JOURNAL STAFF EDITOR: William Balée, oe of Anthropology, Tulane University, New Orleans, LA 70118 wbalee@tulan EDITORIAL ASSISTANTS: Janna Rose and James Welch, Department of Anthropology, Tulane University, New Orleans, LA 70118 BOOK REVIEW EDITOR: Michael K. Steinberg, Department of Geography-Anthropology, University of Southern Maine, Gorham, ME 04038 (mstein@usm.maine.edu) SOCIETY OFFICERS PRESIDENT: Karen R. Adams, Crow Canyon Archaeological Center, Cortez PRESIDENT-ELECT: Jan Timbrook, Santa Barbara Museum of Natural History, Santa Barbara, CA su auanayeie REASURER: Virginia Popper, UCLA Institute of Archaeology, Box 951519 Fowler A-210, Los -1510 eles, CA 90095 CONFERENCE COORDINATOR: Mollie S. Toll, Museum of New Mexico, Office of Archaeological Studies, Box 2087, Santa Fe, NM 87504 BOARD OF TRUSTEES . Kat Anderson, University of California, Davis, CA Leslie Main Johnson, University of Alberta, Edmonton, Alberta, Canada Melinda A. Zeder, Smithsonian Institution, Washington, D.C. Past ngenernes Steven A. Weber, Amadeo M. Rea, Elizabeth S. Wing, Paul Minnis, Cecil Brown, Catherine S. er, Nancy J. Turner, and Deborah M. Pearsall. Permanent board member Steven D. Emslie. The aiios dunes president-elect, secretary /treasurer, and conference coordinator. DITORIAL BOARD Eugene N. Anderson, University of California, Riverside, CA: ethnobotany, China, Maya. Scott Atran, CNRS, Paris, FRANCE: ethnobiological classification, cognition, history of science, Maya. Brent Berlin, University of Georgia, Athens, GA: ethnobiological classification, medical thnoboimny, Maya. Robert A. Bye, Jr., rib Botanico, Universidad Nacional Aut6noma de México, México, D.F., MEXICO: ethno- Pte Mex H. Sorayya Carr, El pene. California: zooarchaeology. Nina Etkin, University of Hawaii, Honolulu, HI: medical ethnobotany, the Pacific. Gayle J. Fritz, Washington University, St. Louis, MO: paleoethnobotany. Terence E. Hays, Rhode Island College, Providence, RI: ethnobiology, Papua New Gui Chris Healey, Northern Territory University, Darwin, AUSTRALIA: ethnozoology, Ca and New Guinea. Eugene S. Hunn, University of Washington, Seattle, WA: ethnobiology, ethnoecology, Maya and Zapotec. Timothy pee Macdonald College of McGill University, Quebec, CANADA: chemical ecology, ethnobotany, East Harriet V. Sohniehs, McGill University, Quebec, CANADA: ethno/human nutrition, First Nations of Canada. David L. Lentz, New York Botanical Garden, Bronx, NY: paleoethnobotany, Mesoamerica, Central Asia. Brien A. — Center for Plant — Missouri Botanical Garden, St. Louis, MO: ethnoecology, plant ervation, oe rai gard Naomi Miller, University of sccitveckiy Philadelphia, PA: paleobotany, ethnobotany, Near Eastern Archaeology. Gary Nabhan, Arizona san Desert Museum, Tucson, AZ: ethnobiology, Sonoran desert cultures. Amadeo M. Rea, San Diego, CA: cultural ecology, zooarc iaeology, ethnotaxonomics. Elizabeth J. Reitz, University of Georgia, Athens, GA: zooarchaeolo Mollie S. Toll, Museum of New Mexico, Santa Fe, NM: prehistoric and historic ethnobiology. The Journal of real is dames semi-annually. Manuscripts for ete and book review sections should be sent to the appr s listed on the inside back cover of this is © Society of oo ISSN 0278-077 COVER ILLUSTRATION: Close up of cattail down in the mouth of a Salish mo rtuary fi Anthropology, A1980). Drawn by J. Rose after a photo by J. Ostapkowicz een Journal of Ethnobiology VOLUME 21, NUMBER 2 WINTER 2001 CONTENTS ETHNOBIOTICA V ETHNOBOTANY OF KU-NU-CHE: CHEROKEE HICKORY NUT SOUP Gayle J. Fritz et al. 1 PURSUING THE FRUITS OF KNOWLEDGE: COGNITIVE ETHNOBOTANY IN MISSOURI'S LITTLE DIXIE Justin M. Nolan 29 CAVES, URSIDS, AND ARTIFACTS: A NATURAL-TRAP HYPOTHESIS Steve Wolverton 55 THE USE OF CATTAIL (TYPHA LATIFOLIA L.) DOWN AS A SACRED SUBSTANCE BY THE INTERIOR AND COAST SALISH OF BRITISH COLUMBIA Joanna Ostapkowicz et al 77 CONTRIBUTIONS TO THE ETHNOBOTANY OF THE CUP’IT ESKIMO, NUNIVAK ISLAND, ALASKA Dennis Griffin 91 TAXONOMIC IDENTITY OF “HALLUCINOGENIC” HARVESTER ANT (POGONOMYRMEX CALIFORNICUS) CONFIRMED Kevin P. Groark 133 BOOK REVIEWS 53; 73,129 yc em wer \ Ny AN ETHNOBIOTICA It's time for me to sign off as your editor, having reached the point of being able to Say, a mission accomplished. And now I have the privilege of introducing your new editor, distinguished archaeologist and ethnobiologist, Naomi F. Miller of the University of Penn- sylvania Museum, who already has served the journal with high energy and competence in her capacity as an editorial board member. I am confident Naomi will make an excellent editor and that you will agree with me in that assessment upon receiving her debut issue (Summer 2002, volume 22, number 1). All article manuscripts should be now submitted directly to Dr. Miller; please consult the inside back cover of this issue for her contact and mailing information. I am pleased to welcome also Darron Collins of the World Wildlife Fund as the journal’s new book review editor. Book reviews should now be sent directly to him at the address given on the inside back cover of this issue. At this time, and on behalf of the editorial board, I would like to thank Michael Steinberg for his superb service during the last three years as the journal’s book review editor, and to wish him well in future endeavors. For their service in reviewing manuscripts over the course of the last three issues, and on behalf of the editorial staff, I gratefully acknowledge the following persons: Eugene Anderson, Cecil Brown, James Carpenter, Helen Sorayya Carr, Alejandro de Avila B., Lydia Nakashima Degarrod, Darna Dufour, Nina Etkin, Jill Forshee, Catherine Fowler, Gayle Fritz, Ted Gragson, Chris Healey, Robert Hill, Sheila Humphrey, Eugene Hunn, Timothy Johns, Allen Johnson, Leslie Main Johnson, Elaine Joyal, Harriet V. Kuhnlein, David Lentz, Dana Lepofsky, Andrew MacWilliam, Judith Maxwell, Will McClatchy, Brien Meilleur, Jay Miller, Naomi Miller, Daniel Moerman, Travis Pickering, the late Darrell Posey, Elizabeth J. Reitz, Mary Riley, Laura Rival, Ted R. Schultz, Les Sponsel, Mike Steinberg, Lena Struwe, Maria Cruz Torres, Nancy J. Turner, Gail Wagner, Steve Weber, and Lyndon Wester. Special thanks are due to Adeline Masquelier and Myriam Huet of Tulane Univer- sity for expert editorial advice on the French abstracts to articles. Finally, appreciation is due to my in-house editorial assistants, here in the Department of Anthropology at Tulane, Janna Rose and James Welch, for their dedicated efforts in helping bring this issue to fruition. May the Journal of Ethnobiology and the Society for Ethnobiology continue to prosper and to advance human understanding of relations among biota, cultures, and languages. Journal of Ethnobiology 21(2): 1-27 Winter 2001 ETHNOBOTANY OF KU-NU-CHE: CHEROKEE HICKORY NUT SOUP GAYLE J. FRITZ, Ph.D. Department es ee Washington University in St. Louis 1114, St. Louis, Missouri 63130-4899 VIRGINIA DRYWATER WHITEKILLER, M.S.W. Department of Social Work, Northeastern State University Tahlequah, Oklahoma 74464 JAMES W. McINTOSH, M.S.W., L.S.W. Health Department, Cherokee Nation, Tahlequah, Oklahoma 74464 ABSTRACT.—A traditional hickory nut soup called ku-nu-che is consumed by many Cherokee people in eastern Oklahoma. A limited number of producers go through a two-stage process of cracking and pounding the nuts—primarily Carya texana—into a mixture of nutmeat and nutshell fragments that they form into balls for distribution to other households. Before being served as soup, these balls are dissolved in boiling water, strained to remove the nutshell fragments, mixed with cooked rice or hominy, and sweetened or salted. We interviewed six makers of ku-nu-che balls and describe their tools, their methods, and their motives for engaging in this labor-intensive practice. We also surveyed other tribal members to ascertain what ku-nu-che means to Cherokee people today. This study docu- ments long-term persistence of an ancient Native American plant food and, in addition, has implications for the interpretation of plant remains from archaeo- logical middens. Key words: hickory nuts, ethnobotany, Cherokee Indians, Native American food plants, paleoethnobotany. RESUMEN.—Una sopa tradicional preparada de nogal americano llamada ku-nu- chee es consumida por mucha gente Cherokee de Oklahoma oriental. Un ntimero contado de personas usan un proceso de preparacién en dos etapas. Primero, quiebran la cascara y después muelen la nuez (especie principal Carya texana) para formar pelotas de una mezcla de fragmentos de cascara y nuez molida que dis- tribuyen a otras familias. Antes de ser usada para sopa, las pelotas se deshacen en agua hirviendo, se cuelan para separar los fragmentos de cascara, se mezcla con arroz 0 maiz cocido y se agrega sal o azucar. Se presenta informacion de seis entrevistas con personas que se dedican a la labor de preparar pelotas. Se des- criben sus herramientas, sus métodos y sus motivos para hacer esta actividad muy laboriosa. También sondeamos a otros miembros del tribu sobre el signifi- cado de ku-nu-chee para la gente Cherokee en la época actual. Este estudio do- cumenta la persistencia larga de una comida tradicional indigena norteamericana y tiene implicaciones para la interpretacién de restos botdnicos de basureros ar- queoldégicos. Z FRITZ et al. Vol2l No.2 RESUME.