of the northern Sierra de Carmen, Coahuila, 345

REFERENCES

Mexico. Mimeographed. U.S. Dept. Interior, Fish and Wildlife Service. TAYLOR, W.

n.d.

1948 1956

1961 1964

W.

The occupation and utilization of arid lands: ancient and modern cultures in north Mexico. Paper presented to Sec. Anthr., Illinois Acad. Sci., 1960. A study of archeology. Mem. Amer. Anthr. Assoc, no. 69. Some implications of the Carbon-14 dates from a cave in Coahuila, Mexico. Bull. Texas Archaeol. Soc, 27: 215-34. Archaeology and language in western North America. Amer. Antiquity, 27: 71-81. Tethered nomadism and water territoriality: an hypothesis. Acts 35th Int. Cong. Amer. (Mexico), pp. 197-203.

1958a Surface survey of the northern valley of Mexico: the classic and post-classic periods. Trans. Amer. Phil. Soc, vol. 48, pt. 5. 1958b The archaeology of the Lena basin and its New World relationships. Pt. 1: Amer. Antiquity, 23: 397-418; pt. 2: ibid., 24: 63-81. TORQUEMADA, J . DE

1943

TOWER, D.

1945

1907 1941

Blood groups of the prehistoric Indians of Coahuila by serological tests of their mummified remains. Year Book Amer. Phil. Soc., 1943, pp. 178-80.

AND F. GONZÁLEZ RUL

1960

Archaeological reconnaissance behind the Diablo Dam, Coahuila, Mexico. Bull. Texas Archaeol. Soc, vol. 31.

TEN KATE, H. F.

1884

1930

1951

C.

Materioux pour servir a l'anthropologie de la Pres-qu'tile Californienne. Bull. Soc. Anthr. Paris, 3: 551-69.

THOMPSON, J. E.

S.

Ethnology of the Mayas of southern and central British Honduras. Field Mus. Natural Hist., Anthr. Ser., vol. 17, no. 2. Canoes and navigation of the Maya and their neighbors. Jour. Royal Anthr. Inst., 79: 69-78.

THOMPSON, R.

H.

1958a [ed.] Migrations in New World culture history. Univ. Arizona, Social Sci. Bull, no. 27. 1958b Modern Yucatecan Maya pottery making. Mem. Soc. Amer. Archaeol., no. 15. TOKYO SCIENTIFIC EXPEDITION

1960

Andes. Rept. Univ. Tokyo Sci. Exped. to the Andes, 1958. Tokyo.

TOLSTOY, P.

1953

346

1947

Some Amerasian pottery traits in north Asian prehistory. Amer. Antiquity, 19: 25-39.

Mexi-

B.

M.

A comparative study of the Mayas and the Lacandones. New York. Landa's Relación de las Cosas de Yucatan. Papers Peabody Mus., Harvard Univ., vol. 18.

TREGANZA, A.

1942

3 vols.

The use of marine Mollusca and their value in reconstructing prehistoric trade routes in the American Southwest. Papers Excavators Club, vol. 2, no. 3.

TOZZER, A.

AND W. C. BOYD

1943

Monarquia indiana. co.

E.

An archaeological reconnaissance of northeastern Baja California and southeastern California. Amer. Antiquity, 8: 152-63. Notes on the San Dieguito lithic industry of southern California and northern Baja California. In Heizer and Lemert, pp. 253-55.

TUOHY, D. R., AND W. C. MASSE Y

n.d. TYLOR, E.

1879

1896

Coiled basketry from central Baja California and its affiliations. MS. B.

On the game of patolli in ancient Mexico, and its probable Asiatic origin. Jour. Anthr. Inst. Great Britain and Ireland, 8: 116-29. On American lot-games, as evidence of Asiatic intercourse before the time of Columbus. Internat. Archiv für Ethnog., suppl. to vol. 9, pp. 55-67.

U H L E , M.

1906

Bericht iiber die Ergebnisse meiner. Südamerikanischen Reise. Proc. 14th Int. Cong. Amer. (Stuttgart), pp. 567-79. 1922 Influencias mayas en el alto Ecuador. Bol. Acad. Nac Hist., 4: 205-40. 1923a Civilizaciones mayöides de la costa pacífica de Sudamérica. Ibid., 6: 87-92. 1923b Toltecas, Mayas y civilizaciones sudamericanas. Ibid., 7: 1-33. 1927 Estudios Esmeraldeños. An. Univ. Central, 39: 1-61. Quito.

REFERENCES

1931

Las antiguas civilizaciones de Manta. Bol. Acad. Nac. Hist., 12: 5-72.

ULLOA, F.

1925

Memorial and relation . . . to isla de la Cedros (1540). In Wagner, 1925.

UMEHARA,

1936

S.

Étude des bronzes des royaumes combatíants. Kyoto.

VAILLANT, G.

1930 1931 1932 1934 1935 1937 1941

Compendium and description of the West Indies, 1629. Tr. by C. U. Clark. Smithsonian Misc. Coll., no. 102.

VÁZQUEZ, F .

1937-44 Crónica de la provincia del santísimo nombre de Jesus de Guatemala de el orden de N. Seráfico Padre San Francisco en el reyno de la Nueva España. 4 vols. Guatemala. VENEGAS, M., AND A. M. BURRIEL

1943

Noticia de la California y de su conquista temporal y espiritual hasta el tiempo presente. 3 vols. Mexico City. (Madrid, 1757).

VERNEAU, R., AND P. RIVET

1912

Ethnographie ancienne de I'Equateur. Vol. 6. Paris.

VERRILL, A.

1927

1937

H.

Excavations in Cocle province, Panama. Mus. Amer. Indian, Heye Found., Indian Notes, vol. 4, no. 1.

VESSBERG,

1929-30 Le makara dans la sculpture de Rev. Arts Asiatiques, 6: rinde. 133-47. Paris. VOGLER, Ε . Β .

1949

C.

Excavations at Zacatenco. Amer. Mus. Natural Hist., Anthr. Papers, vol. 32, pt. 1. Excavations in Ticoman. Ibid., vol. 32, pt. 2. Some resemblances in the ceramics of Central and North America. Medallion Papers, no. 12. The archaeological setting of the Playa de los Muertos culture. Maya Research, 1: 87-100. Excavations at El Arbolillo. Amer. Mus. Natural Hist., Anthr. Papers, vol. 35, pt. 2. History and stratigraphy in the Valley of Mexico. Sci. Monthly, 44: 307-24. Aztecs of Mexico. New York.

VÁSQUEZ DE ESPINOZA, A .

1942

VOGEL, J. P.

W., M. 1752

1934

Vida portentosa del americano septentrional apostol el V. P. Fr. Antonio Margil de Jesus. Mexico.

A new voyage and description of the isthmus of America. Hakluyt Soc, ser. 2, vol. 73.

WAGNER, H.

1924 1925

R.

The voyage of Sebastian Rodriguez Cermenno. California Hist. Soc. Quar., vol. 3, no. 1. California voyages: 1539-1541. San Francisco.

WALLACE, A. F. C ,

1960 1960

W.

A Hohokam platform mound at the Gatlin site, Gila Bend, Arizona. Amer. Antiquity, 26: 244-62.

AND J. E.

1959

1950 1956 W E R R , C.

1960

OFFICER

Report on the Yecora trip, June 1 7 22, 1959. MS. Arizona State Mus., Univ. Arizona.

WAUCHOPE,

1948

ed.

Men and cultures. Selected papers, 5th Int. Cong. Anthr. Ethnog. Sci.

WASLEY, W.

R.

Excavations at Zacualpa, Guatemala. Tulane Univ., Middle Amer. Research Inst., Pub. 14. A tentative sequence of pre-classic ceramics in Middle America. Ibid., Pub. 15, pp. 211-50. [ed.] Seminars in archaeology: 1955. Mem. Soc. Amer. Archaeol., no. 11. H.

A review of northeast Texas archaeology. In Jelks, Davis, and Sturgis, 1960, pp. 35-62.

W E R R , W. S., AND R. BARY

1957

VlLAPLANA, H . DE

1763

The Mosqueto Indian and his golden river, being a familiar description of the Mosqueto kingdom in America (written about 1699). In A collection of voyages and travels, 6: 2 7 9 312.

W A F E R , L.

B.

Un bronze du style Houai, découvert à Rome. Mus. Far Eastern Antiquities, Bull. 9, pp. 127-31. Stockholm.

De monsterkop uit het omlijstingsornament van tempeldoorgangen enkunst. Leyden. nissen in de Hindoe-Javaanse bouw-

The Adena people, no. 2. State Univ. Press.

Ohio

AND C. E. SNOW

1945

The Adena people. Univ. Kentucky Repts. Anthr. Archaeol., vol. 6.

347

REFERENCES W E S T , R.

1961

AND P. PHILLIPS

C.

Aboriginal sea navigation between Middle and South America. Amer. Anthr., 63: 133-35.

W H E A T , J.

1957

W.

C.

Tombs of old Lo-Yang.

Shanghai.

W l C K E , C . R., AND M . BULLINGTON

1960

WILLIAMS,

1952

A possible Andean influence in central Mexico. Amer. Antiquity, 25: 603-05.

WILLEY, G.

R.

1955a The interrelated rise of the native cultures of Middle and South America. In Meggers and Evans, 1955, pp. 28-45. 1955b The prehistoric civilization of nuclear America. Amer. Anthr., 57: 5 7 1 93. 1956a [ed.] Prehistoric settlement patterns in the New World. Viking Fund Pub. Anthr., no. 23. 1956b Review of Tlatilco and the pre-classic cultures of the New World, by Μ. Ν. Porter. Amer. Antiquity, 22: 88-89. 1958 Estimated correlations and dating of South and Central American culture sequences. Amer. Antiquity, 23: 353-78. 1960 New World prehistory: the main outlines of the pre-Columbian past are only beginning to emerge. Science, 131: 73-86.

1932

1959

An archaeological classification of culture contact situations. In Wauchope, 1956, pp. 5-30. Early Ancon and early Supe culture. Columbia Studies in Archaeol. and Ethnol., vol. 3.

1935

348

The Monagrillo culture of Panama. Papers Peabody Mus., Harvard Univ., vol. 49, no. 2.

R.

Sons of the shaking earth.

Chicago.

WOODBURY

Prehistoric skeletal remains from the Texas coast. Medallion Papers, no. 18.

WOODBURY, R. B.

1961

Prehistoric agriculture at Point of Pines, Arizona. Mem. Soc. Amer. Archaeol., no. 17.

AND A. S. TRIK

1953

The ruins of Zaculeu, United Fruit Co.

WOODWARD,

1936 1957 J. 1938

Guatemala.

A.

A shell bracelet manufactory. Antiquity, 2: 117-25.

WORMINGTON, Η.

Amer.

Μ.

Ancient man in North America. Denver Mus. Natural Hist.

YDE,

An archaeological reconnaissance of northwestern Honduras. Tulane Univ., Middle Amer. Research Inst., Pub. 8.

ZEVALLOS MENÉNDEZ, C ,

1960

AND O.

HOLM

Excavaciones arqueológicas en San Pablo. Ed. Casa de la Cultura Ecuatoriana, informe preliminar. Guayaquil.

ZIMMER, H.

1955

The art of Indian Asia. Ser., no. 39.

Bollingen

ZlNGG, R . M .

1940

AND C . R . M c G l M S E Y

1954

Notes on the history and symbolism of the Muskhogeans and the people of Etowah. Phillips Acad., Dept. Archaeol., Etowah Papers, no. 1.

WOODBURY, G., AND E.

AND J. M . CORBETT

1954

H.

Geologic observations on the ancient human footprints near Managua, Nicaragua. Carnegie Inst. Wash., Pub. 596, Contrib. 52.

W O L F , E.

AND OTHERS

1956

STODDARD

Cultural stratigraphy in Panama: a preliminary report on the Giron site. Ibid., 19: 332-43.

WlLLOUGHBY, C . C .

Identification and significance of the cucurbit materials from Huaca Prieta, Peru. Amer. Mus. Novitates, no. 1426, pp. 1-15.

W H I T E , W.

1934

1954

Archaeological Cucurbitaceae from a cave in southern Baja California. SW. Jour. Anthr., 13: 144-48.

AND J. B. BIRD

1949

Negative-painted pottery from Crystal River, Florida. Amer. Antiquity, 10: 173-85.

AND T. L.

B.

1948-49 A double-walled jar from Chihuahua. Kiva, 14 ( 1 - 4 ) : 8-10. 1955 Mogollon culture prior to A.D. 1000. Mem. Soc. Amer. Archaeol., no. 10. WHITAKER, T.

1944

Report on archaeology of southern Chihuahua. Univ. Denver, Center of Latin Amer. Studies, no. 1.

ZUÑIGA, I .

1835

Rápida ojeada al estado de Sonora. Mexico.

INDEX

Abasolo points: in northeastern Mexico, 67, 89, 90 abrading tools: in northeastern Mexico and Texas, 91; in northwestern Mexico, 107; in El Salvador, 151; in Panama, 201-202 Abubaes: language of, 210 Acacia, use of: in northwestern Mexico, 28; in northeastern Mexico, 70 (fig. 8) Acajutla: excavations at, 135, 136 acequias: in northwestern Mexico, 16 n. 17, 2 1 , 22 achiote, use of: in Central America, 220, 229 acorns, use of: in northwestern Mexico, 17 acropolises: in El Salvador, 138; in Honduras, 159 Adena: origin of, 117-118; burial mounds of, 118, 119; domesticated plants of, 119; pipes of, 119, 120; engraved tablets of, 121; status of, 130 Adena Mound: effigy pipe from, 119, 120 adobe, use of: in northwestern Mexico, 7, 14, 2 0 22, 35; in Southwest, 95, 103; in El Salvador, 138, 139, 141, 147, 217; in Honduras, 175, 217; in Mesoamerica, 309, 310 Agalteca: pottery at, 175 agate, use of: in Panama, 207 Agate Basin points: distribution of, 115 agaves, use of: in northwestern Mexico, 33; in northeastern Mexico, 62, 73, 75 (fig. 15), 80 (fig. 2 3 ) , 81 (fig. 2 4 ) ; in Central America, 218, 220 agriculture: in Lower California, 53; in northeastern Mexico, 87, 89, 90, 299, 303; in Southwest, 95, 104; in northwestern Mexico, 99, 100-101, 116; in eastern United States, 117, 124, 125, 130; in Mesoamerica, 130, 266, 267, 303-304; in Central America, 217, 218-219; in West Indies, 235; in Ecuador, 247; in Peru, 266, 267, 299, 303; origins of, in Old and New Worlds, 298-300 Agua Zarca: excavations at, 13 Akwa'ala. See Paipai Alba points: origin of, 128 albinism: in Central America, 209 alcoholic beverages: in Central America, 221, 226, 232 Alexander Pinched: decoration of, 122 alignments, columnar: in Panama, 205, 207 Aljojuca: projectile points at, 128 Alkali Ridge: pottery at, 105 alligator pears. See avocados alligators, representations of: in Northwestern Mexico, 104; in Honduras, 169, 171; in Costa Rica, 189. See also crocodiles Alligator ware: in Costa Rica and Panama, 194 Almagre phase: traits of, 87, 89 Almagre points: in northeastern Mexico, 91 almond, use of: in Central America, 219 altars: in Southwest, 108

Alta Vista phase: pottery of, 103 alter ego motif: in Nicaragua, 182 amaranths: in eastern United States, 116; in Asia, 299, 300 Amaravati: lotus motif at, 284 Amargosa complex: in Lower California, 43, 4 5 47, 5 1 . See also Pinto Basin, Gypsum Cave points amphibians, representations of: in Southwest, 95 Anasazi: Mesoamerican traits of, 95, 97, 100, 105109 passim; culture history of, 105-109 passim; external influences of, 105-106 Ancon: figurines at, 270, 271; pottery at, 272 Angkor Vat: art motifs at, 289 Angostura points: distribution of, 115 Animas phase: in northwestern Mexico, 14 anklets: in Central America, 232 anonas, use of: in Central America, 219 antelope, use of: in northeastern Mexico, 63, 70

(fig. 8)

Antilocapra americana: in Lower California, 4 1 ; in northeastern Mexico, 70 (fig. 8) antler, use of: in northeastern Mexico, 67, 78, 89 ants, use of: in Central America, 223 Anyang: burials at, 307 n. 11 Apache: in northwestern Mexico, 9, 11, 24 Apachian biotic province: in northwestern Mexico, 26 aprons: in Lower California, 53; in northeastern Mexico, 75, 79 (fig. 22) Arawak, culture of: in Central America, 210, 211 n. 1; in West Indies, 234-235, 236, 238-241; in South America, 234, 235 Archaic: in Texas, 91-92; in eastern United States, 111, 113-117, 120, 125, 130 Arctic Small Tool tradition: diffusion of, 120 Arevalo phase: ceremonial centers of, 306, 310 Arid Tropical Zone: in Lower California, 43 armadillo, use of: in Central America, 220 arm bands, gold: in Panama, 207 armor, use of: in Central America, 221, 222 Arrama: language of, 212 arrowpoints: in Lower California, 50, 53; in Texas, 82; in northeastern Mexico, 82, 83, 9 1 ; in Mesoamerica, 128; in eastern United States, 128 arrows: in Lower California, 48 (fig. 9 ) , 50, 5 1 , 53; in northeastern Mexico, 85, 9 1 ; in Southwest, 108; in Central America, 229-230 Arroyo de Malpais: extinct horse at, 17 art, cave: See petroglyphs, pictographs Aspero: maize at, 267 asphaltum, use of: in Ecuador and Mesoamerica, 256 Atacames: pottery at, 256; dental mutilation at, 261, 262

349

INDEX

Atetelco: radiocarbon dates at, 282, 283 Atiquizaya: excavations at, 135 atlantean figures: in Asia and Mesoamerica, 289, 311 atlatls. See spear-throwers avocados, use of: in Central America, 219 Awatovi: murals at, 108 awls, bone: in Lower California, 48 (figs. 8, 9 ) , 49; in northeastern Mexico, 63, 77, 82 (fig. 2 5 ) ; in Southwest, 108 awls, copper: in northwestern Mexico, 103 axes, hand: in northeastern Mexico and Texas, 91 axes, polished stone: in Southwest, 103; in northwestern Mexico, 103; in eastern United States, 115, 120; in Costa Rica, 183-184; in West Indies, 241 Ayala phase: pottery of, 103 Aztatlan horizon: features of, 100, 106-107, 109 Aztec Ruin: copper bells at, 106 Aztecs: migration legend of, 12; pottery of, 259 Aztecoidan languages: in Central America, 210 Azteco-Tanoan languages: in Central America, 210 Babicora phase: in northwestern Mexico, 13-14 Babicora Polychrome: in northwestern Mexico, 13 Babocomari: beads from, 81 Babocomari Polychrome: in northwestern Mexico, 35, 36 bags, fiber: in northeastern Mexico, 75, 78 (fig. 2 0 ) , 79 (fig. 2 2 ) , 90 Baird points: in Texas and northeastern Mexico, 91 Baking Pot: dental mutilation at, 262 Baksei-Chamkrong: temples at, 309 Balcones phase: in Texas, 91-92 ball courts and ball games: in northwestern Mexico, 16 n. 17, 17, 20 (fig. 8 ) , 22, 23; in Southwest, 95, 103; in El Salvador, 143, 147, 151; in Honduras, 161, 175; in historic Central American cultures, 232; in West Indies, 238-241; in Mesoamerica, 240-241, 305 balls, stone: in West Indies, 238, 239 balsas, use of: in Lower California, 53; in Ecuador, 262 balustrades: in Southwest, 106 bands, plaited: in northeastern Mexico, 62, 80

(fig. 23)

bannerstones: in eastern United States, 115 baptism: in Central America, 231 bark, use of: in Lower California, 49, 56; in Central America, 207, 217. See also bark beaters bark beaters, stone: in El Salvador, 136; in Honduras, 174; in Costa Rica, 183. See also bark, use of Barlovento: pottery of, 245, 266, 267 Barranca Tovar: stratigraphy at, 134 (fig. 3 ) , 135, 136 Barriles: investigations at, 197 bars, limestone: in northeastern Mexico, 71 basalt, use of: in northwestern Mexico, 31; in northeastern Mexico, 68, 82; in Ecuador, 245; in Mesoamerica, 309 basketry: in Lower California, 48 (figs. 8, 9 ) , 49, 50, 54, 56, 57; in northeastern Mexico, 74, 79 (fig. 2 1 ) , 84, 85, 90, 115; of Huichol, 102; of

350

western United States, 114; in Central America, 217, 224 Bat Cave: maize at, 299 bats, representations of: in Honduras, 169, 171; in Costa Rica, 184; in Ecuador and Mesoamerica, 256, 259 battue, use of: in Central America, 217 beads, bone: in northeastern Mexico, 70 (fig. 8 ) , 77-78, 83, 85, 86 beads, gold: in Panama, 198, 207; in West Indies, 240 beads, seed: in northeastern Mexico, 70 (fig. 8) beads, shell: in northwestern Mexico, 31, 8 1 , 103; in Lower California, 48 (fig. 8 ) , 49; in northeastern Mexico, 70 (fig. 8 ) , 8 1 , 83; in Southwest, 103; in Central America, 229 beads, stone: in Panama, 198; in historic Central American cultures, 213; in West Indies, 240 beans: in northeastern Mexico, 89, 90, 116, 299, 303; in Southwest, 95; in eastern United States, 124; in Central America, 218-219; in Peru and Mesoamerica, 266 beards, wearing of: in Central America, 223 bears, use of: in northeastern Mexico, 63 bedrock mortars: in northeastern Mexico, 69, 82 beds: in northwestern Mexico, 2 1 ; in Central America, 217; in Mesoamerica, 239, 256, 259; in Ecuador, 243, 256, 259 bells: in northwestern Mexico, 23, 103; in Southwest, 95-97 passim, 103-106 passim, 108; in Honduras, 162; in historic Central American cultures, 222; in Mesoamerica, 305 Berlin tablet: designs on, 121 beverages: in Central America, 221 bifaces: in Lower California, 44; in northeastern Mexico, 67, 9 1 ; in Texas, 9 1 , 92 Big Bend Aspect: traits of, 92 bijagua, use of: in Central America, 216, 224 biotic provinces: in northwestern Mexico, 26, 28; in Lower California, 4 1 - 4 3 ; in northeastern Mexico, 60 birds, representations of: in Southwest, 85, 108; in northwestern Mexico, 104, 107; in eastern United States, 115-116, 121; in Mesoamerica, 121; in Honduras, 169; in Costa Rica, 183; in Panama, 200, 201; in Peru, 273 birds, use of: in Central America, 220, 229 birdstones: in eastern United States, 115-116 bison, use of: in northeastern Mexico, 63 Bixa orellana, use of: in Central America, 220 Black- and Red-line ware: in Panama and Costa Rica, 203, 205 Black-line ware: in Panama and Costa Rica, 197, 201, 203 blades: in Lower California, 44, 45, 46, 49; in northeastern Mexico, 67, 90, 9 1 , 92; in Mesoamerica, 115, 120; in eastern United States, 120, 124; in El Salvador, 135, 151, 153; in Panama, 207; in West Indies, 238; in Ecuador, 245 blankets. See robes bloodletting: in northeastern Mexico, 77; in Central America, 231 blood types: in northeastern Mexico, 85 blowguns, use of: in Central America, 217

INDEX

boats: in Lower California, 53; in Central America, 224, 226, 235-236; in West Indies, 236 boatstones: in northeastern Mexico and Texas, 91 body painting: in Central America, 219, 233 Bolanos culture: survivals of, 100 Bold Geometric: in Honduras, 171, 174 bone, use of: in northwestern Mexico, 23; in Lower California, 48 (figs. 8, 9 ) , 46, 49; in northeastern Mexico, 60, 63, 67, 68, 70 (fig. 8 ) , 77-78, 82 (fig. 2 5 ) , 83, 85, 86, 89, 115; in Southwest, 95, 106; in Central America, 223, 229, 230, 232 Bonham points: origin of, 128 books: in historic Central American cultures, 272 Borjeño: language of, 52 Boruca languages: in Central America, 210, 211 bow and arrow: in Lower California, 53; in northeastern Mexico, 68, 84, 86; in Southwest, 108; in eastern United States, 128; in Mesoamerica, 128; in Central America, 217, 218. See also arrowpoints, arrows bowls, stone: in Southwest, 103; in northwestern Mexico, 103; in Mesoamerica, 115; in El Salvador, 136 bracelets, gold: in Central America, 222, 232 brachycephalization: in prehistoric North America, 117-118 Brangenburg Mound: gourds at, 119 Bravo Valley aspect: arrowpoints of, 82, 83 breast bars and plates, gold: in Panama, 207; in historic Central American cultures, 222 breechcloths. See loincloths Bribri: language of, 210, 211 bronze, use of: in China, 278, 279, 280, 312-313 Brownsville points: in northeastern Mexico, 91 Buena Fe phase: traits of, 21 Buena Vista: Valdivia culture at, 245 bull roarers: in Lower California, 48 (fig. 9) Bulverde points: in northeastern Mexico and Texas, 91 burials: in northwestern Mexico, 33, 36; in southern California, 44; in Lower California, 47, 49, 50, 5 1 ; in northeastern Mexico, 72 (fig. 11), 73, 77, 78, 8 1 , 83-87 passim, 9 1 ; in Texas, 9 1 ; in Southwest, 103, 108; in eastern United States, 117-119 passim; in El Salvador, 139, 141, 148; in Honduras, 162, 163, 164; in Costa Rica, 193, 197, 198; in historic Central American cultures, 226-227; in Ecuador, 259-260, 263; in Colombia, 260; in Mesoamerica, 260, 307 n. 11; in China, 307 n. 11 burial sticks: in northeastern Mexico, 73, 86 burins: in northeastern Mexico, 69; in Ecuador, 245 Burucaca: language of, 210 butterflies, representations of: in Southwest, 108 buttons, bone: in northeastern Mexico, 70 (fig. 8 ) ; in Central America, 223 buttons, stone: in northeastern Mexico, 70 (fig. 8) Cabaret complex: traits of, 238 Cabécar: language of, 210, 211 cabuya, use of: in Central America, 224 cacao, use of: in Central America, 218, 220, 221, 229, 230

Cacaopera-Matagalpa: language of, 212 n. 17 caches: in northwestern Mexico, 3 1 ; in northeastern Mexico, 73; in El Salvador, 137 (fig. 5 ) , 138, 141, 147; in Costa Rica, 193 cacti, use of: in Lower California, 51-52; 54, 56 Caddoan area: ceramics of, 127; projectile points of, 128; pipes of, 129 Caddoan root: in northwestern Mexico, 13-14 Cadegomeño: language of, 52 Caguama Cave: excavations at, 50 Cahokia: dental mutilation at, 129 calabashes. See squashes, pumpkins Calamuya: investigations at, 175 Calathea insignis, use of: in Central America, 216 calendars, ceremonial: in Southwest, 95 calendrical systems: in Mesoamerica, 116, 2 9 1 292; in Central America, 233; in Asia, 291, 292 Calera phase: traits of, 100, 106, 109 California Yumans: culture of, 52-55 passim, 57 Callejon de Huayles: sculpture at, 273 Campana-San Andres: excavations at, 138, 139, 141, 143, 147, 151 Canadian zone: in northeastern Mexico, 60 canals, irrigation: in northwestern Mexico, 21 Candelaria Cave: culture of, 83 cane, use of: in Lower California, 48 (fig. 9 ) , 5 0 51, 53, 55; in Central America, 216-217, 2 2 9 232 passim. See also wood Canis latrans, use of: in northeastern Mexico, 70

(fig. 8)

candeleros: in Honduras, 173 Canelo Brown-on-yellow: in northwestern Mexico, 35 cannibalism: in Central America, 220, 229, 231; in West Indies, 235, 236 canoes. See boats Canutillo culture: traits of, 102, 105; influence of, on Hohokam, 102; influence of, on Mogollon, 105 Canutillo-Chalchihuites cultures: traits of, 98, 103, 104; survivals of, 100; influence of, on Mogollon, capes: in Lower California, 48 (fig. 9 ) , 5 1 , 53, 54; in Ecuador and Mesoamerica, 256, 259 Caribs, culture of: in Central America, 212, 224; in West Indies, 234-238 passim; in South America, 234 Carica papaya, use of: in Central America, 219 Carretas phase: traits of, 14 Carretas Polychrome: in northwestern Mexico, 7 carrying frames: in Central America, 223 Casador period: in northwestern Mexico, 17 Casas Grandes culture: traits of, 3 n. 3, 4, 8 (fig. 4 ) , 9, 10, 11 n. 13, 12, 14, 16, 20, 21, 100, 107, 109; evidences of, in Sonora, 35, 36; external influences of, 106 Catalina phase: pottery of, 248 Catan complex: in northeastern Mexico, 90 Catan points: in northeastern Mexico, 91 cats, representations of: in northwestern Mexico, 104; in Mesoamerica, 121; in eastern United States, 121; in Honduras, 169; in Peru, 273. See also jaguars, mountain lions