—La soupe de Carya spp. appelée ku-nu-che est consommée par de nombreux Cherokees dans la partie Est de l‘Oklahoma. Une petite partie des producteurs adopte un processus en deux étapes de craquage et écrasement des noix (surtout Carya texana) pour en faire un mélange de pulpe de noix et de coquille, ensuite moulé en boules qui sont distribuées a d’autres familles. Pour préparer la soupe, ces boules sont dissoutes dans l’eau bouillante, filtrées pour en enlever les morceaux de coquille, mélangées a du riz cuit, puis sucrées ou salées. Nous avons interviewé six producteurs de boules de ku-nu-che. Nous décrivons leurs méthodes, leurs outils, et les motifs pour lesquels ils se consacrent a cette tache intensive. Nous avons aussi interrogé d’autres membres de la tribu pour comprendre la signification du ku-nu-che pour les Cherokees aujourd’hui. Cette étude fournit des données sur la persistence a long-terme d’une plante nutriti- onelle ancienne des Indiens d’Amérique, et a aussi des implications en ce qui concerne l’interprétation des residus de plantes dans les fouilles archéologiques. INTRODUCTION Cherokee people moving into northeast Oklahoma in the 1820s and 1830s were probably relieved to find an abundance of hickory trees (Carya Nuttall spp.) (Juglandaceae). Hickories would have been a welcome sight because nuts of thick- shelled species were—and still are—the basic ingredient of a traditional soup-like dish known as ku-nu-che (or “ga-nu-ge” or “conutchie” or “kinugee,” among other variants). Hickory nuts had been a dietary staple in the Eastern Woodlands for thousands of years before the transition to American Indian agriculture, and the nuts remained a central ingredient in cuisines of indigenous farming societies before and after the arrival of Europeans. Ku-nu-che is still today prepared in the households of members of the Western Cherokee Nation, with its seat of govern- ment in Tahlequah, Oklahoma, and those living in the southern Appalachian Mountains, homeland of the Cherokees before most were forced west in the early nineteenth century, before and during the Trail of Tears in 1838-1839. This study began primarily as an attempt to observe modern hickory nut Processing in order to gain insights into the ways hickory nutshell entered the archaeological record. Archaeologists look to ethnographic and ethnohistoric de- scriptions of plant use in order to understand better how plant remains and ar- tifacts may have been deposited in archaeological sites; in other words, to gain taphonomic and contextual insights. Interest by archaeologists in traditional food- ways increased during the 1970s and 1980s in conjunction with ecological ap- Proaches to archaeology, accompanied by large-scale recovery of plant and animal remains through newly developed methods including flotation. Hickory nutshell is the most abundant type of food plant in many archaeobotanical assemblages, sometimes outweighing the ubiquitous wood charcoal. This is especially true for samples from the Archaic period (8000-1000 b.c.e.), but some Mississippian pe- riod (1000 c.e. to European Contact) sites are also dominated by thick hickory nutshell. Archaeologists have looked to historical and early ethnographic sources for descriptions of native nut processing techniques, but none have, to our knowl- edge, considered the living Cherokee m d and make ku-nu-che rere epane women who gather-hickory, nuts a eee ee eee hatin a Winter 2001 JOURNAL OF ETHNOBIOLOGY 3 The fact that this food is still made and consumed by many Cherokee people today says a great deal about persistence of native values and appreciation of long-standing traditions. Although it might seem as if this particular tradition is in danger of disappearing, we found evidence to indicate that, because so many Cherokees continue to appreciate ku-nu-che, the incentive exists to ensure its avail- ability in the foreseeable future. Our objectives are, therefore, both ethnobotanical and ethnoarchaeological: to document in as much detail as possible contemporary ku-nu-che making processes as practiced by Cherokees in and around Tahlequah; to discuss the meaning and significance of ku-nu-che in modern Cherokee society; and to explore the ecological and archaeological implications of the harvesting and processing of hickory nuts by Cherokee people today. KU-NU-CHE IN MODERN CHEROKEE COUNTRY, OKLAHOMA Most adults and teenagers, and even many children, who live or grew up as members of a Cherokee community in northeastern Oklahoma are aware of ku- nu-che. They may not eat it often, but it is available at gatherings such as holiday and birthday dinners, church socials, and family reunions. Ku-nu-che is usually distributed in the form of balls (Figure 1), which can be purchased directly from individuals who process the raw nuts. Ku-nu-che balls are also sold at tribal health clinics, community grocery stores, and Cherokee Nation governmental offices. As a friend of one of our consultants said during an interview in Tahlequah, ‘’’“When FIGURE 1.—Uncooked ku-nu-che ball on right; nutshell sifted from one cooked ku-nu-che ball in front; undissolved nutmeat sifted from cooked ku-nu-che ball on left; cooked hickory nut soup in jar at rear. Note: one ball mixed with water and hominy or rice fills two or three jars 4 FRITZ et al. Vol. 21, No. 2 you get away from here, nobody knows about ku-nu-che, but around here, every- body knows it.” Whitekiller (second author of this paper) recalls that ku-nu-che was made from scratch, starting with nuts collected from trees in and near their yards, by members of her grandparents’ generation living in the cluster of homes and gar- dens owned by the extended Drywater family on the outskirts of Tahlequah. Mc- Intosh (third author of this paper) remembers a jar of cooked ku-nu-che soup often being available in the refrigerator of his grandparents’ rural home in Mays County when he was growing up. Grandchildren and other family members were free to help themselves to cold or reheated ku-nu-che as a snack whenever they desired. Ku-nu-che was (and still is) commonly served at church dinners. Mem- bers of the congregation serve themselves from a large pot, usually ladling it into styrofoam cups using a gourd dipper or large spoon. One of Whitekiller’s brothers was such a frequent visitor to the ku-nu-che pot as a child that he was jokingly called ““ku-nu-che boy” by the other Cherokee children playing nearby. The rich- ness of this dish, however, causes most people to consume it in moderate amounts. Ku-nu-che balls are made by a limited number of Cherokees. Many others buy the balls, which tend to be about the size of softballs, for approximately $5.00 to $6.00 each. The price of a ku-nu-che ball a few decades ago was $2.00. People who actually gather, crack, and pound hickory nuts and make ku-nu-che balls for distribution to others have special tools, although individuals who make a few balls each year for use by the immediate family may use only common household tools such as hammers or mallets. Sellers of ku-nu-che balls have been known to advertise on local call-in radio swap-meet shows or in newspapers that specialize in non-retail, person-to-person sales. Most knowledge about availability is, how- ever, spread by word of mouth. After providing historical background information, we introduce six serious ku-nu-che makers and tell how, where, and why they produce and distribute balls. We follow them through the steps of gathering or acquiring the nuts from others, of cracking, sieving, pounding, and forming the balls, of distributing (usually selling) the balls, and of cooking them. We then discuss motives for making ku- nu-che today and assess attitudes of Cherokee teenagers that make us optimistic about the survival of this traditional food. Finally, we briefly explore the archae- ological implications of modern ku-nu-che making. ARCHAEOLOGICAL, HISTORICAL, AND NUTRITIONAL BACKGROUND The archaeological record attests to the importance of hickory nuts in subsis- tence strategies of native peoples in the Eastern Woodlands as far back as Late Paleoindian times, 8300 b.c.e. (Detwiler et al. 1998). Middens dating to the Middle and Late Archaic periods, 6000-1000 b.c.e., typically contain masses of charred hickory nutshell, indicating that hickory nutmeat was a staple food, possibly the single most important plant food for many Woodland foragers (Asch, Ford and Asch 1972; Gardner 1997; Lopinot 1982; Yarnell and Black 1985). Prodigious amounts of charred hickory nutshell in archaeological sites might exaggerate the dietary importance of this resource due to its mass, its density and subsequent durability, and to the likelihood that cracked pieces of nutshell were deliberately Winter 2001 JOURNAL OF ETHNOBIOLOGY 5 burned more often than the remains of other food plants that do not make useful fuel. Even so, the “nutritional superiority” (Gardner 1997:175) in terms of caloric content and protein complement of hickory nuts over other nuts reinforces the claim that hickories were a “first-line” food resource (Asch, Ford, and Asch 1972) for foragers in what is now the eastern United States. Gardner (1997) points out that only 340 g dry weight of hickory nutmeat is required to supply 2200 kcal intake, compared to 427 g of acorn and 604 g of maize. The fat content of hickory nuts is double that of acorns and approximately sixteen times that of maize, a fact that ‘“‘may have been of considerable nutritional importance to Eastern Wood- lands foragers’ (Gardner 1997:162). Hickory nuts are higher than either acorns or maize in eight out of the ten essential amino acids, falling only slightly lower than maize in leucine and slightly lower than acorns in lysine (Gardner 1997:164). An assemblage of human paleofeces from Salts and Mammoth Caves in Ken- tucky demonstrates that hickory nuts were frequently consumed during the mid- dle first millennium b.c.e., at a time and place where native seed gardening had been integrated into the economy of hunters and gatherers (Yarnell 1969). Munson (1986) and Gardner (1997), in fact, hypothesize that management of nut groves (girdling and clearing to favor highly productive trees) was ecologically conducive to local domestication of plants such as