351

105

INDEX

caves, use of: in northwestern Mexico, 7, 12, 13, 16-18 passim, 29, 35, 36; in Lower California, 47, 49, 50; in northeastern Mexico, 62-64 passim, 84, 86; in western United States, 114; in Honduras, 162; in Panama, 201; in historic Central American cultures, 230. See also rock shelters celts: in eastern United States, 120, 125; in eastern Mexico, 120; in Honduras, 174; in Panama, 198, 207; in West Indies, 237, 241 Central Texas aspect: arrowpoints of, 82 Cerquin: investigations at, 177 ceremonials: in Lower California, 54-55; in northeastern Mexico, 77, 78; in Southwest, 95-97 passim, 104, 106, 107-109; in northwestern Mexico, 107, 108; in eastern United States, 117-119 passim; in Central America, 224-227 passim, 2 3 0 233 passim; in Mesoamerica, 306-310 Cerrito Blanco: excavations at, 46—47 Cerro Cuevosa: excavations at, 47, 49 Cerro de las Mesas: pottery of, 259; dental mutilation at, 262 Cerro de las Remojadas: dental mutilation at, 262 Cerro del Zapote: excavations at, 134 (fig. 3 ) , 135, 138, 141 Cerro Mangote: preceramic remains at, 201, 202 Cerro Sechin: sculpture at, 273 chacuacos. See pipes chains, copper: in northwestern Mexico, 203 chalcedony, use of: in Costa Rica, 183-184 Chalchihuites culture: traits of, 99, 100, 102-105 passim, 109. See also Canutillo, CanutilloChalchihuites Cham: voyages of, 284; art of, 287; vaults of, 294 Chame: pottery at, 203 Chametla culture: traits of, 100, 102 Chan Chan: cut clay mosaics at, 275 Changuena: language of, 210, 211 Chaupihuaca: burials at, 260 Chavin de Huantar: as cult center, 303, 309, 310; masonry at, 309 Chavin culture: pottery of, 248, 268, 270-273 passim; agriculture of, 303; early temple centers of, 308; sculpture of, 310-311; gold work of, 311; Chinese contacts of, 313-314 chenopodium: in eastern United States, 116 Chiapa de Corzo: pottery of, 248, 249, 250, 253, 254, 269, 271; dental mutilation at, 262 Chibchan languages: of Central America, 210 Chicanei phase: pottery of, 249, 273; temple centers of, 309 Chichen Itza: gold ornaments at, 197; lotus motif at, 284, 285, 287; mohara at, 287; atlantean figures at, 311 Chichimecs: definition of, 59; in Central America, 211 Chihuahua culture: Mesoamerican traits of, 100 Chihuahua polychromes: distribution of, in northwestern Mexico, 7, 13, 23, 34 Chihuahua-Zacatecas zone: in Mexico, 60 Chila White: relationships of, 250 chili pepper: in northeastern Mexico, 89, 116; in Southwest, 95; in Central America, 219, 220; in Ecuador, 254. See also pepper

352

Chinautla Polychrome: in Honduras, 161 Chira complex: pottery of, 268, 269 Chiricahua Apaches: in northwestern Mexico, 11 Chiricahua stage: in northwestern Mexico, 29 Chiriqui Alligator ware: in Panama, 205 Chiriqui Red-line ware: in Costa Rica and Panama, 197 Chirripo: language of, 211 Chiru: language of, 210 chisels: in Panama, 198 Chisos focus: in Texas, 92 Choco: language of, 210 Cholula: projectile points at, 128 Cholula Polychrome: influence of, in Peru, 274— 275 Chombo phase: pottery of, 248, 249, 250, 253 Chontal: language of, 212, 214 n. 26; hairdress of, 223; trade of, 229 Chontales: sculpture at, 181 (fig. 1) choppers: in northwestern Mexico, 29; in Lower California, 44, 48 (fig. 9 ) , 50; in northeastern Mexico, 63, 68 (fig. 6 ) , 69, 9 1 , 115; in Texas, 91; in western United States, 114; in Mesoamerica, 115; in Panama, 201; in Ecuador, 245 chopping tools: in southern California, 44, 49 Chorotega: physical features of, 209; language of, 210, 211 n. 2, 213, 214; towns and houses of, 215, 217; foods of, 220; ornaments and hairdress of, 223; textiles of, 224; social and political organization of, 225, 228; burials of, 2 2 6 227; trade of, 229; warfare of, 230; medical practices of, 231; calendar of, 233 Chorotega-Mangue: language of, 210, 212, 214 n. 23; sites attributed to, 215 n. 29; foods of, 219, 220; religion of, 220, 221, 230, 231; clothing of, 221, 222; social organization of, 228 Chorrera phase: traits of, 247-249 passim, 253, 269, 271 Chorrera-Tejar phase: pottery of, 247, 250 Chorti: in El Salvador, 134 Chuchures: language of, 210, 211 Chupadero Black-on-white: in northwestern Mexico, 14, 21 Chupicuaro culture: northward diffusion of, 100; pottery of, 101-102 Chupicuaro Red-on-buff Polychrome: decoration of, 101 Ciboney: distribution of, in West Indies, 234, 235, 236, 239 Cienegas complex: in northeastern Mexico, 62-63, 75 (figs. 15, 16), 80 (fig. 2 3 ) , 82 (fig. 25) cigars, smoking of: in Central America, 221 Cihuatan: excavations at, 143 circlets, gold: in Panama, 207 cliff dwellings: in northwestern Mexico, 12, 13, 16, 35 Cliffton points: in northeastern Mexico, 84 (fig. 27) clothing: in Lower California, 48 (fig. 9 ) , 5 1 , 53, 54; in northeastern Mexico, 62, 73-74, 75, 79 (fig. 2 2 ) , 85, 93; in western United States, 114; in Central America, 221-223; in West Indies, 236; in Ecuador and Mesoamerica, 256, 259

INDEX Clovis points: in northwestern Mexico, 16, 29, 36; in Lower California, 44; in eastern United States, 113; in Costa Rica, 181; in Panama, 181 Club International: discoveries at, 138, 141, 148, 170 clubs, grooved wooden: in northeastern Mexico, 71; in western United States, 114. See also rabbit sticks clubs, stone: in Costa Rica, 183; in Asia and Polynesia, 277-278 Coahuila: culture sequences of, 60-87 Coahuila complex: definition of, 63-81, 82 (fig. 2 5 ) , 83-86 passim, 90, 93 Coahuiltecan languages: of northeastern Mexico, 93-94 Coastal Plain complex: definition of, 84-86 passim, 89, 92 Coatlinchan: pottery at, 143 coatimundi, use of: in northeastern Mexico, 63 coca, use of: in Central America, 219, 221 Cocal phase: in Honduras, 167 Cochími: location and language of, 52 Cochise culture: in northwestern Mexico, 18, 29, 30, 35, 36 Cochise: crania of, 87 Cocle: gold of, 197; stone artifacts of, 197; pottery of, 203, 205 Cocle Polychrome: in Panama, 203 coconuts, use of: in Central America, 218 Cocopa: location of, 28; agriculture of, 53 Cody knife: appearance of, 115 Coles Creek: platform mounds of, 126 collars and yokes, stone: in West Indies and Mesoamerica, 240 columns, square: in Southwest, 106 colonettes: in Asia and Mesoamerica, 289, 291 color symbolism: in Southwest, 95; in Asia and Mesoamerica, 292 Comanche: in northeastern Mexico, 87 combs, use of: in Central America, 223 comelagatoazte: in Central America, 230 Comondu culture: traits of, 4 1 , 48 (fig. 9 ) , 5 0 51, 57 compounds: in northwestern Mexico, 32, 33; in Southwest, 95, 97 Conchagua Vieja: investigations at, 153 Concho: linguistic and cultural relationships of, 11 Conchos phase: in northwestern Mexico, 14; in Guatemala, 248, 249, 250, 253, 254 Conchos White-to-buff: relationships of, 254 confessions: in Central America, 231 containers, water: in Lower California, 53-54 Convento phase: traits of, 18, 19 (fig. 7) Copador Polychrome: in El Salvador, 138, 139, 141, 155, 161; in Honduras, 161; in Nacaragua, 182 Copan: external relationships of, 136, 138, 139, 141, 145, 147, 148, 151, 152, 153, 155, 162, 167, 170, 171, 174, 177, 178; pottery of, 157, 161, 164, 273; sculpture and architecture of, 159; gold ornaments of, 197; dental mutilation at, 262 Copena: stone celts of, 120 copper, use of: in northwestern Mexico, 7, 23,

103, 107; in Southwest, 103; in eastern United States, 116, 119; in El Salvador, 139, 141; in Honduras, 174; in Costa Rica, 193; in Panama, 198-201; in historic Central American cultures, 224; in Ecuador, 263 Cora: location of, 52 cordage: in Lower California, 5 1 , 53, 55; in northeastern Mexico, 70 (fig. 8 ) , 73, 78, 85, 115; in western United States, 114; in Central America, 222, 224, 230 cord-marked pottery: in Mesoamerica, 123, 302; in eastern United States, 123 cores, flint and other stone: in northeastern Mexico, 68-69; in eastern United States, 117, 124; in Ecuador, 245 Coroban: investigations at, 153 Corobici: language of, 210-213 passim; clothing of, 223; marriage patterns of, 225 corpse desiccation: in Central America, 226 Corral Falso: burials at, 260 cosmologies: of Southwest, 98; of Mesoamerica, 98 Coto: language of, 210, 211 Cotorra phase: pottery of, 248, 249, 253 cotton, use of: in northeastern Mexico, 83, 85, 90, 116, 303; in Southwest, 95; in Central America, 217, 218, 221-224 passim, 229, 230; in West Indies, 240; in Old and New Worlds, 299-300 Coumarouna panamensis, use of: in Central America, 219 Couri complex: traits of, 238 courts: See plazas Cowan Creek: radiocarbon date at, 119 Coxcotlan Cave: culture sequence of, 115 coyotes, representations of: in northwestern Mexico, 104 cradles: in northeastern Mexico, 72 (fig. 11), 73, 85, 86; in Ecuador, 243, 259 crania: in northwestern Mexico, 29; in Lower California, 47, 57, 86; in northeastern Mexico, 86; in eastern United States, 117 cranial deformation: in Mesoamerica, 121; in Central America, 225 cremation: in northwestern Mexico, 33; in Southwest, 103; in Honduras, 162; in historic Central American cultures, 226-227 crocodiles, representations of: in Panama, 200, 201, 205. See also alligators Crooks: pottery at, 121 crosses: fiber, in northeastern Mexico, 62 (fig. 2 ) , 80 (fig. 2 3 ) ; in Southwestern ceramics, 108; in Mesoamerican art, 121 Crystal River: pottery at, 124 Cuba: languages of, 210, 211 Cucurbita, use of: in northeastern Mexico, 89, 90, 116; in Peru, 267 Cueva: language of, 210; culture of, 217, 218, 222-224 passim, 227-230 passim cuffs, gold: in Central America, 222 Culebras: maize at, 267, 271; pottery at, 271 Culhuacan: projectile points at, 128 Culiacan culture: traits of, 100, 109 Cuna: language of, 210 Cipisnique phase: pottery of, 248; agriculture of, 303; dating of, 309

353

INDEX

curassow, use of: in Central America, 220 Curayacu: pottery at, 268, 269, 272; figurines at, 270, 271 curing societies: in Southwest, 95 Cyperaceae, use of: in northeastern Mexico, 86 dagolitos: in West Indies, 238 dantas, use of: in Central America, 220, 229, 230 darts, use of: in Central America, 229. See also lances, spears Davis: radiocarbon dates at, 125-126 decoys, use of: in Central America, 217 deer, use of: in Lower California, 49, 5 1 , 53-55 passim; in northeastern Mexico, 78, 83 (fig. 2 6 ) ; in Central America, 217, 220, 222, 232 deities: in Lower California, 55; in Southwest, 95, 98, 108; in Mesoamerica, 98, 102, 109, 129, 236, 288, 291; in eastern United States, 129; in El Salvador, 135; in Nicaragua, 181; in Panama, 200, 201; in historic Central American cultures, 230-231, 233; in West Indies, 235, 236, 238, 241; in Asia, 288, 291 demography: of Lower California, 53, 54; of eastern United States, 118; of Mesoamerica, 304, 308 dental mutilation: in eastern United States, 119; in Mesoamerica, 129, 261-263; in Southwest, 129; in Ecuador, 260-263 dentate stamping: in Mesoamerica, 121-122; in eastern United States, 122 depilatories: in Central America, 223 Desert culture: in northwestern Mexico, 18, 29, 30, 35, 36, 99, 101; in northeastern Mexico, 61; in Texas, 9 1 , 92; in Southwest, 104; in western North America, 114, 115 Desert Hohokam. See Hohokam dewclaws, use of: in northeastern Mexico, 70

(fig. 8)

Diablo phase: traits of, 23, 87, 89 diadems: in West Indies, 240 dibbles. See digging sticks Dieguieño: location of, 52; culture of, 54-55 digging sticks: in northeastern Mexico, 7 1 ; in western United States, 114; in Central America, 218, 219 Dili phase: pottery of, 249, 250, 253 Dipteryxoleifera, use of: in Central America, 219 Discorea, use of: in Central America, 219 discs, gold: in Panama, 198; in historic Central American cultures, 223 discs, stone: of Huichol, 102 dogs, use of: in southern California, 55; in Central America, 217, 220, 229 dogs, representations of: in Panama, 200 Dongson culture: diffusion of, 277, 300 doorways, T-shaped: in northwestern Mexico, 2 1 , 35 Dorasque: language of, 210, 211 Double Butte Cave: effigy pahos from, 108 doughnut stones: in Southwest and northwestern Mexico, 103 drains, subterranean: in northwestern Mexico, 2 1 , 22

354

drilling, of stone: in eastern United States and Mesoamerica, 120 drills, chipped stone: in northeastern Mexico, 69, 91; in Texas, 9 1 ; in Ecuador, 245 drugs, use of: in Central America, 219, 221, 226; in Asia, 293 drums: in northwestern Mexico, 23; in Costa Rica, 193; in historic Central American cultures, 232; in Asia, 312 Duran points: in northeastern Mexico, 65 (fig. 3) dyes, use of: in Central America, 219, 229 eagles, representations of: in Southwest, 108 Early Guañape: pottery of, 268-269 Early Polychrome: in Costa Rica, 184-189, 197, 237 Early Tutishcainyo: pottery of, 273-274 ear ornaments: in eastern United States, 117, 118, 119; in Mesoamerica, 119; in Costa Rica, 189; in historic Central American cultures, 222; in West Indies, 241; in Ecuador, 248 earthworks, ceremonial: in eastern United States, 117 Easter Island: script of, 311-312 eccentric flints: in eastern United States, 128 Edwards Plateau aspect: traits of, 91 effigy vessels, pottery: in Southwest, 106, 128; in eastern United States, 128; in El Salvador, 143, 147, 152; in Honduras, 164; in Panama, 201; in Peru, 272 El Angel: burials at, 260 El Arbolillo: pottery at, 249, 250, 253; agriculture, houses, and burials at, 304; stone artifacts at, 304-305 El Arenal: burials at, 260 El Carmen: burials at, 259 El Cauce: pottery at, 170 El Hatillo ware: in Panama, 205 n. 1 El Inga: artifacts from, 245 El Llano: burials at, 260 El Opeño: burials at, 260 El Paso Brown: in northeastern Mexico, 63, 83, 93 El Paso Polychrome: in northwestern Mexico, 14, 21 El Prisco period: pottery of, 124 El Puerto: investigations at, 159 El Trapiche: pottery at, 121-122, 136, 248 El Triunfo: excavations at, 145, 147 elbow stones: distribution of, in West Indies, 240 elephants, representations of: in Asia and Mesoamerica, 288 elk, use of: in northeastern Mexico, 60 emeralds, use of: in Panama, 207 Escalera phase: pottery of, 249 Escoria: language of, 210, 211 Eslabones phase: pottery of, 127 Esmeraldas: pottery of, 256; dental mutilation at, 261 Espantosa points: in northeastern Mexico, 66 (fig. 4 ) , 67 Esperanza phase: traits of, 143 Estrella phase: pottery of, 101; figurines of, 102 Estrella Red-on-gray: description of, 101

INDEX

ethnohistorical studies: in northwestern Mexico, 9-16 passim, 23, 24, 28, 99; in Lower California, 51-53, 57; in northeastern Mexico, 60, 61, 85, 93-94 Etowah: maize at, 119; knives at, 128 Eudeve: location of, 7 n. 5 Falcon focus: radiocarbon date for, 9 1 , 92 Farone Apache: in northwestern Mexico, 111 Fat Burro Cave: materials from, 62, 63, 65-78 passim, 81-85 passim feathers, use of: in northeastern Mexico, 78; in Southwest, 108; in Central America, 221-223 passim, 229-231 passim; in West Indies, 240; in Ecuador and Mesoamerica, 256, 259 feces, human: in sites of northeastern Mexico, 78, 81 Felis concolor, use of: in northeastern Mexico, 85 fenced fields: in Central America, 219, 227 fertility cults: in Southwest, 95, 104; in Mesoamerica, 104, 109, 256; in Ecuador, 256 festivals, interband: in Lower California, 55 fibers, plant, use of: in northwestern Mexico, 33; in Lower California, 33, 48 (fig. 9 ) , 5 1 , 53, 55; in northeastern Mexico, 62, 63, 67, 71, 73-77, 80 (fig. 2 3 ) , 85, 86; in western United States, 114 fiber-tempered pottery: history and distribution of, 116, 122, 301, 302 figurines: in northwestern Mexico, 31, 35; in Lower California, 54; in Southwest, 101, 102, 118; in Mesoamerica, 102, 124, 256, 259, 305, 306; in eastern United States, 117, 118, 120; in El Salvador, 134, 135, 136, 139, 148; in Honduras, 136, 161-164 passim, 173-174, 191, 193; in Costa Rica, 189, 191, 194; in Panama, 191, 193, 194; in Nicaragua, 191, 193; in historic Central American cultures, 227; in West Indies, 241; in Ecuador, 246, 247, 256, 259; in Peru, 269-270, 306 Fine Line Polychrome: in El Salvador, 148, 151 Fine Orange: distribution of, 174, 178 fire drives: in Central America, 217 fire-making, tools for: in Lower California, 54; in northeastern Mexico, 71-73 passim first-fruits ceremonies: in Lower California, 54, 55 fishes, representations of: in Lower California, 5 1 ; in Peru, 273 fishhooks: in Lower California, 54, 56; in Central America, 174, 218 fishing: in Lower California, 53-57 passim; in eastern United States, 117; in Central America, 215, 217, 218, 220, 224, 229; in West Indies, 234, 236 Flacco phase: traits of, 89, 90 flaking tools: in northeastern Mexico, 78, 82 (fig. 25) floors, prepared, in caves: of northeastern Mexico, 64 Florence Mound: radiocarbon date at, 119 flutes: in Central America, 232 folklore: of Central America, 231-233; of West Indies, 236, 241; of Asia and Mesoamerica, 291, 292

Folsom points: in northwestern Mexico, 29; in eastern United States, 113 food-gathering: in northwestern Mexico, 7, 99, 101; in Lower California, 53, 54; in northeastern Mexico, 63-65, 69, 89, 90, 115; in western United States, 114; in eastern United States, 114-117; in Mesoamerica, 115 food, preparation and preservation of: in Central America, 219-221; in West Indies, 239 food resources: of Lower California, 41-43; of northeastern Mexico, 63-64 footprints, human: in El Salvador and Nicaragua, 180-181 foreshafts, dart: in northeastern Mexico, 63 Forraje period: in northwestern Mexico, 17 fortifications: in El Salvador, 153; in Honduras, 215 n. 28; in Panama, 216; in historic Central American cultures, 228 Fort Walton: pottery of, 122 Four Mile Polychrome: decoration of, 108 foxes, use of: in Lower California, 54 Fragua points: in northeastern Mexico, 65 (fig. 3 ) , 67, 91 Francesca phase: pottery of, 249 Frankston focus: arrowpoints of, 82 Fresno points: in northeastern Mexico, 84 (fig. 27) Frightful Cave: materials from, 62, 63, 65-68 passim, 71-82 passim, 85 fringes, fiber: in northeastern Mexico, 75 Frio points: in northeastern Mexico and Texas, 91 frogs, representations of: in Panama, 200 frogs, use of: in Central America, 220, 223 Frontera phase: traits of, 9 1 , 92 Fugitive Red: in Costa Rica and Panama, 194 Furcraea cabuya, use of: in Central America, 218 Galveston Bay focus: arrowpoints of, 82 Ganesa, representations of: in Asia, 288 Gatlin: ball court at, 103 Georgetown phase: pottery of, 101 Gibson aspect: age of, 125-126 Gila Butte phase: ball courts of, 103 Gila Plain: in northwestern Mexico, 30 Gila polychromes: in northwestern Mexico, 14, 23; origin of, 107 Gila Red: in northwestern Mexico, 36 Gileno Apache: in northwestern Mexico, 11 girdles: in West Indies, 240 Giron: pottery at, 203 glaze paint, use of: in northwestern Mexico, 7 Gliricidia maculata, use of: in Central America, 218 glues, use of: in Central America, 217 Gobernador points: in northeastern Mexico, 65 (fig. 3) god impersonation: in Lower California, 55; in Southwest, 95 gold, use of: in El Salvador, 139, 141; in Nicaragua, 182; in Costa Rica, 184, 193, 196, 197, 207; in Panama, 193, 194, 196-201 passim, 207; in historic Central American cultures, 2 2 1 226 passim, 229, 230, 232; in West Indies, 240,

355

INDEX

241; in Ecuador, 260-263; in Mesoamerica, 261-263, 311; in Peru, 311 goldsmiths: in Central America, 215, 224, 230 gongs: in West Indies, 241 gorgets, shell: in Southwest, 103; in northwestern Mexico, 103; in eastern United States, 116 gouges, Clear Fork: in northeastern Mexico and Texas, 9 1 , 115 gourds, use of: in Lower California, 5 1 ; in northeastern Mexico, 89, 116, 299; in eastern United States, 116, 119, 124; in Central America, 217, 219, 220, 221, 224 granaries: in northwestern Mexico, 35 grass, use of: in Lower California, 53; in Central America, 218, 224 gravers: in northeastern Mexico and Texas, 91 greaves, gold: in Panama, 207 Greenhouse: platform mounds at, 126 grooved rocks: in northeastern Mexico, 69, 71 (fig. 9) Guacamayo ware: in Panama, 201, 202-203 Guaicura: language and culture of, 52-54 passim Guaicururian languages: in Lower California, 5 1 52 Guañape: relationships of, 245 Guanexico: language of, 212 Guangala phase: pottery of, 254-255 Guasave phase: traits of, 36, 96, 106, 107, 109 guasuma, use of: in Central America, 221 Guatemalan Tropical Flint maize: in eastern United States, 118 Guatuso: language of, 210 guavas, use of: in Central America, 219 Guaymi: language of, 210, 211; clothing of, 222; marriage patterns of, 226; social organization of, 228; religion of, 230, 231; games of, 232 Guazuma ulmifolia, use of: in Central America, 221 Guerra phase: traits of, 89, 90 Guetar: language of, 210, 211 Guilielma gasipaes, use of: in Central America, 218 Gypsum Cave points: in Lower California, 45-46, 49, 5 1 , 57 hachas: in El Salvador, 139 hair, use of: in Lower California, 54; in northwestern Mexico, 62, 67 hairdress: in Central America, 223 Haldas: pottery at, 268, 269; figurines at, 271 Haliotis, use of: in Lower California, 48 (fig. 9) hammocks: in Central America, 217, 224, 229; in West Indies, 239-240; in Mesoamerica, 239 Han period: traits of, 283-284, 294 Harness: maize at, 118, 119 Harris Village: ornaments at, 81 harvest ceremonies: in Southwest, 95 Hayes points: origin of, 128 head-dresses: in Lower California, 5 1 ; in northwestern Mexico, 107 head ornaments: in Central America, 222, 223 head rests: in Central America, 217 helmets, gold: in Panama, 207; in historic Central American cultures, 222; in Mesoamerica, 305

356

hematite, use of: in northeastern Mexico, 78 hennequen, use of: in Central America, 218, 224 Henrietta focus: arrowpoints of, 82 Hesperaloe funifera, use of: in northeastern Mexico, 73, 75 (fig. 15) Hibiscus tiliaceus, use of: in Central America, 218 hide processing: in northeastern Mexico, 78 hieroglyphic writing: in Mesoamerica, 311-312 hoes: shell, in northeastern Mexico and Texas, 9 1 ; stone, in eastern United States, 125 Hohokam: external influences of, 13, 23, 30, 3 4 36, 105, 106; Mesoamerican traits of, 96-105 passim Hokaltecan languages: of northeastern Mexico, 93-94; of Central America, 210 Hokan-Siouan languages: of northeastern Mexico, 93-94; in Central America, 210 Holmul: pottery at, 155; dental mutilation at, 262 honey, use of: in Central America, 220; in West Indies, 237 hoof-covers, use of: in northeastern Mexico, 70

(fig. 8), 85

hooks: copper, in northwestern Mexico, 103; harvesting, in Lower California, 48 (fig. 9 ) , 50-51 Hooper Ranch Pueblo: kachina cult at, 108 Hopewell: culture of, 117, 118, 130; burial mounds of, 118, 119; figurines of, 120; Mesoamerican traits of, 119-124 Hopi: ceremonies of, 96; mythology of, 108; kachinas of, 108; marginal position of, 109 horse, extinct: in northwestern Mexico, 17 houses: in northwestern Mexico, 18-23 passim, 31-34 passim; in northeastern Mexico, 89; in Southwest, 104; in eastern United States, 125; in Central America, 216-217 Huaca Negra: excavations at, 268 Huaca Prieta: pottery at, 268; figurines at, 271; agriculture at, 299-300 Huastec: location of, 59 Huastecan pottery: in northeastern Mexico, 90 Huatabampo complex: status of, 36, 37, 100 101 Huchiti: location and language of, 52, 57 Hueco caves: beads from, 81 Hueco phase: traits of, 92 Huerigas Polychrome: in northwestern Mexico, 7 Huichol: culture of, 97, 98, 102 Humboldtiana montezuma: in northeastern Mexico, 62 Hunter: radiocarbon date from, 122-123 hunting: in northwestern Mexico, 7; in Lower California, 5 1 , 53, 56; in northeastern Mexico, 64, 89, 90; in eastern United States, 113-117; in western United States, 114; in Central America, 215, 217; in Ecuador, 247 hunting cults: in Southwest, 95 idols: in Lower California, 55; in Central America, 230; in West Indies, 235, 238, 241 Idria columnaris, use of: in Lower California, 40 (fig. 2 ) , 42 Ignacieño: language of, 52 iguanas, use of: in Central America, 220 incense burners: in El Salvador, 134, 138, 139, 143, 145, 147, 148; in Honduras, 161, 173; in