sumpweed (lua annua L.) and chenopod (Chenopodium berlandieri Moq.). Southeastern American Indians did not abandon nut harvesting even after intensification of maize (Zea mays L.) agriculture at 800- 1200 c.e. Early European explorers and entrepreneurs enjoyed hickory nut foods and oils (Battle 1922; Talalay et al. 1984), and described groves near Indian vil- lages where nut trees were managed in an orchard-like fashion (Hammett 1992). Several European observers described the pounding of nuts and rendering of milk-like emulsion and oil. William Bartram, who lived among the Creeks in Georgia at the close of the eighteenth century, wrote: I have seen above an (sic) hundred bushels of these nuts belonging to one family. They pound them to pieces, and cast them into boiling water, which, after passing through fine strainers, preserves the most oily part of the liquid: this they call by a name which signifies Hiccory milk; it is as sweet and rich as fresh cream, and is an ingredient in most of their cookery, especially homony and corn cakes (Harper 1958:25). In 1799, Benjamin Hawkins made the additional observations that Creek hickory nut processors pounded the nuts in a mortar and winnowed the pieces ‘’to free the kernels as much as possible from the shells” (cited in Talalay et al. 1984:352). Hawkins also distinguished between hickory nut oil, which was separated when it rose to the top after water was added to the winnowed, pounded nuts, and “the milk,” which remained below and was not separated. John Lawson's observation of hickory nut use by unspecified Carolina Pied- mont natives in 1701 is unusual in that it describes consumption of solid nutmeat fragments rather than milk or oil. Lawson (1709:98) observed nuts broken “very small betwixt two stones till the Shells and Kernels are indifferent small; And this Powder you are presented withal in their Cabins, in little wooden Dishes; the Kernel dissolves in your Mouth, and the shell is spit out.” Lawson (1709:98- 99) described another dish, however, ‘‘the Soup which they make of these Nuts, 6 FRITZ et al. Vol. 21, No. 2 beaten, and put into Venison-Broth, which dissolves the Nut and thickens, whilst the Shell Precipitates, and remains at the bottom. This Broth tastes very rich.” Use of finely pounded hickory nutmeat to flavor and thicken soups and gruels obviously persisted among the eastern Cherokees. Writing about early nineteenth century Cherokee diet in the southern Appalachian region, Malone (1956:132) mentioned that a “tasty and frequent dish was ca-nu-chi (or car-nut-chee), con- sisting of corn meal mush mixed with crushed hickory nuts.” It is not clear if Malone was applying a term he knew only from the twentieth century to early nineteenth century hickory nut soup or if he had evidence for much earlier use of the food’s name. We assume it is a very old Cherokee word. [t appears in a 70-page manuscript written in English by a Cherokee woman, Wahnenauhi, sent by her from Oklahoma in 1889 to the Bureau of American Ethnology (Keys 1966). Wahnenauhi, whose English name was Lucy Lowrey Hoyt (Mrs. Lucy L. Keys after her marriage) graduated in 1855 from the Cherokee Female Seminary in Tahlequah and recorded valuable historical and cultural information in this man- uscript. Major John Wesley Powell, Director of the B.A.E., wrote to her, however, “You will ... understand that its value to the Bureau is comparatively small,” attempting to justify a purchase price of $10 (Kilpatrick 1966:182). In the manu- script, Keys (1966:194) tells about a band of eastern Cherokees who migrated as far west as the Rocky Mountains in the early eighteenth century (before 1730) to get away from White settlers: Although the greater part of the Tribe was very unwilling to have them leave, yet, finding their efforts to persuade them to remain, were unsuc- cessful, they assisted them in making preparations for the journey: some furnished “pack ponies,” while others loaded them with ‘“Cuh-whe-si, tah’”’ [hominy], ‘“Cuh-nuh-tsi,” dried venison, and other things. ... In an editorial footnote, Kilpatrick (in Keys 1966:194) describes cuh-nuh-tsi as a soup made of hominy and crushed hickory nuts and says the Cherokee people consider it to be their national dish. Myra Perry (1974), whose M.A. thesis focuses on wild plant foods used by Cherokees living on or near the Quallah Reservation in North Carolina, recorded a description of “ko-nu-chie” processing during her independent fieldwork, quot- ing Lish Sneed, a Cherokee elder, as specifying that hickory nuts are pounded between two stones, but the shells and meats not separated by hand because “you know that you can’t shell a hickory nut.” According to instructions provided to Perry by Geneva Welch, another elder, the fine, greasy meal is formed into a ball about two inches in diameter and dissolved into a quart of boiling water: ‘As it melts, you have soup. You would describe it more or less as a beverage. Sweet- ening with sugar is optional’ (Perry 1974:40). This recipe calls for the formation of balls, which were not mentioned in earlier accounts, but it does not break nut iC — sig oo of cracking and pounding that we found to be the A slightly earlier recipe collected from the same region, ho i the two-stage Process. In the book Cherokee Cooklore (Ulmer aa scare tailed instructions for making hickory nut soup (“ga-nu-ge’’) are shared b A ie Lossiah, granddaughter of the nineteenth century Cherokee chief John vies rris Winter 2001 JOURNAL OF ETHNOBIOLOGY 7 recipe, collected in 1949-1950, is quoted here in full here because it corresponds closely to the way ku-nu-che is made today in eastern Oklahoma, yet retains traditional aspects that we know about only as memories: Gather hickory nuts or scalybarks, dry on a rack before the fire. When the nuts are dry crack them by using a large flat rock placed in a flat basket lined temporarily with a cloth, use a smaller rock to pound the nuts when placed on the larger rock. When the nuts are all cracked sieve them through a sieve basket. Place the kernels and small hulls that passed through the sieve in the corn beater and pound until the substance can be made into balls. Roll this into balls until ready for use. These balls will keep fresh for several days if the weather is not too warm. When ready for Hickory Nut Soup place a ball or more in a vessel that will hold water, pour boiling water over the balls while stirring con- stantly. If this is made into a thick soup it may be served with any type bread or dumpling. If it is made into a thin soup it may be used as a drink. As soon as enough soup has been poured off to leave a very thick mixture more water may be added. Do not drink the very last of the mixture because that is where the little bits of hulls are (Ulmer and Beck 1951:48). CHEROKEE KU-NU-CHE BALL MAKERS Six experienced producers of ku-nu-che balls generously shared their methods and motives with us during the course of this study (1996-1999). Their tools and techniques might not be representative of all Cherokees who engage in the crack- ing and pounding of hickory nuts today. These individuals all speak English in addition to the Cherokee language and all live in easily accessible locations, facts that might distinguish them from non-English speakers in more remote, rural areas, although only two of their households had telephones in 1999. Nevertheless, our consultants were all raised in families where Cherokee was spoken and where traditional Cherokee values were taught. Two are retired men, Blue Rock and Daniel Beaver, both of whom lived until recently in Tahlequah (pop. 10,400) and made ku-nu-che balls in their homes. Sadly, Blue Rock passed away in October, 1999. Daniel Beaver moved to a smaller town in northeastern Oklahoma at approximately the same time. Narcy Holcomb is a homemaker whose children are teenagers and young adults. Mrs. Holcomb lives in a rural community a few kilometers southeast of Tahlequah and makes ku-nu-che balls in and near a shed behind her house. Ramona and Charley Carey are a semi-retired couple with grown children. They live in a rural community 20 km west of Tahlequah and move between a shed behind their house and their kitchen when making ku-nu-che balls. Patrick Bearpaw is a 21-year old college student and musician who, when not at school in Muskogee, Oklahoma, lives a few kilometers east of Jay, a town of 2,220 souls located 70 km north of Tahlequah. He makes ku-nu-che balls on the porch of his parents’ home. We observed only Blue Rock and Narcy Holcomb in the actual process of cracking and pounding nuts. Daniel Beaver has recently retired from making ku- 8 FRITZ et al. Vol. 21, No. 2 nu-che, but allowed us to examine his tools. The Careys and Patrick Bearpaw were interviewed during the summer, when balls are rarely made, but they, too, demonstrated their tools and described their production methods. GATHERING THE NUTS Modern makers of ku-nu-che balls either gather nuts themselves from acces- sible trees they know to be good producers or barter bags of nuts collected by people who furnish them to primary producers in exchange for a few balls. Sev- eral of our consultants pursue both strategies, remaining flexible from year to year. Ramona and Charley Carey rely primarily on nuts gathered themselves from the property of non-Indian neighbors who grant permission without any interest in using the nuts themselves or in receiving ku-nu-che balls. We spoke to nobody who goes into heavily wooded areas to collect hickory nuts, even though hickory trees comprise one of the dominant genera of the oak- hickory forests of northeastern Oklahoma. Most if not all gathering takes place in anthropogenically-opened locations: yards, parks, savanna-like pastures and hay- fields, and fence rows. Patrick Bearpaw, for example, frequently gathers nuts from the grounds of his church—Pineridge Baptist Church—on the outskirts of Jay. Clients who bring him bags of nuts usually gather them in their yards. We spec- ulate (but have no firm evidence) that many of the hickory trees left standing on Cherokee-owned property have been recognized as valuable sources of nuts for ku-nu-che, like native pecan trees left uncut across the Southeastern United States. Ongoing selective management whereby the heaviest nut producers (the . FIGURE 2.