INDEX

Costa Rica, 197; in Ecuador, 256, 259; in Mesoamerica, 256, 259 iron, use of: in Lower California, 51 iron pyrites, use of: in Mesoamerica, 261; in Ecuador, 262 irrigation, canal: in Southwest, 95, 104; in Central America, 218 Irritilia: location of, 94 Isla Conchaqua: pottery at, 151, 153 Issaquena: absence of platform mounds at, 126 Ixtlan: burial at, 260 Jacome: location of, 7 n. 5 jade and jadeite, use of: in Costa Rica, 183-184, 189, 193, 197; in historic Central American cultures, 223; in West Indies, 237; in Mesoamerica, 261, 262, 305; in China, 278, 282 jaguar fat, use of: in Central America, 220 jaguars, representations of: in El Salvador, 145, 147; in Costa Rica, 189; in Panama, 197, 200; in Ecuador, 256, 259. See also cats Jaina: pottery at, 239; dental mutilation at, 262 Jano: location of, 7 n. 5, 11; revolt of, 11 Jaral: investigations at, 162 Jeddito Black-on-yellow: decoration of, 108 Jicaque: location of, 179; language of, 210, 212 jocotes, use of: in Central America, 220, 221 Jomon pottery: similarities to, in Mesoamerica, 123; in Ecuador, 246 Jonuta: dental mutilation at, 262 Jora complex: traits of, 65, 67, 68 (fig. 6 ) , 70 (fig. 8 ) , 72-76 passim, 78, 81-83, 86 (fig. 2 9 ) , 87, 93 Jora points: in northeastern Mexico, 64 (fig. 3 ) , 67 Jornada branch: in northwestern Mexico, 14; pottery of, in northeastern Mexico, 83 Jova: location of, 7 n. 5, 28; language of, 11 n. 11; origin of, 13 Juchipila culture: survivals of, 100 Jumano: language of, 11 n. 11 jus primae noctis: in Central America, 225 kachina cult: origin of, 108, 109 Kamia: location of, 52 Kaminaljuyu: external relationships of, 136, 138, 139, 141, 249; dental mutilation at, 262; mirrors at, 279, 280; ceremonial centers at, 306, 309 kaolin, use of: in northwestern Mexico, 7 Karatus plumieri, use of: in Central America, 218 Khmer: voyages of, 284; planned architectural assemblages of, 310 Kickapoo: in northern Mexico, 9 Kiliwa: location of, 52, 53; culture of, 54, 55 Kinosternum sonorense, use of: in northeastern Mexico, 82 (fig. 25) kivas: in northwestern Mexico, 18, 2 1 ; in Southwest, 95, 96, 108, 109 Knight Village: maize at, 118, 119 knives, stone: in northeastern Mexico, 83, 9 1 ; in Texas, 9 1 ; in Mesoamerica, 115, 128; in eastern United States, 117, 118; in West Indies, 238 knots: in northeastern Mexico, 77

Kolomoki: radiocarbon dates at, 126 Kuaua: murals at, 108 Kwataha: identification of, 108 La La La La La La La La La La La La

Ceiba: investigations at, 162, 167 Concepcion: investigations at, 197 Cueva Pintada: pictographs at, 36 Jolla complex: in Lower California, 43, 44 Morita: excavations at, 13 Ola: pottery at, 148, 177 Paz points: in Lower California, 46, 49 Peña: burials at, 198 Perra phase: traits of, 87, 89, 90, 299 Piedra: dental mutilation at, 262 Playa: excavations at, 29, 30, 33, 34 Quemada: pottery at, 101-102; ball courts at, 103 La Quemada culture: See Malpaso-Canutillo La Rama: pottery at, 151 La Roma: burials at, 260 La Salta phase: pottery of, 127 La Tolita: pottery at, 256; dental mutilation at, 261 La Venta: stone work at, 120, 121, 139, 309; cranial deformation at, 121; pottery at, 121, 122, 248, 269, 271, 273; figurines at, 270; temple center at, 308-309, 310 La Venta Coarse Brown: decoration of, 122 La Venta Coarse Buff: decoration of, 122 La Victoria: external relationships of, 247-248 labrets: in Central America, 223 Lagenaria, use of: in Lower California, 5 1 ; in eastern United States, 119; in Peru, 267 Lagoa Santa: crania of, 86 Laguna phase: traits of, 87, 124, 302 Lagunero: location of, 94 Lake Borgne Incised: decoration of, 122 Lake Chapala: sites near, 44, 45 Lampsilis siliquoidea, use of: in northeastern Mexico, 70 (fig. 8) lances, use of: in Central America, 217, 224, 229. See also darts, spears Langtry points: in northeastern Mexico, 90, 91 Las Aldas. See Las Haldas Las Charcas: traits represented at, 136, 306, 310 Las Flores: materials from, 162, 167, 169, 171, 173, 174 Las Haldas: radiocarbon dates from, 268; figurines at, 270; as cult center, 303, 310 Las Huacas: excavations at, 189 Las Joyas phase: traits of, 100, 103, 105 Las Palmas culture: traits of, 47-50, 5 1 , 56, 57 Las Tolas de Huaraqui: burials at, 259-260 Las Trincheras: excavations at, 3 n. 3, 28, 32, 33 Las Vegas ( H o n d u r a s ) : investigations at, 175 Las Vegas Polychrome: in El Salvador, 147, 151 Late Chou: traits of, 278-280 passim, 282, 284, 313 Late Polychrome: in Costa Rica, 189 Late Woodland: characteristics of, 126 Latter-day Saints: in northwestern Mexico, 9 Laurentian: crania of, 117 Laymon: culture of, 50; language of, 52

357

INDEX

Lemaireocereus gríseas, use of: in Central America, 219 Lenca: in El Salvador, 134, 152, 153; in Honduras, 179; language of, 210, 212, 213, 214; possible sites of, 215 η. 28; houses and furniture of, 216, 217; weapons and warfare of, 217, 230; textiles of, 224; trade of, 229; social and political organization of, 229 n. 55; religion of, 230, 231; calendar of, 233 Lepus californicus, use of: in northeastern Mexico, 70 (fig. 8) Lerma phase: traits of, 87, 89, 90 Lerma points: in northeastern Mexico, 67, 245; in Mesoamerica, 245; in Ecuador, 245 Lewisville: age of, 113 ligatures, ornamental: in Central America, 225 limestone, use of: in northeastern Mexico, 68-71 passim, 82 Lincoln Black-on-red: in northwestern Mexico, 20 linguistic families: of Lower California, 51-52; of Central America, 210-214 Lipan Apache: in northeastern Mexico, 87 Lipan phase: in northern Mexico, 14 litters, use of: in Central America, 223-224 Little Colorado polychromes: in northwestern Mexico, 21 Llano complex: in northwestern Mexico, 16, 29, 36; in Lower California, 44 loincloths: in northeastern Mexico, 85; in Central America, 221, 222 Loma San Gabriel culture: distribution of, 99, 100, 101 loops, wooden: in northeastern Mexico, 72 (fig. 11), 74 (fig. 13), 82 Loreto blades: in Lower California, 46, 49 Los Calpules: pottery at, 177 Los Chiqueros: burials at, 260 Los Frailes: excavations at, 49 Los Higos: investigations at, 159 Los Llanitos: materials from, 147, 148, 151, 152, 171 Los Llanitos Polychrome: in El Salvador, 147, 151, 171 Los Naranjos: traits represented at, 162, 167, 174 Loson: artifacts from, 245 lotus motif: in Asia and Mesoamerica, 284-287, 288-289, 295, 297 Lower Austral zone: in northeastern Mexico, 60 Lower Colorado Buff ware: in northwestern Mexico, 30 Lower Tres Zapotes phase: pottery of, 249, 253 Luna ware: in Nicaragua, 182; in Costa Rica, 189, 191 macanas, use of: in Central America, 229 macaws, representations of: in northwestern Mexico, 23. See abo parrots Machalilla: traits of, 246-247 Macro-Chibchan languages: in Central America, 210 Macro-Nawan languages: in northwestern Mexico, 11 n. 11 Macro-Otomanguean languages: in Central America, 210

358

Madden Lake: projectile points at, 201 maguey. See agaves maize: in northwestern Mexico, 14, 16, 17, 19; in Lower California, 5 1 ; in northeastern Mexico, 89, 90, 116, 303; in Southwest, 95, 102, 119, 299; in eastern United States, 118, 119, 124, 125; in Central America, 218-221 passim, 226, 229, 230; in West Indies, 239; in Peru and Mesoamerica, 266, 267-268, 303; in Asia, 299, 300 majagua, use of: in Central America, 218 majolica: in Honduras, 161 makara: in Asia and Mesoamerica, 287-288 Malpaso-Canutillo culture: status of, 100; pottery of, 102 mameys, use of: in Central America, 220 mammoths: association of, with man, in Mesoamerica, 113-114, 115 Mamom phase: pottery of, 249 Managua: settlement pattern at, 215 n. 29 manati, hunting of: in Central America, 218 Mandeville: figurine at, 120; pottery at, 124 mandibles, animal: use of, in northeastern Mexico, 78, 81 (fig. 2 4 ) , 82 (fig. 25) Mangue: language of, 213, 214 Mangus Black-on-white: in northwestern Mexico, 19 Manihot utilissima, use of: in Central America, 218 manioc. See yuca manos: in northwestern Mexico, 29—31, 35—36, 107; in Lower California, 50, 54; in northeaste m Mexico, 68, 82; in Mesoamerica, 115; in El Salvador, 148; in Honduras, 174; in Panama, 197 Marajoara: pottery of, 312, 313 marble, use of: in Honduras, 173, 277-283 passim Marcos de Colon: fortifications at, 215 n. 28 Maribio: language of, 210, 213; political organization of, 228; religion of, 230-231 marine worm tubes, use of: in northeastern Mexico, 70 (fig. 8 ) , 8 1 , 93 markets: in Central America, 215, 216, 220, 229 Marksville: pottery of, 121, 123, 124 maroma. See regurgitation Marquesas: Shang style in, 312 masked dancers: in Southwest, 95, 108, 109 masks: in Lower California, 5 1 , 54; in El Salvador, 139; in Honduras, 162; in historic Central American cultures, 226; in Ecuador, 256, 259; in Mesoamerica, 256-259, 279, 285; in Asia, 285, 287 masonry, core: in Southwest, 106 masonry, stone: in northwestern Mexico, 7, 14, 32; in Southwest, 97, 103 mastodon: late survival of, 113 Matagalpa: in El Salvador, 134; language of, 210, 212, 214 Matapalo phase: pottery of, 254-255 Matina: language of, 211 matting: in Lower California, 48 (fig. 9 ) ; in northeastern Mexico, 74-75, 84-86 passim, 115; in western United States, 114; in Central America, 217, 218

INDEX

mauls, grooved: in northwestern Mexico, 103 Mayan languages: in Central America, 210 Mayo: location of, 28; pottery of, 31 Mayran complex: in northeastern Mexico, 70 (fig. 8 ) , 72 (fig. 11), 75-78 passim, 83-87 passim, 89, 90, 92, 93 Mazapan period: pottery of, 143 Mechora: language of, 212 Medaños phase: traits of, 13, 14, 18 medicine stones: in Southwest and northwestern Mexico, 103 Medio period: traits of, 20, 22, 23 Mesa de Guaje phase: food remains of, 303 mescal bean, use of: in northeastern Mexico, 81 Metate Cave: excavations at, 50 metates: in northwestern Mexico, 2 9 - 3 1 , 35-36, 103; in southern California, 44; in Lower California, 48 (figs. 8, 9 ) , 50, 54; in northeastern Mexico, 68, 82, 83, 9 1 ; in Southwest, 103; in eastern United States, 127; in El Salvador, 135, 136, 148, 153; in Honduras, 174, 175; in Costa Rica, 183, 193, 197, 237; in Panama, 198; in West Indies, 237 meteorites, use of: in northwestern Mexico, 11 n. 13 mica, use of: in Honduras, 174 Michilia Red-filled Engraved: in northwestern Mexico, 103 microblades: in eastern United States, 120; in Ecuador, 245 midden circles: in northeastern Mexico and Texas, 92 Middle Chou: traits of, 278, 280, 313 Middle Polychrome: in Costa Rica, 184, 189 Mier focus: in northeastern Mexico and Texas, 92 migrations: in Southwest, 96, 98-99, 106, 107; in eastern United States, 125; in El Salvador, 153; in Costa Rica and Nicaragua, 180-181; in Peru, 272 milling stones: in northwestern Mexico, 2 9 - 3 1 , 35-36, 103, 107; in southern California, 44, 45; in Lower California, 48 (figs. 8, 9 ) , 50, 53; in northeastern Mexico, 68, 82, 83, 9 1 ; in Southwest, 103; in Mesoamerica, 115; in eastern United States, 127; in El Salvador, 135, 136, 148, 153; in Honduras, 174, 175; in Costa Rica, 183, 193; in Panama, 198; in West Indies, 237 Milnesand points: distribution of, 115 Mímbrenos Apache: in northwestern Mexico, 11 Mimbres: influence of, in northwestern Mexico, 12, 14, 16, 20; copper bells of, 106 Mimbres Black-on-white: in northwestern Mexico, 20 Mimbres Polychrome: in northwestern Mexico, 20 Mirador phase: pottery of, 247-248 Miraflores phase: relationships of, in El Salvador, 136 mirrors, bronze: in China, 280 mirrors, mosaic: in Southwest, 95-97 passim, 103, 106, 108; in northwestern Mexico, 103; in El Salvador, 139, 141; in Costa Rica, 183; in Guatemala, 279, 280; in Mesoamerica, 280 Misquito: language of, 210, 213, 214; settlement patterns of, 215; houses of, 216; clothing of,

222; ornaments of, 223; canoes of, 224; social organization of, 229; string records of, 235 missionaries, Spanish: in Lower California, 46—47, 50, 56-57; in northeastern Mexico, 60, 84 Mississippian: maize of, 118, 119; ornaments of, 119; pottery of, 124; Mesoameriean traits of, 125-130 Misumalpan languages: in Central America, 210 Mitla: stone mosaics at, 275 Mixteca-Puebla style: in northwestern Mexico, 107; in eastern United States, 129 moats: in Central America, 215 Modoc Shelter: crania of, 117 Mogollon: influence of, in northwestern Mexico, 9, 13, 14, 16, 18, 20, 35-36; relationships of, to Anasazi and Hohokam, 104, 105, 109; Mesoamerican traits of, 98, 100-102 passim, 104-105, 108 Mogollon Village: beads from, 81 Mohenjo Daro: cotton at, 300 Momil: pottery of, 269, 274; figurines of, 271 Monagrillo culture: description of, 201-202, 2 6 6 267, 301; external relationships of, 245 money, copper-ax: in Ecuador and Mesoamerica, 263 monkeys, representations of: in El Salvador, 138; in Honduras, 167; in Panama, 200 monkeys, as food: in Central America, 220 Monqui-Didia: language of, 52 Monte Alban: pottery of, 259; dental mutilation at, 262; sculpture at, 273, 282; temple center at, 309; glyphs at, 312 Monte aspect: in Texas, 92 Monte Fresco phase: pottery of, 248-250 passim, 253 Montell points: in northeastern Mexico and Texas, 91 Monte Negro: dental mutilation at, 262 Morhiss focus: in Texas, 91 Morhiss points: in northeastern Mexico, 124 mortars: in Lower California, 55; in northeastern Mexico, 69, 82, 83; in Southwest, 103; in northwestern Mexico, 103; in Panama, 202 mosaics, carved: in Mesoamerica and Peru, 275 mosaics, turquoise: in Southwest, 97, 103, 106, 108; in northwestern Mexico, 103 mounds, ceremonial: in northwestern Mexico, 6 n. 17, 20 (fig. 8 ) , 18, 2 1 , 22, 33, 103, in northeastern Mexico, 89, 118; in Southwest, 95, 96, 99, 103, 106; in eastern United States, 125, 126; in El Salvador, 138, 139, 143, 147; in Honduras, 161, 162, 175, 177, 215 n. 28; in Costa Rica, 193; in Mesoamerica, 307, 310 mounds, burial: in northwestern Mexico, 36; in eastern United States, 117, 118, 119, 124, 307; in Panama, 202; in Mesoamerica, 307 mounds, house: in northeastern Mexico, 89; in Southwest, 107; in Honduras, 175, 177; in Costa Rica, 183; in historic Central American cultures, 217 Moundville: pottery at, 129 mountain lion, use of: in northeastern Mexico, 85 mourning anniversaries: in northeastern Mexico, 89

359

INDEX

multistoried structures: in northwestern Mexico, 22 (fig. 9 ) , 23 mural paintings: in Southwest, 95, 108-109, 129 musical instruments: in northwestern Mexico, 23; in Costa Rica, 193 Naco ( H o n d u r a s ) : investigations at, 161-162 Naco phase: traits of, 161, 175 Naco Polychrome: in Honduras, 161 naguals: in Central America, 231-232 Nahua languages: in Central America, 210, 211, 213-214 Nahuat languages: in Central America, 210 Nahuatl languages: in Central America, 210 Nakipa: location of, 52 Nal-Tel maize: in northeastern Mexico, 89 Nandaime: pottery at, 182 narcotics. See drugs Nata: language of, 210, 211 Natchez: religion of, 129 Natural Clay ware: in Panama, 202 necklaces, gold: in Central America, 222 needles: of agave spines, in northeastern Mexico, 63, 89 (fig. 2 3 ) ; of copper, in northwestern Mexico, 23, 103 negative-painted pottery: in northwestern Mexico, 102, 103, 105; in Southwest, 105; in eastern United States, 124, 125, 127; in Mesoamerica, 124, 125, 127; in Ecuador, 253-254; in Peru, 272 nets and netting: in Lower California, 48 (figs. 8, 9 ) , 49, 5 1 , 53-56 passim; in northeastern Mexico, 7 1 , 73, 75, 83, 85, 86, 90; in western United States, 114; in Central America, 217, 218 net sinkers: in Honduras, 161 new fire ceremonies: in Southwest, 95 Newt Kash Rock Shelter: gourds and sunflowers at, 119 Nicarao: language of, 210-214 passim; towns and houses of, 215-217 passim; economic life of, 217-221 passim; religion of, 220, 230, 231; social and political organization of, 221, 224, 225, 228; clothing and ornaments of, 222, 223; textiles of, 224; cranial deformation among, 225; burials of, 227; lawways of, 227, 228; trade of, 229; warfare of, 229, 230; medical practices of, 231; games of, 232; calendar of, 233 Nicoya Polychrome: in El Salvador, 135, 138, 141, 143, 151, 155; in Costa Rica, 187, 189 nísperos, use of: in Central America, 219 Nogales phase: traits of, 87, 89, 90, 302 Nogales points: in northeastern Mexico, 89 Nogales Polychrome: in northwestern Mexico, 36 Nolan points: in northeastern Mexico and Texas, 91 nomadism: in northwestern Mexico, 64, 65, 89, 90; in Central America, 215 Nopal Shelter: artifacts from, 63, 65 (fig. 3 ) , 66 (fig. 5 ) , 68 (fig. 6 ) , 69, 7 1 , 84 (fig. 26) nose ornaments: in Southwest, 108; in Panama, 207; in historic Central American cultures, 222, 223

360

oars. See paddles obsidian, use of: in northwestern Mexico, 3 1 ; in Lower California, 50; in Mesoamerica, 121, 304-305; in El Salvador, 135, 136, 151, 153; in West Indies, 237; in Ecuador, 245 obsidian dating: in Ecuador, 245 Ocampo phase: traits of, 89, 299 ocarinas: in Central America, 232 oceanic travel: in eastern Asia, 292-293, 309, 3 1 3 314 ochre, use of: in Lower California, 49; in Mesoamerica, 304 Ocos phase: pottery of, 121-123 passim, 247-248, 249, 253, 269, 271, 272, 301-302; ceremonial centers of, 306 Odocoileus hemionis, use of: in Lower California, 41 Oldwaw language. See Ulva Olivella, use of: in Lower California, 48 (figs. 8, 9) Olmec: ceremonial mounds of, 118; stone art of, 120, 121, 139, 310-311, 313; relationship of, to Hopewell, 120; stone celts of, 120; cranial deformation of, 121; pottery stamps of, 121; influences of, in Central America, 139, 184; pottery of, 248, 269, 273; masks of, 279-280; temple centers of, 309; contacts with Chinese of, 3 1 3 314 Ometepe Island: pottery of, 182 Ootam: culture history of, 99, 104 Opata: location of, 7 n. 5, 10, 11 n. 5, 28; language of, 11 n. 11; origin of, 13; mythology of, 23, 24 Opuntia, use of: in northeastern Mexico, 81 (fig. 24) ornament painting: in northeastern Mexico, 69 ornaments: in northwestern Mexico, 23, 3 1 , 33, 34, 103, 107; in Lower California, 48 (fig. 8 ) , 49-50; in northeastern Mexico, 69, 70 (fig. 8 ) , 81; in Southwest, 103, 108; in eastern United States, 116, 117, 119, 124; in El Salvador, 152; in Costa Rica, 184, 189, 197; in Panama, 198, 203, 205; in historic Central American cultures, 222-223, 226, 229, 322; in West Indies, 240; in Ecuador, 248 Otinapa Red-on-white: in northwestern Mexico, 105 Otomac: ball game of, 240 Otomanguean languages: of Central America, 210, 212 ovens, pit: in northwestern Mexico, 33 Ovis canadensis: in Lower California, 41 oysters, use of: in Lower California, 49-50 pachisi: relationship of, to patolli, 291, 295, 314 Pachycormus discolor: in Lower California, 43 pads, fiber: in northeastern Mexico, 80 (fig. 23) paddles and oars: in Lower California, 53; in Central America, 224; in West Indies, 236 pahos, effigy: in Southwest, 108 Paila Cave: culture of, 83 painting, body: in Lower California, 51 Paipai: location of, 52, 53; culture of, 55 Pajarito Plateau Biscuit ware: decoration of, 101

INDEX

Pala period: art of, 289 Palenque: dental mutilation at, 262; lotus motif at, 284, 288-289; representations of trees at, 289 Paleo-Indians, occupations by: in northwestern Mexico, 16, 29, 36; in Lower California, 44; in eastern United States, 113, 130; in Mesoamerica, 113-114, 115; in Central America, 180181 palettes, stone: in Southwest, 103 palisades: in Central America, 215 Palmar: Valdivia culture at, 245 Palmar ware: in Nicaragua, 182, 184 palmas: in Mesoamerica, 279, 282 Palmillas points: in northeastern Mexico, 124 palms, use of: in Lower California, 48 (fig. 8 ) , 49; in Central America, 216, 218, 220, 221, 229 Palo Blanco: pottery at, 177 Paloparado: pottery at, 34 Panelled Red ware: in Panama, 205 Panuco: pottery of, 122; projectile points of, 128 Papago: location of, 28; culture history of, 99 Paparo: language of, 210 papaya, use of: in Central America, 219, 220 Paquime: traits represented at, 10, 16, 2 1 , 23 Paquime phase: of northwestern Mexico, 23 Paraíso: sculpture at, 159 parrots, use of: in Southwest, 106, 108. See also macaws patolli: relationship of, to pachisi, 291, 295, 314 Pavon: pottery at, 249, 253 Paya: language of, 210, 214; sites attributed to, 215 n. 28; weapons and warfare of, 217, 230; salt sources of, 220-221; clothing of, 222; water transportation among, 224, 226; marriage patterns of, 225; religion of, 231 Pearl Islands: pottery of, 203 Pecos: crania of, 86 Pecos River focus: in Texas, 92 Pedernales points: in northeastern Mexico and Texas, 91 Pedro Carbo: burials at, 263 pegs, wooden: in northeastern Mexico, 73 pejibaye, use of: in Central America, 219, 221, 230 pendants, metal: in northwestern Mexico, 103; in Panama, 200 pendants, shell: in northwestern Mexico, 3 1 ; in northeastern Mexico, 70 (fig. 8) pendants, stone: in northeastern Mexico, 70 (fig. 8 ) ; in El Salvador and Costa Rica, 152; in Panama, 198; in historic Central American cultures, 223 Pendleton Ruin: excavations at, 13 penis covers: in Central America, 221-222, 229 Peninsular Yumans: languages of, 52; cultures of, 54-55, 57 peppers: in northeastern Mexico, 89, 116, 299; in Southwest, 95; in Central America, 219, 220. See also chili pepper Peralta complex: in northwestern Mexico, 29 Perdiz points: in northeastern Mexico, 84 (fig. 7) Pericú: location of, 44, 47, 52, 57; language of, 52; culture of, 53, 56; crania of, 86

Perros Bravos phase: traits of, 19, 20 pestles: in Southwest, 103; in northwestern Mexico, 103; in Mesoamerica, 115; in Costa Rica, 183; in Panama, 198; in West Indies, 239 petroglyphs: in Lower California, 50 (fig. 10), 51; in northeastern Mexico, 82, 87; in Costa Rica and Panama, 197; in West Indies, 240. See also pictographs peyote, use of: in northeastern Mexico, 84 pictographs: in northwestern Mexico, 36; in northeastern Mexico, 87. See abo petroglyphs Piedra Gorda: excavations at, 49 Piedras Negras: dental mutilation at, 262 pigeons, use of: in Central America, 220 pigs, wild: use of, in Central America, 220, 229 pillows, fiber: in northeastern Mexico, 80 (fig. 23) Pilon phase: traits of, 19 Pima: location of, 28; culture history of, 99 Pima-Papago maize: in Southwest and northwestern Mexico, 102 Pinctada mazatlanica, use of: in Lower California, 49 pineapples, use of: in Central America, 218, 221 pins, bone: in Lower California, 49 Pinto Basin complex: in Lower California, 31, 44-45, 49, 5 1 , 57 Pinto points: in Lower California, 31, 45, 49, 5 1 , 57 Pinto Polychrome: origin of, 107 pipes: in Lower California, 48 (fig. 9 ) , 5 1 , 54; in northeastern Mexico and Texas, 9 1 ; in northwestern Mexico, 107; in Southwest, 108, 129; in eastern United States, 119, 129; in Mesoamerica, 129 Pipil: in El Salvador, 134, 143, 145, 152 Pipil Polychrome: in El Salvador, 135 pitfalls, use of: in Central America, 217 Plainview points: distribution of, 115 Plano culture: definition of, 114, 115 plaques, stone: in El Salvador, 139 plaster, use of: in El Salvador, 138, 139, 141, 143; in Honduras, 161 Playa complex: See San Dieguito Playa de los Muertos: figurines at, 136, 163; excavations at, 162, 163; pottery at, 163-164, 165 (fig. 3 ) , 166 (fig. 4 ) Playa de los Muertos phase: traits of, 162, 248, 249, 250, 253 plazas: in northwestern Mexico, 16 n. 17, 20 (fig. 8 ) , 21-22; in northeastern Mexico, 89; in Southwest, 95-98 passim; in eastern United States, 125, 126; in El Salvador, 138, 143, 147; in Costa Rica, 183; in historic Central American cultures, 215; in Mesoamerica, 310 Pleistocene fauna: in northwestern Mexico, 17; in Lower California, 4 1 , 44, 57; in Mesoamerica, 113-114, 115; in eastern United States, 1 1 3 114 Plumbate ware: in El Salvador, 135, 138, 141, 143, 151, 155; in Honduras, 167, 174, 178; in Nicaragua, 182 pochteca: in Southwest, 99 Pocosi: language of, 211 Point of Pines: painted stone slabs at, 108

361

INDEX

Point Peninsula: pottery of, 301 poisoning, of fish: in Central America, 218 Pokomam: in El Salvador, 134 popcorn: in Mesoamerica and Peru, 303 porcelain: in Lower California, 51 Portales points: distribution of, 115 Posorja: Valdivia culture at, 245 Poton: language of, 213 n. 2 1 , 214; sites attributed to, 211 pottery: in northwestern Mexico, 3 n. 3, 4 n. 4, 7, 12-14 passim, 18-23 passim, 30, 3 1 , 33-37 passim, 99, 100-107 passim, 275; in Lower California, 47, 5 1 , 54; in southern California, 55; in northeastern Mexico, 63, 82-84 passim, 89, 90, 116, 124, 127, 303; in Southwest, 9 5 98 passim, 101, 104-108 passim, 127-128; in Mesoamerica, 101-102, 107, 121-124, 127-128, 143, 239; in eastern United States, 121-129 passim, 301; in El Salvador, 134-135, 170, 171; in Honduras, 157-159 passim, 248-250 passim, 253; in Nicaragua, 170, 182; in Costa Rica, 182, 184-191, 193-197 passim, 207, 224, 248, 250, 253-255 passim; in Panama, 194, 196, 197, 200-207 passim, 266, 301; in historic Central American cultures, 229, 232; in West Indies, 235, 237, 241; in Ecuador, 243, 245-259 passim, 266, 301; in Colombia, 266, 267, 269, 274, 301; in Peru, 266-274 passim; in Venezuela, 267, 274, 301; in Asia, 283-284, 301 Pottery Mound: murals at, 108-109 Poverty Point: mounds and earthworks at, 118; artifacts at, 118, 120 prayer sticks: in Southwest, 108 Preceramic horizon: in northwestern Mexico, 16; in West Indies, 234 prickly pears, use of: in northeastern Mexico, 81 (fig. 2 4 ) , 85 priests: in Southwest, 95; in eastern United States, 129; in Mesoamerica, 129, 305; in Central America, 224, 225, 231; in West Indies, 235 Progreso White: relationships of, 250 projectile points: in northwestern Mexico, 29, 3 1 , 36; in Lower California, 44, 45, 46, 48 (fig. 9 ) , 49, 53; in northeastern Mexico, 65-68 passim, 71, 82-84 passim, 89-93 passim, 115, 124, 125; in eastern United States, 113-114, 115, 125, 128; in western United States, 114; in Plains area, 115; in Mesoamerica, 124, 125, 128; in El Salvador, 151; in Costa Rica, 181; in Panama, 198, 201, 207; in Ecuador, 245 prostitution: in Central America, 227 Protula superba, use of: in northeastern Mexico, 81 Pucará: sculpture at, 273 Pueblo Bonito: Mesoamerican traits at, 105-106 Pueblo Indians ( m o d e r n ) : external relationships of, 14, 16; ceremonies of, 96, 97; Mesoamerican traits of, 97; mythology of, 109 Pueblo Nuevo: traits represented at, 197, 202, 203 Puerto Hormiga: pottery at, 266, 301 pumice, use of: in Mesoamerica and West Indies, 237 pumpkins: in northeastern Mexico, 116, 303. See also squashes