—Blue Rock in the process of crackin room of his home in Tahlequah, Oklahoma. & nuts inside a cardboard box in the bed- Winter 2001 JOURNAL OF ETHNOBIOLOGY 9 id FIGURE 3.—Narcy Holcomb’s Stage 1 cracking tools. “thrifty” trees) are favored by clearing away competitors for sun and root space probably differs little from pre-Contact management practices in the homeland of the Cherokees and other Southeastern tribes. Although hickory trees growing in closed-canopy forests produce fewer nuts that are much harder to gather given the undergrowth and unchecked competition from squirrels (Talalay et al. 1984), some are likely to have been gathered on occasion, especially during hard times. Wilma Mankiller, former Principal Chief of the Cherokee Nation, for example, includes hickory nuts as one type of wild plant food gathered by her large family when she was a child, along with walnuts, wild onions, dandelions, poke, mush- rooms, berries, and wild grapes (Mankiller and Wallis 1993:34). She does not specify, however, that the nuts were gathered in the woods. All source trees shown to us belong to the species Carya texana Buckl., by far the most common upland hickory in northeastern Oklahoma. Although tree books, including Trees of Arkansas (Moore 1986), refer to this species as the Blac Volb21.Nor2 FRITZ et al. asn [euos -Jod JO} dARY 0} puke auooul yuaweddns oj, spied awioy Jy} UT SUIMOIS SaaJ} WIOIY Ajysouu ‘syuelp> pue spuary Aq paystumz jod yoj}s uod jsed UT (iJequiog) yeq YJIM UPd daJJOD ase] aseq [JOU IP] -n3urj}da1 uo joTTeur ejow apeul-wojsn> yenboyey, ul awoY JO WOOIpag (6661 “90 Aeme passed) szapje paryey ajdoad dayYOIIYD Jayjo pue ‘Ajurey ‘Spudatsy 0} Ayrpiqeyreae oinsua oy, aye -O] payedsun wos Jjesisy pasrayyes duos ‘s}UaT]D pue spuaty Aq paystuiny ued ul o81e] ut apysad UuapooM apeUul-woysn> japue -[09 UaYD}Fy prepurysg aseq [e}our punos uo oo} [eJou apeu-woysn> yenbaryey eeu vase [eins ut au0Yy jo pred pue pays JoyeuUaUoYy dAyoVy Asuow 3utr -puads urea 0} pure S}UST]D PO SJoyye} 0} ayo-nu-ny Ayddns of, syudr]o fa poeystuiny Jjesumry pasoyyes autos ued daJJOO aBie] Ul ap}sod uapoom apeul-puepy] sajoy payound ay yprym ysnory} [Mog suse Jaya) UT A}LARIUOD YIM YOI yey UO aysad uapoony, Avf reou vase [ein ul asnoy jo ypiod yuo1y yuapnys a8atjoo pyo-1eaA-]Z SaAjasuiey} pue ‘Ayrumey ‘spuaty 0} Ayyiqeyreae ainsua Of, saimjsed pue spoom uedo ut pur] sioqysreu uerp ay Ivgel) apeuwl-wojsny ysnoryy peyound safoy [euonIppe YJIM Japuejod [eV [eiseped uo peop -utjAD uo a8pam uo] -pIPy pur “pars “pays SIOp]e Peiyal-TUTsag spueLy 0} Ajddns 0} pue awioout yuautayddns oy sjuar> Aq paystuiny awios ‘syed [eo] ul jJjesurry pasayjzes aurog as -UI [RJOUT YIM xoq yy81am ureyind [ee PaArasgo JON qrey|s aya1NUOD UO FOUTURPY yenb -a]ye], ul yuauyedy JOppe paige} SOAQ -oul AIRUILIg :synu Aso -yory JO saam0g (Sut -punod .10}) S]OO} Z a8e}S :syudU -ojdut Sunsis (Sut -yoeId IOJ) S]OO} [ 38K} ‘ays Surssal019 :punoi3yoeg oY oN quiosjopH] Adie Medivag youyeg Aare euowey pue Aayreyd JaAvag jarueq :($)1ayeY ‘S[OO} JToy} Pue Stayeur ayo-nu-ny jo ArewumMs— | _qIGVL Winter 2001 JOURNAL OF ETHNOBIOLOGY 11 Hickory, Cherokee ku-nu-che makers do not use that common name. Other than distinguishing “scalybarks” (C. ovata [Miller] K. Koch), they classify hickory trees simply as “‘hickernuts.’’ None of them expressed a preference for any particular kind of hickory nut, although a few specified that “pignuts’” (the local name for C. cordiformis [Wang] K. Koch), are too bitter. Thick-shelled species such as C. lacinosa (Michaux f.) Loudon and C. ovata (Miller) K. Koch that grow primarily on terraces of larger streams are known to be gathered by people who have access to them. Cherokee people like and eat pecans (C. illinoensis [Wang] K. Koch), few of which grow in the Cookson Hills surrounding Tahlequah, but cannot use them for ku-nu-che either because of the hard, sharp septal tissues or the failure of pecan meat to form the correct constituency of ku-nu-che when pounded, or both (our consultants disagreed on the limiting factor). Black walnuts (Juglans nigra L.), like pecans, are much easier than hickories to shell by hand and do not lend themselves to mass pounding due to the ridged nutshell that traps bitter-tasting residue from the messy outer hull. Yields of hickory nuts, like those of pecans and walnuts, fluctuate from year to year. Producers and consumers of ku-nu-che expect that nuts will be rare dur- ing bad years, and take the fluctuations in stride. We managed to purchase a few balls from Narcy Holcomb in June, 1999, even after two consecutive bad years, but most people today seem reconciled to wait for the next bumper crop. Fortu- nately, this occurred in the fall of 1999. THE PROCESS OF MAKING KU-NU-CHE BALLS Before processing can begin, hickory nuts must be dried for several weeks. Boxes or bags of whole nuts are left near a wood stove if either the nut gatherer or the ku-nu-che ball maker has one in their home. The meat of well-dried nuts separates more readily from the shell than does the meat of freshly fallen nuts. Ku-nu-che producers use a diverse array of tools for cracking and pounding hickory nuts, but people we interviewed all divide the process into two main stages. First, nuts are cracked one at a time. Narcy Holcomb uses and Blue Rock used custom-made metal tools (See Table 1 for a summary) that were welded for them in machine shops. Blue Rock’s nutcracker was a mallet made from two hollow metal pipe segments welded at right angles to each other (Figure 2). The openings at the ends of the shorter segment, which come into actual contact with the nuts, are covered by metal. Blue Rock set each nut, one at a time, on a base consisting of a flat, rectangular iron block approximately 25 cm long, 13 cm wide, and 4 cm thick. As shown in Figure 2, the block was set inside a cardboard box, and cracked nuts were then pushed off to the sides. Narcy Holcomb uses a metal, semi-cylindrical cracking tool designed by her husband, with an expanded, flat working end opposite a rounded end that she covers with a cut-off sock to protect her hand (Figure 3). For supporting the nuts she uses an iron base set inside a box, like Blue Rock, but her metal base is round. Daniel Beaver and the Careys also use metal cracking tools: a standard hammer and a large (18-20 cm long), unhafted, firewood-splitting wedge, respectively. Mr. Beaver cracks nuts on a con- crete slab. The Careys crack nuts on top of a cylindrical iron pedestal less than FRITZ et al. Vol. 21, No. 2 5 ie 3, * oe" qe as = of at - FIGURE 4.—Patrick Bearpaw holding the wooden pestle passed down from his grandfa- ther, Lee fatermelon, to his father and then from his father to him. Patrick ica We wide end of the pestle to crack nuts on a flat rock with a concavity in the cent =a 4 he uses the narrower end to pound the sifted nuts inbide-n cbffendmad er, and he uses Winter 2001 JOURNAL OF ETHNOBIOLOGY 13 13 cm in diameter and approximately 10 cm high. This pedestal is placed inside a box lined with a towel or other fabric. Several of the older consultants said they preferred metal to stone hammers because metal will not spall, but acknowledged that they used stone tools in the past or else had observed others using smooth, round rocks. Patrick Bearpaw uses the slightly wider end of a ca. 1.2 meter long wooden pestle that was passed down from his grandfather for cracking nuts (Figure 4). He cracked the nuts on top of a flat rock that he reported had become increasingly concave with use. Because this rock had been lost after the winter of 1998-99, he expected to search stream beds for a new rock for the 1999 season. Nuts must be cracked one at a time in order to avoid contaminating the ball with worms or with bitter, spoiled nutmeat. Patrick Bearpaw’s wooden pestle is wide enough on the nut cracking end to handle several nuts at a time, but he stressed that he cracks one nut at a time—occasionally two at the very most—so that he will not have to discard good nutmeat mixed with bad during multiple crushing. One or two initial blows reveal whether or not the nutmeat is usable. Each good nut is cracked into rather large pieces during the cracking stage. After five to ten whacks, the fragments—shell and all—are pushed off the metal, stone, or concrete base onto the lining of the box, and the next nut is cracked. Between the first stage (cracking) and second stage (pounding), larger pieces of nutshell are removed by sifting. Narcy Holcomb uses a large-holed aluminum colander (Figure 5), and the Careys use a standard kitchen colander through which they have punched a number of larger holes (Figure 6). Blue Rock used a 2 Ib 7 oz coffee can with screwdriver-sized holes punched in the bottom (Figure 7), and Patrick Bearpaw uses a plastic bow] with holes punched in it. The Careys and Patrick Bearpaw save the nutshell to be used as fuel in their wood-burning stoves. The others currently discard the nutshell, although Blue Rock burned it when he lived in the country and had a wood stove. The second stage involves pounding the nutmeat together with the small pieces of nutshell that passed through the holes of the sifter. This process is nec- essary not only to crush the solid fragments into very small pieces, but also to release the fats into an oily or “gummy” constituency that allows the meal to be shaped into balls. Our consultants engage in pounding for 30 minutes or more per batch. The Careys use large batches—a dishpan-full—and have a large wood- en mortar, so it can take 50 minutes of pounding before the meal is ready to be shaped into balls. Pounding tools and basins, again, vary according to the individual specialist. The Careys, who use a traditional, hollowed-out wooden tree trunk or “stump” as a mortar (Figure 8), have the most unconventional “pestle,” custom-made from four segments of ca. 1.4 meter long reinforcing bar (“rebar”) welded at one end onto the long sides of an unhafted sledge hammer head (Figure 9). Narcy Hol- comb uses a ca. 60 cm long wooden pestle custom-made by her husband for pounding ku-nu-che (Figure 10). She sits on a chair and pounds inside a large tin can. Blue Rock used an aluminum baseball bat (Bombat™ brand) to pound inside an iron stockpot (Figure 11). Daniel Beaver uses a heavy cylindrical steel curtain weight to pound ku-nu-che inside a square-sided wooden box that he made and affixed to a wider and heavier wooden base for steadiness. A square sheet of thin 14 FRITZ et al. Vol. 21, No. 2 FIGURE 5.