362

Punin: calvarium from, 245 Punta Arenas de Posorja: Valdivia culture at, 245 Punta Pescado: excavations at, 49 Purple Polychrome: in El Salvador, 138, 139, 141 pyramidal structures: in northwestern Mexico, 2 1 ; in Asia and Mesoamerica, 291, 310 pyro communication: in northwestern Mexico, 21 quails, use of: in Central America, 220 quarries, stone: in Lower California, 44 quartz crystals, use of: in northwestern Mexico, 7 Quelapa: pottery at, 152, 171; ball courts at, 175 Quemada: external influence of, 21 Quepo: language of, 210, 211; ceremonies of, 232 Quetzalcoatl: cult of, in Southwest, 106, 109; in Nicarao calendar, 233 quids, fiber: in northeastern Mexico, 73, 8 1 ; in western United States, 114 rabbits, use of: in Lower California, 5 1 , 54; in northeastern Mexico, 86; in Central America, 220 rabbit sticks: in Lower California, 54; in northeastern Mexico, 7 1 . See also clubs, grooved wooden racing, foot: in Lower California, 55 radiocarbon dates: in northeastern Mexico, 62, 63, 75, 89, 9 1 , 245, 299; in Mesoamerica, 113, 304, 309; in eastern United States, 113, 118, 119, 122, 126, 301; in Costa Rica, 184, 187, 193; in Panama, 197, 201, 202, 203; in Southwest, 299; in Venezuela, 234; in Peru, 268, 303; in Ecuador, 263, 301; in Columbia, 301 rafts, use of: in Central America, 224 rain ceremonies: in Southwest, 95, 104; in Mesoamerica, 104, 109 Rama: language of, 210, 212 Ramos phase: traits of, 14 Ramos Polychrome: in northwestern Mexico, 23 Rancho Peludo: pottery at, 301 Rapid Style Polychrome: in El Salvador, 138, 139, 141 rasps, musical: in northwestern Mexico, 23; in northeastern Mexico, 71 rattles: in Central America, 232; in Mesoamerica, 305 rattlesnakes, use of: in northeastern Mexico, 62 Red-line ware: in Panama, 205 Red Ochre complex: burials of, 119 Red Paint complex: burials of, 119 Red ware: in Panama, 202, 205 reeds. See canes reels, netting: in northeastern Mexico, 71 Refugio points: in northeastern Mexico, 67, 90 regurgitation, induced: in Lower California, 55 Renner: maize at, 118, 119 Repelo complex: traits of, 90 reservoirs, water: in northwestern Mexico, 16 n. 17 resin, use of: in Central America, 229 revolt: of Pueblo Indians, 21 Reyes phase: traits of, 21 Ridge Ruin: burials at, 108 rings: gold, in Central America, 222; stone, in West Indies, 238

INDEX

Rio de Sonora culture: definition of, 12 Rio Tunal phase: pottery of, 105; culture history of, 106 roadways: in Costa Rica, 193; in historic Central American cultures, 223-224 robes: in Lower California, 54; in northeastern Mexico, 75, 85; in western United States, 114; in Central America, 221, 222 rocker-stamping: distribution of, 121, 301; in Mesoamerica, 121-122, 125, 248; in Honduras, 163; in Ecuador, 245, 301; in Peru, 272 rock grooves: in northeastern Mexico, 69 rock middens: in northeastern Mexico, 69, 7 1 , 82, 83; 9 1 , 94 Rockport focus: arrowpoints of, 82 rock shelters, use of: in Lower California, 50; in northeastern Mexico, 63, 64; in western United States, 114. See also caves rocks, thermally fractured: in northwestern Mexico, 23; in northeastern Mexico, 69, 81, 9 1 , 92 Rosario Mission: excavations at, 47 rosettes, fiber: in northeastern Mexico, 75, 80

(fig. 23)

Sacaton phase: relationship of, to Trincheras culture, 34 Sacaton Red-on-buff: in northwestern Mexico, 30 sacbes: in Mesoamerica, 310 n. 13 sacrifice, human: in northwestern Mexico, 2 1 ; in Southwest, 95; in Central America, 229, 2 3 0 231 sails, use of: in Central America, 224; in West Indies, 236 Saint Johns Polychrome: in northwestern Mexico, 21 Salado culture: in northwestern Mexico, 14, 2 1 , 35, 36; culture history of, 106; pottery of, 107 salt, sources of: in northwestern Mexico, 4, 7; in Central America, 220-221, 229 Salt Red: in northwestern Mexico, 36 San Borjita: petroglyphs at, 50 (fig. 10), 51 San Cayetano: ornaments at, 81 San Dieguito complex: in Lower California, 4 3 44 San Jose: stone vessels at, 173; dental mutilation at, 262 San Lorenzo Red-on-brown: description of, 101 San Lorenzo Tenochtitlan: sculpture at, 311 San Marcos: pottery at, 171 San Pedro stage: in northwestern Mexico, 29-31 passim, 36 San Salvador: pottery at, 171; sculpture at, 288 Sanci: stupa at, 311 sandals: in Lower California, 48 (fig. 9 ) ; in northeastern Mexico, 62, 73-74, 85, 93; in western United States, 114; in Central America, 222 Sandia: culture of, 113 sandstone, use of: in northwestern Mexico, 7 Santa Ana ( D u r a n g o ) : ornaments at, 81 Santa Ana ( H o n d u r a s ) : artifacts from, 162, 167, 169, 171, 173, 174, 177 Santa Catalina Mission: excavations at, 46-47 Santa Cruz phase: relationships of, 34

Santa Cruz Polychrome: in northwestern Mexico, 35, 36 Santa Elena Pablo: burials at, 260 Santa Helena ware: in Nicaragua, 182 Santa Isabel Iztapan: mammoth and artifacts at, 245 Santa Maria Polychrome: in Panama, 201 Santa Marta Cave: culture sequence of, 115; pottery of, 248 Santa Rita: artifacts at, 162, 163 Santa Rita de Copan: sculpture at, 159, 167-169 passim, 171, 173-175 passim Santa Rosa ( E l Salvador): pottery at, 151 Santa Rosa-Swift Creek: pottery of, 123 Sarigua complex: in Panama, 203 scalp ceremonies: in Southwest, 95 scarification, body: in Central America, 223 Scarified ware: in Costa Rica and Panama, 196197, 202 scarifiers: in northeastern Mexico, 77, 80 (fig. 23) Schroeder: ball court at, 103 scraper-planes: in Lower California, 45; in western United States, 114; in Mesoamerica, 115 scrapers: in northwestern Mexico, 29; in Lower California, 44; in northeastern Mexico, 68 (fig. 6 ) , 82, 83, 93; in western United States, 114; in Panama, 201; in West Indies, 238; in Ecuador, 245 sculpture, stone: in El Salvador, 139, 145, 152; in Honduras, 159; in Nicaragua, 181-182, 237; in Costa Rica, 183, 193, 196, 197; in Panama, 196-198 passim; in West Indies, 237; in Peru, 273, 275; in Mesoamerica, 273, 275, 310-311 seals, use of: in Lower California, 54 seats: stone, in Costa Rica, 183; wooden, in historic Central American cultures, 217. See also stools secret societies: in Lower California, 55 seeds, use of: in Lower California, 56; in northeastern Mexico, 70 (fig. 8) selenite, use of: in northeastern Mexico, 69, 70 (fig. 8) Selin phase: traits of, 167, 170 Sells Plain: in northwestern Mexico, 30 Sensenti: investigations at, 177 Seri: location of, 28; figurines of, 31; doublebladed paddles of, 53 serological studies: in northeastern Mexico, 85 serpentine, use of: in Panama, 207 serpents, representations of: in northwestern Mexico, 2 1 , 23, 29, 104, 107, 108; in Southwest, 108, 109; in El Salvador, 145; in Honduras, 169, 171; in Peru, 273; in Mesoamerica, 279, 287, 289, 291; in Asia, 279, 287, 289, 291 settlement patterns: in northwestern Mexico, 7, 12-15, 18, 21-23, 32-35 passim; in Southwest, 95; in eastern United States, 117, 125; in El Salvador, 138, 143, 147; in Honduras, 175; in historic Central American cultures, 214-217 sexual perversion: in Central America, 227 shamanism: in Lower California, 5 1 , 54-55; in Central America, 231; in Mesoamerica, 305 Shang period: tombs of, 307 n. 11; drums of, 312 sharks, representations of: in Panama, 200

363

INDEX

shark teeth, use of: in Lower California, 49 shell middens: in northwestern Mexico, 30, 3 1 , 36; in Panama, 201; in Ecuador, 245 shellfish, as food: in Central America, 220; in Ecuador, 245 shells, trade in: in northwestern Mexico, 9, 30, 31, 34 shells, use of: in northeastern Mexico, 70 (fig. 8 ) , 78, 83, 89, 103, 107; in Southwest, 95, 103; in Honduras, 169; in historic Central American cultures, 218, 221, 223, 224, 229, 232; in West Indies, 241 shields: in Central America, 229, 230 ships, use of: in Asia, 292-293 shirts: in Central America, 222 Shumla points: in northeastern Mexico, 90 Sierra de Tamaulipas: culture sequences of, 8 7 90, 302-303 Sikyatki: dental mutilation at, 129 Sikyatki Polychrome: description of, 108 silk-grass, use of: in Central America, 218 silver, use of: in northwestern Mexico, 10 n. 8; in Central America, 224; in China, 280 Sinagua: culture history of, 106 Sinaloan biotic province: in northwestern Mexico, 26 sinkers, net: in northeastern Mexico and Texas, 91 sipapu: origin of, 98 Sitio Conte: pottery at, 201, 203, 205; burials at, 207 skirts: in Lower California, 53; in Central America, 221, 222 slavery: in Central America, 227, 229 Smoked ware: in Panama, 205 Snaketown: beads at, 8 1 ; platform mounds at, 104 socio-political groups: in Lower California, 54, 56; in northeastern Mexico, 91-92; in Southwest, 95; in historic Central American cultures, 2 2 7 229; in West Indies, 235, 238-239, 240 sockets, antler: in northeastern Mexico and Texas, 91 Socorro points: in northeastern Mexico, 66 (fig. 4) sodasalt, use of: in northwestern Mexico, 7 Sonoran biotic province: in northwestern Mexico, 26; in Lower California, 42-43 Sophora secundiflora, use of: in northeastern Mexico, 81 sotol, use of: in northeastern Mexico, 63, 93 sour sap, use of: in Central America, 220 Southeastern Ceremonial complex: traits of, 1 2 7 129 spades, manufacture of: in Central America, 218 spatulas: in Lower California, 48 (fig. 8) spears: in Lower California, 5 1 , 56; in northeastern Mexico, 63, 67, 7 1 ; in western United States, 114 spear-throwers: in Lower California, 48 (fig. 8 ) , 49, 56, 57; in northeastern Mexico, 68, 71; in western United States, 114; in Central America, 229 speech-scrolls: in Ecuador, 243 Spinden hypothesis: evaluation of, 266, 267 spindle whorls: in northwestern Mexico, 107; in

364

El Salvador, 148; in Honduras, 161; in Mesoamerica, 305 Spiro: knives at, 128 Spondias purpurea, use of: in Central America, 221 Spondylus princeps, use of: in northeastern Mexico, 70 (fig. 8 ) spoons, use of: in Central America, 220 squashes: in northeastern Mexico, 89, 90, 116, 299, 303; in Southwest, 95; in Central America, 220; in Peru, 266, 267. See also pumpkins stairways: in northwestern Mexico, 2 1 ; in Southwest, 106 stamps, pottery: in Southwest, 106; in Mesoamerica, 121; in El Salvador, 136; in Honduras, 163; in Costa Rica, 189; in West Indies, 241; in Ecuador, 243, 256, 259; in Peru, 271 stars, representations of: in northwestern Mexico, 104 Starr points: in northeastern Mexico, 91 statues. See figurines, idols staves, wooden: in Costa Rica, 193 stelae: in El Salvador, 145; in Honduras, 159 stone, use of: in northwestern Mexico, 6, 7, 14, 16, 23, 29-32 passim, 35-36, 103, 107, 120; in Lower California, 31, 44, 45-46, 48-51 passim, 53-55 passim, 57; in southern California, 44, 49; in northeastern Mexico, 63-71 passim, 78, 82-85 passim, 89-93 passim, 115, 124, 125, 245; in Texas, 9 1 , 92, 115, 128; in Southwest, 95-97 passim, 103, 106, 108, 115, 120; in Mesoamerica, 102, 115, 120, 121, 124, 128, 237, 240, 245, 261, 262, 273, 275, 280, 282, 289, 291, 3 0 4 305, 309-311 passim; in eastern United States, 113-121 passim, 124, 125, 127, 128, 130; in western United States, 114, 115; in El Salvador, 135, 136, 139, 141, 145, 148, 151-153 passim; in Honduras, 159, 161, 162, 173, 174, 175, 2 7 7 283 passim; in Nicaragua, 181-182, 237; in Costa Rica, 181, 183-184, 189, 193, 196-198 passim, 237; in Panama, 196-198 passim, 2 0 1 202, 205, 207; in historic Central American cultures, 213, 223, 229, 230; in West Indies, 2 3 7 241 passim; in Ecuador, 245; in Peru, 273, 275; in Polynesia, 277-278; in Asia, 277, 278, 282, 289, 311; in Guatemala, 279, 280 stools, wooden: in Costa Rica, 193; in historic Central American cultures, 220. See also seats strips, fiber, self-wrapped: in northeastern Mexico, 80 (fig. 23) stucco, use of: in Central America, 216 stupas: in India, 311 Subtiaba: language of, 210 Suerre: language of, 210-212 passim; houses of, 216; head rests of, 217 Sula Fine Orange: in Honduras, 174 Sulphur Spring stage: in northwestern Mexico, 29 Suma: location of, 7 n. 5; linguistic and cultural affiliations of, 11 Sumu: language of, 210, 213; houses of, 216; salt sources of, 221; clothing of, 222; scarification by, 223; canoes of, 224; physical deformation among, 225; weapons of, 229; religion of, 231; string records of, 233

INDEX sun, religious behavior associated with: in Southwest, 95, 104, 108; in northwestern Mexico, 104 sunflowers, use of: in eastern United States, 116, 119; in northeastern Mexico, 303 Supanecan languages: in Central America, 210 suttee burial: in Central America, 226 sweat houses: in southern California, 55 sweet potatoes, use of: in Central America, 219 swizzle sticks, use of: in Central America, 220, 221 sylviculture: in Central America, 215, 218, 219 tablets, stone: in northeastern Mexico and Texas, 91; in eastern United States, 121; in Mesoamerica, 121, 273; in China, 278 tablets, wooden: in lower California, 54; in Costa Rica, 193 Tacacho: language of, 210, 213, 228 Tacuba: dental mutilation at, 262 Taguaca: language of, 210, 212, 213, 214; houses of, 216; religion of, 220, 230, 231; weapons of, 229 Taino. See Arawak Tajin: murals at, 282 Tajin style: Asiatic features of, 278, 279, 282, 283, 313 Talamanca: language of, 211; use of palms by, 218; foods of, 220; social organization of, 228; religion of, 231 n. 60 Talanga: pottery at, 177 Tamaulipan province: in northeastern Mexico, 60 Tamaulipecan languages: in northeastern Mexico, 93-94 Tamaulipas: culture sequences of, 89-90 Tammany Pinched: decoration of, 122 tapirs. See dantas Tarahumar: location of, 9, 94 Tardio period: traits of, 20 (fig. 8 ) , 2 1 - 2 3 passim tattooing: in northeastern Mexico, 77; in Central America, 219, 223; in Mesoamerica, 279-280; in New Zealand, 279-280; in China, 279-280 Tauaxka language. See Taguaca Taulepa: language of, 213 n. 21 Taylor points: in northeastern Mexico, 68 (fig. 6 ) , 91; in Texas, 91 Tazumal: traits represented at, 138-139, 141, 143, 147, 151, 170 Tchefuncte: pottery of, 122, 124 Tchefuncte Incised: decoration of, 122 Tchefuncte Stamped: description of, 122 Tecoatega: association of, with Nicarao, 215 n. 29 teeth, mutilation of: See dental mutilation teeth, animal, use of: in Central America, 223 Tegucigalpa ware: in Honduras, 177 Tehuacan: pottery at, 301 Tejar phase: pottery of, 253-254 Tenampua: traits represented at, 151, 175, 177 Tenampua Polychrome: in Honduras, 175, 177 Tenayuca: projectile points at, 128 Teotihuacan: pottery at, 124, 282, 283-284; projectile points at, 128; external influences of, 141, 155; dental mutilation at, 262; murals at, 282 Tepeaca: dental mutilation at, 262 Tepecano: culture of, 98

Tepehuan: culture history of, 94, 99 Tepetitlan: projectile points at, 128 Teribi. See Térraba Térraba: language of, 210, 211 terraces, agricultural: in northwestern Mexico, 7, 33 terraces, architectural: in northwestern Mexico, 32, 33 Texas: relationships of, to northeastern Mexico, 59, 67, 69, 7 1 , 81-93 passim, 115, 116; culture sequences of, 91-92 textiles: in northwestern Mexico, 11 n. 13; in northeastern Mexico, 78, 83, 84, 303; in Southwest, 95; in Costa Rica, 184; in Panama, 207, 217; in historic Central American cultures, 220, 221, 224, 226-229 passim; in Peru, 267, 275 textile impressions, on pottery: in Mesoamerica, 122, 123, 161, 302; in Peru, 272 thatch, use of: in Central America, 216 Three Circle Red-on-white: influence of, in northwestern Mexico, 105 three-pointed stones: in West Indies, 240, 241 Three Rivers Red-on-terracotta: in northwestern Mexico, 20 Tiahuanaco: sculpture at, 273 Ticoman: pottery at, 101, 253 tigers. See cats, jaguars Tikal: temples at, 309 Tizon Brown: in Lower California, 47 Tlaloc cult: in Southwest, 108, 109 Tlaloc, representations of: in El Salvador, 135, 143, 145, 152 Tlamimilolpan: radiocarbon dates at, 883 Tlatilco: pottery at, 101, 121, 122, 161, 178, 248, 249, 253, 254, 271-272; figurines at, 246 (Table 1 ) , 305; radiocarbon dates at, 304 n. 9 tobacco, use of: in Central America, 231 Tohil Plumbate: in Honduras, 178, 189; in Nicaragua, 182; in Costa Rica, El Salvador, and Guatemala, 189 Toltecs: migrations of, 99, 106; pottery of, 259 Tonala: sculpture at, 273 tongs, fire: in northeastern Mexico, 71 Tonjagua: structures at, 215 n. 28 Tonto Polychrome: in northwestern Mexico, 23; origin of, 107 Topawa phase: traits of, 107 tortillas: introduction of, into Central America, 219-220 tortoise shell, use of: in Lower California, 54; in northeastern Mexico, 82 (fig. 2 5 ) ; in Central America, 218 Tortugas points: in northeastern Mexico, 85, 89 towers, stone: in northwestern Mexico, 6 (fig. 3 ) ; in Southwest, 106 Toyah points: in northeastern Mexico, 84 (fig. 27) trade: in northwestern Mexico, 9, 30, 31, 34; in northeastern Mexico, 90; in Southwest, 101; in eastern United States, 117; in El Salvador, 151; in Honduras, 161; in Costa Rica, 189, 193, 197; in Panama, 197; in historic Central American cultures, 229, 235-236; in West Indies, 236, 237; in Ecuador, 267; in Peru, 274; in Asia, 284, 292-293

365

INDEX

trails: in northwestern Mexico, 7, 10 n. 8, 10 n. 9, 21 Transition zone: in Mesoamerica, 59 transportation devices: in Central America, 2 2 3 224, 236; in West Indies, 236; in Asia, 292-293 traps, use of: in Central America, 217 Travesía: traits represented at, 162, 167 tree decoration: in Lower California, 55 tree houses: in Central America, 216 trees, representations of: in Asia and Mesoamerica, 289 Tres Zapotes: pottery at, 259; figurines at, 270 tribute, payment of: in Central America, 224, 228 trincheras: definition of, 33 Trincheras culture: traits of, 30-36 passim Trincheras Polychrome: distribution of, 33, 34 Trincheras Purple-on-red: distribution of, 30, 3 2 36 passim Trincheras Red: distribution of, 33, 34 trophy heads: in Central America, 231 Troyville: platform mounds at, 126 trumpets: in Southwest, 95, 97, 103, 106; in northwestern Mexico, 103; in Central America, 232 tubes, use of: in Lower California, 54, 55 Tucson Polychrome: in northwestern Mexico, 23 Tula: destruction of, 99 Tula-Mazapan: expansion of, 126 tumbaga, use of: in Panama, 198-201 tumplines, use of: in Lower California, 48 (fig. 9 ) , 51 turkeys, use of: in northwestern Mexico, 7; in Central America, 220 Turner: maize at, 118, 119 turquoise, use of: in Southwest, 97, 103, 106, 108; in northwestern Mexico, 103; in Honduras, 162; in Mesoamerica, 262, 305 turtles, methods of taking: in Central America, 218 turtles, representations of: in northwestern Mexico, 104; in Panama, 200 Turucaca: language of, 210, 211 Twaka. See Taguaca twin deities: in Southwest, 95; in northwestern Mexico, 104; in Mesoamerica, 109 Typha, use of: in northeastern Mexico, 86 Tzakol phase: pottery of, 273 Uaxactun: stone vessels at, 173; pottery at, 249; dental mutilation at, 262; temple center at, 309-310 Ulua Bichrome: in Honduras, 162, 163, 175 Ulua language. See Ulva Ulua Marble Vases: in Honduras, 173 Ulua Polychrome: distribution of, 134, 138, 139, 141, 148, 151, 152, 155, 157, 159, 161, 162, 164-174, 175, 177, 178 Ulva: language of, 210, 212, 213, 214; sites attributed to, 216; cranial deformation of, 225; religion of, 231 Upper Austral zone: in Mesoamerica, 60 Upper Sonoran zone: in Lower California, 42, 43 Upper Tres Zapote: pottery of, 259

366

Urban Civilization horizon: in northwestern Mexico, 2 1 - 2 3 Urinama: language of, 210, 211 Usulutan technique: in El Salvador, 135, 136, 139, 145, 147, 153, 163; in Nicaragua, 182 Utaztecan languages: in northeastern Mexico, 9 3 94; in northwestern Mexico, 99; in Central America, 210, 211 Uxmal: lotus motif at, 284, 285 Valdivia culture: radiocarbon dates of, 121, 245; traits of, 245-247, 266-269 passim, 271, 272, 301, 306 Valshni Red: in northwestern Mexico, 36 vaults: in Honduras, 159; in Asia, 294 Venado Beach: radiocarbon dates at, 197; pottery at, 201, 203, 205 Ventana Cave: excavations at, 43; crania of, 87 Vermetus: incorrect identification of, 70 (fig. 8 ) , 81 Viceyta: language of, 211 Viejo period: traits of, 18, 20 Village-farming Community horizon: traits of, 1 7 21 Villalba: investigations at, 197 Vinette I: pottery of, 122-123, 301 voladores: in Central America, 230 Voto: language of, 210, 211, 212; marriage patterns of, 225; political organization of, 228 Walker-Giltnore: radiocarbon date at, 119 walnuts, use of: in northeastern Mexico, 68 wands, wooden: in Southwest, 108 warfare: in northwestern Mexico, 23, 24; in Lower California, 52-53, 56-57; in northeastern Mexico, 64; in Central America, 223, 228-30 passim; in West Indies, 236 Wariho: location of, 28 Warrau: location of, 234 Washingtonia: in Lower California, 40 (fig. 2) Waterfall Cave: excavations at, 16 water resources: in Lower California, 40, 4 1 ; in northeastern Mexico, 64-65 wattle-and-daub, use of: in northeastern Mexico, 89 weapons: in historic Central American cultures, 229-230; in West Indies, 241; in Asia and Polynesia, 277-278. See also bow and arrow, lances, spears Weedin Island: radiocarbon dates from, 126 Weedin Island Punctate: decoration of, 122 whales, use of: in Panama, 207; in historic Central American cultures, 223 wheeled toys: in Asia and Mesoamerica, 289 whistles: in Lower California, 48 (fig. 9 ) , 5 1 ; in Honduras, 173-174; in Ecuador and Mesoamerica, 256, 305 White-line on Red: in Nicaragua, 182 White Monochrome: relationships of, 250 White Slip: relationships of, 250 wood, use of: in Lower California, 48-51 passim, 53-57 passim; in northeastern Mexico, 63, 67, 68, 71-74 passim, 82, 8 4 - 8 6 passim, 89, 9 1 ; in Southwest, 106, 108; in western United States,

INDEX

114; in eastern United States and Mesoamerica, 128; in Costa Rica, 193; in Panama, 201; in historic Central American cultures, 207, 2 1 6 221 passim, 223-224, 226, 229, 230, 235-236; in West Indies, 236, 240, 241; in Ecuador, 243, 259, 262; in Mesoamerica, 285, 287, 291; in Asia, 285, 289, 291. See also cane Woodland culture: pottery of, 116-117, 123, 124; in eastern United States, 116-119, 301-302; Mesoamerican traits in, 117-125; maize of, 1 1 8 119 Woolwa. See Ulva world quarters, concept of: in Southwest and Mesoamerica, 98, 109; in northwestern Mexico, 104 wrenches, shaft: in northeastern Mexico, 71 X Fine Orange: in Honduras, 174, 189; in Costa Rica, Guatemala, and El Salvador, 189 Xicaque. See Jicaque Xipe Totec, representations of: in Mesoamerica, 143; in El Salvador, 152 Xipe cult: in Central America, 231 Xochicalco: art style of, 282 xolotls, representations of: in northwestern Mexico, 104 Xoxo: dental mutilation at, 262

Yagul: dental mutilation at, 262; stone mosaics at, 275 yams: in Central America, 219 Yaqui: location of, 28 Yarumela: traits represented at, 136, 157, 165 yerba flecha: clinical effects of, 7 n. 6 Yohoa Monochrome: in Honduras, 162-163 yokes, stone: in Mesoamerica, 282 Yoxiha: dental mutilation at, 262 yuca, use of: in Central America, 218-222 passim; in West Indies, 235, 239 Yucca, use of: in northeastern Mexico, 73 Yuman languages: in Lower California, 51-52 Zacateco: language of, 94 Zacatenco: pottery at, 249, 250, 253, 269, 273, 301; agriculture at, 304; houses and burials at, 304; stone artifacts at, 304-305, 306 Zambo: language of, 213 Zape: relationships of, 9, 13 zemis: in West Indies, 235, 238, 241 ziggurat: concepts represented by, 310 Zomoto: language of, 212 Zoned Bichrome: in Costa Rica, 182, 184 Zumatlan: copper-ax money at, 263 Zuni: as Puebloan frontier, 109

367