—Metal colander used by Narcy Holcomb to sift larger pieces of nutshell after cracking and before pounding. yr ff , FIGURE 6.—Metal colander used by the Careys to remove larger pieces of nutshell between cracking and pounding. Winter 2001 JOURNAL OF ETHNOBIOLOGY 15 metal is inserted into the inside floor of the box to form the pounding surface. This avoids splintering and allows the box to last longer. Ku-nu-che balls tend to vary between 7 cm and 9 cm in diameter. Towards the end of his career, Daniel Beaver began selling the ku-nu-che meal loose inside plastic baggies rather than shaping it into balls, reasoning that the first step in the soup-making process is to break the ball back up into loose meal or dissolve it in hot water. DISTRIBUTING THE PRODUCT All but one of our experts sell their products without advertising. Ku-nu-che balls are sometimes commissioned ahead of time, with avid patrons furnishing the maker with more than enough hickory nuts to meet the buyers’ needs, as mentioned earlier. Other interested clients begin inquiring about availability in November and December, and information spreads through the grapevine. Three of our consultants had no telephones, and it is likely that quite a few people who end up with their ku-nu-che balls also live without telephones. Word of mouth, therefore, is still a key mechanism for ku-nu-che distribution. Much of this com- munication takes place at church gatherings. Patrick Bearpaw volunteered that people drive to his house to buy ku-nu-che balls from 50 or 60 miles (up to 100 km) away. Whitekiller and McIntosh have both observed balls in the offices of employees of the Cherokee tribal government and at Cherokee-run hospitals and health clinics. These balls had either been purchased on the premises or were available for purchase if one were to ask. Some FIGURE 7.—Coffee can with holes punched through the bottom used by Blue Rock to remove larger pieces of nutshell before Stage 2 pounding. 16 FRITZ et al. Vol. 21, No. 2 FIGURE 8.—Hollowed-out tree trunk used b y the Careys as a mortar for pounding already cracked and sifted hickory nuts. of these balls, in the recent past, were produced by Blue Rock and Daniel Beaver. Patrick Bearpaw, Narcy Holcomb, and the Careys easily sell as many balls as they want to distribute out of their homes. such as birthdays and ann served ku-nu-che sou Tsa-La-Gi Heritage iversaries. Ramona Carey and her associates have Pp made from the previous year’s balls on the grounds of the Center for many years during Cherokee National Holiday, Winter 2001 JOURNAL OF ETHNOBIOLOGY 17 Danby FIGURE 9.—The Carey's pestle: four rods of “rebar” (reinforcing bar) welded onto a sledgehammer head. The rods are taped together with duct tape. Mr. Carey holds the tool upside down in this photograph to display the pounding end. which is held over Labor Day Weekend, in early September. The Careys also enjoy eating ku-nu-che at the monthly gatherings of their large family. COOKING HICKORY NUT SOUP A ku-nu-che ball contains many small fragments of nutshell. Two balls, both made by Narcy Holcomb, were weighed separately before cooking and the nut- shell weighed afterwards, having been strained through a flour sifter. The balls were found to consist of between 22% and 25% nutshell by weight. The recipe from Cherokee Cooklore (Ulmer and Beck 1951), provided above, does not call for straining to remove the nutshell, but rather for leaving a residue of nutshell frag- ments in the bottom of the pot. Everyone we consulted, however, including several 18 FRITZ et al. Vol. 21, No. 2 <4 ee An, Saat \ Sent 1 a a. Ce S Se ee Fw AN d by Narcy Holcomb for pounding nuts inside a metal cooks who buy ku-nu-che balls but do not make them themselves, remove the shell fragments after dissolving the ball in hot water. The recipe in Cherokee Cook- lore is also unusual in that it does not hominy or rice, although it does mention bread or dumplings. Our experiences until soft but leaving much water , She put the ball into a one-quart (ca. 1 about two cups (ca. 0.5 liter) of hot water from add boiling water). The ball melted into a milky Winter 2001 JOURNAL OF ETHNOBIOLOGY 19 FIGURE 11.—Blue Rock’s baseball bat (Bombat™) in use as a pestle, pounding cracked and sifted nuts inside a stockpot. emulsion. Breaking up the lumps with a fork, Mrs. King poured this thick white fluid into a bow] through a standard flour sifter to remove the nutshell, and then added the hickory solution to the hot rice and unabsorbed water. Patrick Bearpaw said that a cloth is used for straining nutshell in his family, and Ramona Carey uses a sifter without a metal stirring apparatus (Figure 12). As the published recipe indicates, degree of thickness is a matter of personal preference. Few Cherokees today, young or old, eat ku-nu-che without sweetening it with sugar. Due to health concerns, Mrs. King adds artificial sweetener rather than sugar. A few people we talked to said they know someone who prefers salt to sugar, and salting rather than sweetening the soup seems to have been more common in the past. Ku-nu-che is served hot, but eaten at community gatherings after it has cooled to room or outdoor temperature. The high fat content causes the soup to thicken as it cools. Those fortunate enough to have leftovers in their refrigerators can enjoy cold ku-nu-che. WHY MAKE KU-NU-CHE BALLS TODAY? The people we interviewed who spend many hours each year cracking and pounding hickory nuts engage in this task for three main reasons. First, they are making a product that other members of their family and community desire. The product is particularly significant because it is a traditional Cherokee food, passed down through countless generations and key to the survival of their ancestors during famines. Narcy Holcomb’s commitment extends to reintroducing ku-nu- che to native Muscogee communities near Okmulgee, west of Cherokee country, 20 FRITZ et al. Vol. 21, No. 2 : ae RK— zr. ; te, % % | Sy ~ amg, 8 > —— FIGURE 12.—Ramona Carey’s sifter for straining out small pieces of nutshell while cooking hickory nut soup. Other cooks use muslin cloth or flour sifters with rotating handles. where the tradition of hickory nut pounding seems to have been discontinued (Muscogee people prefer sofkee, a food more traditional for their tribe, made from sour corn meal mash). Second, most Cherokee nut processors enjoy eating ku-nu- che very much and want more than a few balls for themselves and their house- holds. Blue Rock was the most avid ku-nu-che eater encountered during our study. His response to the question, “Why do you make ku-nu-che?” was “I love ku-nu- che.” Three of our consultants (see Table 1) also engage in the making of ku-nu- che in large or small part for economic reasons. When Patrick Bearpaw’s father stopped making ku-nu-che four or five years ago, Patrick saw an opportunity to earn a significant amount of money as a teenager while working at home. At the same time, however, he knew he was providing relatives and other members of his father’s clientele. The work is very hard, ev- were younger because the past. When sources were no longer available, antitative information such as the number cracking, and pounding nuts; total number of balls amount of money earned. Blue Rock volunteered that had earned $400.00 selling ku-nu-che that season. He had one more burlap bag full of nuts to proce ss at the time. Few if any of our consultants make more than 100 balls pe r year, and the going price is $5.00 to Winter 2001 JOURNAL OF ETHNOBIOLOGY 21 $6.00 per ball. Patrick Bearpaw’s mother told about a year when her husband made 100 balls within a few days and had badly swollen hands. In addition to those who make dozens of ku-nu-che balls for distribution, there are many other Cherokee men and women, probably helped by adolescents and children, who crack, sift, and pound enough hickory nuts for their families to have hickory nut soup at Thanksgiving, Christmas, or some other special oc- casion. THE FUTURE OF KU-NU-CHE In order to assess how younger Cherokees viewed ku-nu-che, a survey was administered in April, 1996, by Whitekiller to 28 female students living in the dormitory and attending Sequoyah High School, a Cherokee Nation tribally op- erated boarding school in Tahlequah. The students ranged in age from 14 to 19 years and represented 18 various North American tribes, diverse in singular or multi-tribal heritage. Twenty-four (86%) of the students claimed tribal affiliation belonging to either the Cherokee or Creek, with each of these tribes representing an equal number of 12. In addition to the question determining tribal affiliation, students were asked the following questions: What is your favorite food? Do you know what ku-nu-che is? (A NO answer terminated the survey.) Have you ever eaten ku-nu-che? . If so, on what occasion: (a) family gathering; (b) church gathering; (c) cul- tural gathering such as a stomp dance; (d) holiday such as Christmas, New Year’s Day, or Thanksgiving; (e) no special occasion? How often do you eat ku-nu-che: (a) one time only; (b) about once a year; (c) about once a month; (d) about once every two weeks or more often? How important do you think ku-nu-che is to Indian culture: (a) very im- portant; (b) somewhat important; (c) not at all important? Do you know how to make ku-nu-che? If so, who taught you how to make it: (a) grandparent(s); (b) parent(s); (c) other relative; (d) someone else? If you do not know how to make ku-nu-che, would you be interested in learning how to make it? If YES, who would you ask to teach you: (a) grandparent(s); (b) parents(s); (c) other relative(s); (d) someone else? BON ee o Eighteen (64%) of the students listed their favorite food as being pizza or hamburgers, with the remainder naming their preferred cuisine as Mexican, Chi- nese or “Indian tacos.”” When asked about ku-nu-che, 14 (50%) of the students answered they did not know what it was, nor had they ever tasted it. Of this number, nine (64%) claimed no Cherokee tribal affiliation. Four others identified multi-tribal lineage including Cherokee. One student who identified herself as being only Cherokee did not know what ku-nu-che was and had never tried it. Fourteen (50%) students reported they knew what ku-nu-che was, and 12 of the 14 had tried this food. For those who were familiar with ku-nu-che, 11 (79%) named themselves as being Cherokee, while three (21%) claimed no Cherokee tribal affiliation. The two respondents who stated they had not tried it claimed Native heritage to more than the Cherokee tribe. When asked about the occasion(s) 22 FRITZ et al. Vol. 21, No. 2 on which ku-nu-che was served, nine students (64% of those familiar with it) reported it was served on holidays such as Christmas and Thanksgiving. Two (14%) answered that ku-nu-che was present at family gatherings, with the re- maining three respondents indicating they had seen this food served at cultural gatherings, church meetings, and for no special occasion, respectively. In response to the question, ‘“how often do you eat ku-nu-che?,” five (42% of those who had eaten it) answered they had it once a year. Four (33%) reported they had tried it one time, two (17%) had it about once a month, and one (8%) indicated she ate it about once every two weeks or more often. Six (50%) of the 12 students who had eaten ku-nu-che responded they believed it to be ‘very important’ to Indian culture. Out of these six, two indicated they knew how to make it and were taught to make it, in one case by her parents and in the other case by ‘someone else.’ The remaining four stated they would be interested in learning how to make ku-nu-che, with two indicating they would ask their parents or another relative to assist her. Two responded they would ask someone other than family to teach them how to make it. Five (42%) of the students stated they believed ku-nu-che to be ‘somewhat important’ to Native culture. Four of these five indicated they would be interested in learning how to make it and would ask their grandparents or other relatives to teach them. One responded that although she thought ku-nu-che was somewhat important to Native culture, she had no interest in learning how to make it. All of these students with the exception of one responded that they were members of the Cherokee tribe. The remaining student (not identified as Cherokee) indi- cated she thought ku-nu-che was not at all important to Native culture and she had no interest in learning how to make it. To summarize the results of this survey, half of the 28 high school females residing in a Native American boarding school in Tahlequah and representing various tribal affiliations were familiar with ku-nu-che. Most of the students who knew what ku-nu-che was and had eaten it identified themselves as Cherokee and reported they had eaten the food at least once a year during holidays. All but one of this group of students indicated they believed ku-nu-che was very important or somewhat important to Native culture. Most in this group who did not know how to make it expressed a desire to learn and stated they would ask their parents or another relative to teach them. In spite of a drop-off in frequency of ku-nu-che use during the late twentieth century, the tradition is stil] fairly strong. A demand for the balls exists in Tah- vet i and women of various ages have demonstrated their willingness to take on the work of cracking and pounding hickory nuts after available sources dried RDS ee ee Pease a hi ait eal Winter 2001 JOURNAL OF ETHNOBIOLOGY 23 rate, and no extended training period is required. Any motivated, able-bodied person can do the job. Incentives are both economic and cultural. ARCHAEOLOGICAL IMPLICATIONS OF MODERN KU-NU-CHE The early ethnohistoric record, to our knowledge, makes no mention of solid balls made of sifted and pounded hickory nutmeat mixed with smaller pieces of nutshell. Therefore, archaeologists have emphasized liquid products, especially the milk and oil rendered from boiling nutmeat and cracked nutshell (e.g., Gard- ner 1998; Reidhead 1981; Talalay et al. 1984). Cherokee people commonly use metal tools to make ku-nu-che balls today, but most use stone or wooden tools in at least one stage of the process, and all report that their ancestors used stone and wooden tools to make ku-nu-che in both the near and distant past. We can think of no technological reasons to dismiss the practice of forming nutmeat and shell into balls before European contact, and good reasons to infer that native people—especially those who were not fully sedentary or who gathered hickory nuts some distance from their dwellings—reduced the weight and bulk of the nuts by making balls close to the source. This would have been easier than car- rying either bags of whole nuts or pots or skins full of oil when overland transport was necessary. This strategy might not have been workable in parts of the country where warm weather persists into late autumn, because balls would have spoiled within weeks without refrigeration. Many parts of the Eastern Woodlands, how- ever, are cool enough by November for storing oily balls for several weeks at least. Archaeological reports that include only counts or only weights of nutshell are inadequate for determining whether or not an assemblage represents the ac- tual cracking and pounding stages. A low total nutshell weight—even with a relatively high nutshell fragment count—might be interpreted as indicating that hickory nuts were insignificant at a site where a great deal of pre-sifted nutmeat mixed with many small pieces of shell had been imported from elsewhere. A ratio such as number of fragments of nutshell divided by their weight would be more revealing than count or weight alone, although post-depositional factors at specific sites must be carefully considered. A second ethnoarchaeological implication of modern ku-nu-che making is that the process involves two main stages: first cracking and then pounding, with sifting in between. Cracking is conducted one nut at a time so that bitter nutmeat and worms do not contaminate the meal. This is significant for at least two rea- sons. First, archaeologists who have experimented with cracking hickory nuts found the process to be more time-efficient when they started and ended with a wooden mortar and pestle in which numerous nuts could be crushed all at the same time, rather than reducing nuts to small pieces using only a grinding stone and hand-held mano that could crush only one or two nuts at a time (Reidhead 1981). It seems, however, that the process is not initiated in the mortar, although Patrick Bearpaw does use a wooden pestle to crack nuts—one at a time—over a large stone base. Regardless of which tools are used for Stage I cracking, each nut is whacked only a few times. Cherokee ku-nu-che ball makers do not use Stage I tools to render nuts into small pieces. Instead, they eliminate large pieces of coarsely cracked shell by sifting, then transfer the loosened nutmeat and small- 94 FRITZ et al. Vol. 21, No. 2 er shell fragments to a mortar or mortar-like metal container for pounding into fine particles. Another implication of the two-stage process is obviously that two sets of tools would have been seen as necessary or at least highly desirable. Archaeolo- gists tend to associate wooden mortars and pestles with maize rather than nuts, but mortars—both wooden and bedrock—may have been used for thousands of years before maize was introduced into eastern North America. It is extremely interesting that wooden mortars are identified as ‘‘ku-nu-che blocks” or “ku-nu- che stumps” at two historical sites in eastern Oklahoma: Tahlonteeskee (Figure 13), and the birthplace of Sequoyah, inventor of the Cherokee alphabet. This ter- minology would be consistent with a developmental sequence in which wooden mortars retained their original Cherokee name even after they came to be used mostly for pounding maize rather hickory nuts. Even in regions where all nuts were rendered directly into milk or oil rather than into an intermediate solid ball form, a two-stage process means that two sets of tools were probably involved whenever possible. A final lesson learned from modern ku-nu-che makers is that they consider hickory nutshell to make good fuel, and some people burn it in their wood stoves even today. This is a minor point, but the question has been raised during dis- cussions of taphonomy and the degree to which hickory nutshell is over-repre- sented in the archaeological record (see Lopinot 1982:729). Frequent use of nut- shell for fuel is likely to have increased the numbers of fragments in the archae- ological record of open sites and wet rockshelters, even though many specimens would have burned to ash in the process. If hickory nutshell had not been rou- tinely and purposefully burned as fuel, a far higher proportion would have rotted away over the years. CONCLUSIONS Hickory nuts were for thousands of years a staple food and the source of cooking oil and soup stock used by ancestors of the Cherokees and other Eastern North American Indians. After intensification of maize agriculture (ca. 1000 c.e.), hickory nuts remained a highly valued supplement. They were the source of fla- vorful oil and stock used for cooking various dishes in which maize was usually re pereey ingredient. Hickory nuts also constituted a critical fallback or famine ood in years when crops failed. Several hundred years after initial European at the Practice of rendering hickory oil apparently ceased, but the process si Big ae then pounding nuts and shells en masse and storing them in the i Ais s to be cooked either alone or with hominy in soup-like dishes sut- ived. Throughout the twentieth century, the Cherokee dish known as ku-nu-che isted ae Raliaey as a highly appreciated and frequently-to-occasionally served traditional Makers of ku-nu about their tools. No en tools are unavail -che balls in Oklahoma today are pragmatic and flexible thie mes: implements are used if old-style stone or wood- e. Metal tools have advantages over stone tools that might 1 Ie eens m8 wooden tools that might decay. Several producers have built Sioned unique tools used only by themselves for the purpose of crack- N J Winter 2001 JOURNAL OF ETHNOBIOLOGY FIGURE 13.—Traditional wooden mortar and pestle on display at Tahlonteeskee, near Gore, Oklahoma, capital of the Western Cherokee Nation between 1828 and 1839. The sign in the window shows a Cherokee woman using a mortar and pestle, along with the words ““Ga-Na-Ge Ka-No-Na: The big end gives weight to pound corn (selu) or hickory nuts (ga- nu-ge) in the (ka-no-na) or stump.” 26 FRITZ et al. Vol. 21, No. 2 ing or pounding hickory nuts. Sifting baskets have been replaced by metal col- anders and sifters. In spite of the popularity of new, modern tools and the use of tools that look very different from their ancient counterparts, the process always proceeds through the stages of cracking nuts one at time, sifting out the larger pieces of nutshell, and then pounding the smaller pieces of shell and nutmeat until enough oils are released to allow the maker to form the mixture into balls. The making of balls out of hickory nuts may not have been described ethnohis- torically, but we see no reason to doubt that the practice has considerable antiq- uity. Today, hickory nut soup is served less frequently than in the past, but all signs point to its survival. Although the production of ku-nu-che balls consumes a good deal of time and demands physical labor, no lengthy apprenticeship or extraordinary skills are required, and appropriate tools can be purchased or fash- ioned without great expense. Younger Cherokees demonstrate the motivation to carry on the tradition out of dedication to their heritage, a desire to reinforce cultural identity, and a sense of responsibility to satisfy the desires of elders, combined with the incentive to make extra money. We hope that ku-nu-che will be enjoyed by tens of thousands of Cherokees for generations to come. ACKNOWLEDGMENTS We are grateful beyond measure to Patrick Bearpaw, Daniel Beaver, Ramona and Char- ley Carey, Narcy Holcomb, and Blue Rock for sharing their valuable knowledge. We extend sincerest sympathy to the Rock family and thank them for allowing us to observe some of the last ku-nu-che making hours of their beloved grandfather, father, and husband. Among the many other people who helped us during this project, we are indebted to Florence Cornell, Johnny and Luke Drywater, Lois Johnson, Melvina King, Julie Reid, and Betty Smith. Mrs. Cornell’s aid was especially crucial in that she helped conduct the survey at Sequoyah High School and shared ku-nu-che traditions. Mrs. Smith told us how she makes ku-nu-che and provided useful comments on a draft of the manuscript. Sincere thanks go to Fiona Marshall, Nancy Turner, and Patty Jo Watson for also reading earlier drafts of the paper and suggesting improvements. Maria Bruno and Pascal Boyer kindly furnished the Spanish and French abstracts. Research was partially funded by a Faculty Research Grant from the College of Arts and Sciences, Washington University in St. Louis. LITERATURE CITED ~~ rae ee oe I. FORD, Southeastern Archaeological Confer- are ~ . 1972. Paleoeth- ence Bulletin 41:25 (Abstracts of the nobotany of the Koster Site: The Archa- 55th Annual Meeting of the Southeast- ic Horizons. Illinois State Museum Re- ern Archaeological Conference, Green- a Bs of Investigations 24, Springfield. ville, South Carolina. angst pices - 1922. The domestic | GARDNER, PAUL S. 1997. The ecological pe pasha : ni earn aborigi- structure and behavioral implications 0 fe nthropologist 24:171- mast exploitation strategies. Pp. 161- 178 in People, Plants, and Landscapes: Studies in Paleoethnobotany, Kristen J. Gremillion (editor). University of Ala- ama Press, Tuscaloosa. HAMMETT, JULIA E. 1992. Ethnohistory of aboriginal landscapes and land use DETWILER, KANDACE R., RENEE B WALKER, AND SCOTT C. MEEKS. 1998. Berries, bones, and blades: recon- structing Late Paleoindian subsistence economy at Dust Cave, Alabama. Winter 2001 in the southeastern United States. Southern Indian Studies 41:1-—50. HARPER, FRANCES. 1958. The Travels of William Bartram: Naturalist’s Edition. manuscript: Historical sketches of the Cherokees, — with some of their customs, traditions, and superstitions, Jack Frederick Kilpatrick (editor). thropological Papers 77, pp. 175-213. Washington, D.C. KILPATRICK, JACK FREDERICK. 1966. In- troduction. In The Wahnenauhi manu- script: Historical sketches of the Cher- okees, together with some of their cus- tome, traditions, and superstitions, by Lucy L. Keys, pp. 179-183. Smithsonian Institution, Bureau of American Ethnol- ogy Bulletin 196, Anthropological Pa- pers 77. Washington, D.C. LAWSON, JOHN. 1709. A New Voyage to Carolina: Containing the Exact Descrip- tion and Natural History of That Coun- try; Together with the Present State Thereof; and a Journal of a Thousand Miles, Travel’d Thro’ Several Nations of Indians; Giving a Particular Account of Their Customs, Manners, etc. [s.n.] (no publisher’s name: this is an original edi- tion in Special Collections). London. LOPINOT, NEAL H. 1982. Plant macrore- mains and paleoethnobotanical impli- cations. Pp. 671-854 in The Seat Mills Archaeological Project: Hum Adaptation in the Saline River Valley, Richard W. Jefferies and Brian M. Butler (editors). Center for Archaeological In- ee Southern Illinois Universi- ty, Carbondale. MALONE, HENRY THOMPSON. 1956. Cherokees of the Old South. University of Georgia Press, Athens MANKILLER, WILMA, AND MICHAEL WALLIS. 1993. Mankiller, A Chief and JOURNAL OF ETHNOBIOLOGY 27 Her People. St. Martin’s Press, New rk. WIGHT M. 1986. Trees of Ar- kansas wits revised edition). Arkansas estry Commission, Little Rock. MUNSON: PATRICK J. 1986. fidaxy sil- viculture: a subsistence revolution in of the Eastern Woodlands, Carbondale, coral PERRY, M eae 1974. Food u wild” plants 7 Cherokee iis M.S. thesis (Food Science), University of Tennessee, Knoxville. REIDHEAD, VAN A. 1981. A Linear Pro- ciety, Prehistory —— Series Vol. VI, 0. 1, Indianapo TALALAY, LAURIE, DONALD R. KELLER, and PATRICK J. MUNSON. 1984. Hick- relevant to their aboriginal exploitation in eastern North America. Pp. 338-359 in Experiments and Observations on Aboriginal Wild Plant Food l Utilization in Eas rth America, Patrick J. Munson Saal ndiana Historical se tory Series, Vol. VI, No. 2 is. ULMER, MARY and SAMUEL E. BECK. 1951. Cherokee Cooklore. Published by Mary and Goingback Chiltoskey, Mu- seum of ie Cherokee Indian, Cherokee, RIC the paleofeces. Pp. 41-54 in The Prehis- y, Patty Jo Watson (editor). Illinois State Museum, Reports of Investigations 16, Spring- field. , and M. JEAN BLACK. 1985. Tem- s fr ica. a me jer oie A r 94-107. Journal of Ethnobiology 21(2): 29-51 Winter 2001 PURSUING THE FRUITS OF KNOWLEDGE: COGNITIVE ETHNOBOTANY IN MISSOURI’S LITTLE DIXIE JUSTIN M. NOLAN Department of Anthropology, University of Missouri Columbia, MO 65211 ABSTRACT.—This study investigates ethnobotanical knowledge variation in Little Dixie, a folk cultural region in Central Missouri. Data were obtained from twenty and twenty “novices” who free-listed the names and uses for wild plants and rated them according to cultural usefulness, ecological value, beauty, and overall appeal. It is hypothesized and demonstrated that novices privilege species that are perceptually distinctive and ecologically abundant, while experts emphasize species with high use potential. Accordingly, novices emphasize beau- ty, a form-based variable, in their evaluation of listed species, while experts em- phasize cultural utility, a function-based variable. These results suggest that the acquisition of ethnobotanical expertise entails a shift from morphological, imag- istic information processing to the cognitive assimilation of abstract, utilitarian factors gained through learning and cultural experience. Key words: folk biology, cognition and expertise, free-listing, U.S. regional cul- RESUMEN.—Fste trabajo investiga la variacién del conocimiento etnobotanico en Little Dixie, una regién cultural popular en Misuri central. Los datos se obtuvi- eron de veinte ‘‘expertos” y veinte ‘‘novatos’’ que escribieron una lista al azar de los nombres y los usos de plantas silvestres y las calificaron de acuerdo a la utilidad cultural, valor ecolégico, belleza, y el atractivo general que tienen. Se hace hipsétesis y se demuestra que los novatos privilegian las especies de plantas que son perceptualmente distintivas y ecol6gicamente abundantes, mientras los ex- pertos hacen hincapié en las especies que tienen potencial alto de utilidad. Como corresponde, los novatos acentdan la belleza, una variable basada de forma, en su evaluacién de especies puestas a lista, mientras los expertos ponen énfasis en la utilidad cultural, una variable basada de la funcién. Estos resultados sugieren que la adquisicién de competencia etnobotanica conlleva un cambio morfolégico, pro- cesamiento de. informacién basada de imagenes a la asimilaci6n cognitiva del resumen, factores utilitarios ganados por el aprendizaje y la experiencia cultural. RESUME.—Cette étude examine la variation de connaissances éthno-botaniques dans le Little Dixie, une région culturelle du Missouri central. Les données ont été obtenues de vingt ‘‘experts’’ et vingt “novices” qui ont énuméré les noms et les usages de plantes sauvages et les ont évaluées selon leur utilité culturelle, leur valeur écologique, leur beauté, et leur attrait général. Il est démontré que les nov- ices privilégient les espéces qui sont perceptuellement distinctes et abondantes dans l’environnement alors que les experts prétent d’avantage attention aux es- péce qui ont un usage potentiel élevé. En conséquence, les novices soulignent la beauté, une variable basée sur la forme, dans leur évaluation des espéces énu- mérées alors que les experts soulignent I’utilité culturelle, une variable basée sur 30 NOLAN Vol. 21, No. 2 la fonction. Ces résultats suggérent que l’acquisition d’expertise éthno-botanique présuppose une modification allant du traitement morphologique et imagée de l’information a l’assimilation de facteurs abstraits et utilitaires grace a l’étude et a l’expérience culturelle. INTRODUCTION Ethnobiological knowledge is a complex phenomenon based fundamentally on human recognition of the perceptual and functional attributes that characterize living things. Over the past two decades, considerable progress has been made toward understanding how people transform their natural worlds into meaning- ful cultural categories (e.g., Brown 1984, Hunn 1982, Berlin 1992, Medin and Atran 1999, Ford 2001, etc.). Relatively neglected, however, is the study of variation within ethnobotanical knowledge systems. Research indicates that the differences in how people perceive biological domains are related to levels of respondent expertise, whereby experts have access to more kinds of information about a do- main than do novices, resulting in different patterns of domain organization. For instance, Boster and Johnson (1989) demonstrate that novices rely on mostly mor- phological cues when learning about and classifying marine fishes, while experts make use of morphological signals in addition to utilitarian information gained through personal experience. However, it remains yet undetermined whether or not experts and novices emphasize common referential features in their concep- tualization of plants or if they maintain separate patterns of ethnobotanical cog- nition. To answer the question, this project will explore the structure of ethno- botanical knowledge among residents of a regional culture in the U.S. Midwest. SCOPE OF THE STUDY . A defining feature of expertise is the ability to recognize and process multiple inds of information about a cognitive domain. For example, becoming an expert : ing from expert- pats understanding of physics problems (Chi et al. 1981) a yok nates ( esgold et al. 1988), to studies of how connoisseurs and amateurs appreciate wine (Solomon 1997) and art (Hekkert and Van Wieringen 1997) wo hypotheses stem from these collect ; ive findings. Gi d differences in how experts and novices appr po tbe SeenRPne oach and process information about Winter 2001 JOURNAL OF ETHNOBIOLOGY 31 Ne FIGURE 1.—Little Dixie Counties of Missouri. a domain, it follows that novice and expert plant users emphasize different focal attributes in their cognitive articulation of wild botanicals. That is, novices are expected to prioritize species that are perceptually distinctive and ecologically abundant, while experts should focus on species with salient use potential. Sec- ondly, it is proposed that novices prioritize beauty, a form-based variable, in their appreciation of plants, and that experts emphasize utility, a function-based vari- able, in their plant evaluations. DESCRIPTION OF THE STUDY REGION “Little Dixie’ is the name given to the corridor of gently rolling farmland that straddles the Missouri River in the central section of the state. In an historical account of slavery and cultural life in Little Dixie, R. Douglas Hurt (1992) pro- poses a map of the area that includes Callaway, Boone, Cooper, Howard, Saline, Lafayette, and Clay counties (Figure 1). Situated roughly between the corn belt and the Ozark Mountain region, Little Dixie represents a transition zone of the United States where the glaciated plains join the Interior Highlands to the south. The landscape is ecologically diverse, and supports between 80 and 90 native plant species that are absent or rarely found elsewhere in the state (Yatskievych 1999). The region’s physiographic character is one of rolling prairies, savannas, upland forests, and sandstone bluffs along the streams and rivers. Oak, hickory, and cedar predominate in the timbered hills and bluestem-dominated tallgrasses carpet the fields and savannas. Birch, maple, poplar, and willow are common along the bottomlands of the Missouri River and its numerous tributaries. The Cultural Landscape.—Little Dixie has been described as ‘’a section of central Missouri where Southern ways are much in evidence—an island in the Lower 32 NOLAN Vol. 21, No. 2 Midwest settled mostly by migrants from Virginia, Kentucky, Tennessee, and the Carolinas, who transplanted social institutions and cultural expressions to the new landscape” (Marshall 1979:400). Many of the early migrants were prominent families whose plantations and fortunes were built around farming tobacco, hemp, cotton, and indigo across the farmlands of the Upper South. These wealthy aristocrats brought with them their Southern culture, including a plantation econ- omy that involved the use of slaves and the sale of crops to the commercial market. Other settlers of Little Dixie included subsistence farmers, merchants, builders, and teachers who also originated from Kentucky and Virginia. While the Civil War brought an end to slavery and plantation life in Little Dixie, the tenacious Upper South cultural heritage has persevered in lives and minds of the people. The distinctly Southern identity of Little Dixie is apparent today through the local dialect, antebellum architecture, foodways, traditional music, and the strong influence of the Democratic party (Crisler 1948; Marshall 1979, 1981; Skill- man 1988; Hurt 1992). Agriculture remains a strong component of the present- day economy in Little Dixie, where soybean, hay, wheat, corn, cattle, and hogs are commonly raised. The economic base has diversified considerably to include education, health care services, manufacturing, and a strong retail and wholesale industry, each of which has brought growth and progress to the region. Wild Plants, Social Relations, and Group Identity—The people of Little Dixie are devoted to a lifestyle of relative independence. One of the ways in which people maintain and express their self-sufficiency is through the frequent and regular Procurement of wild plants for a variety of purposes. A number of local species are valued for their purity and wholesomeness, and, in some cases, for their rarity. Whether enjoyed as food, taken as medicine, or valued aesthetically, wild plant Procurement plays an important role in the social lives of the women and men of Little Dixie. The knowledge and work required in locating these plants from the outdoors and preparing them for personal use is developed over time by participating in family walks outdoors, helping out in the kitchen, and listening to the stories of mothers, fathers, and grandparents. Procuring and sharing wild plant resources symbolizes a neighborly communion with the local landscape, the sharing of personal skill, effort, and craftsmanship, a reverence for traditional customs, and the expression of group identity. METHODS AND MATERIALS ae a _ aie the Patterns of variation in ethnobotanical knowledge were BEAN a Pia Dixie, 20 experts and 20 novice (non-expert) consultants fom the seven counties within Little Dixie’ t of the respondents were selected from ixie’s borders. Mos to vary substantially among ex to ensure an adequate repres Winter 2001 JOURNAL OF ETHNOBIOLOGY 33 included both males and females with both commercial and non-commercial in- terests in wild plant use. Some experts operate private herbal practices, others sell botanical products at stores or from their homes through mail-order business or have contracts to cultivate selected species, while others are simply local people— from farmers to schoolteachers—who have exceptional knowledge of local flora. Novices also included male and female Little Dixie natives of mixed ages, but for whom wild plant collecting is neither a commercial activity nor a serious hobby. Both expert and non-expert consultants were selected by reputation (Martin 1995), followed by the “snowball’’ technique (Bernard 1994) in which one respondent recommends another, who in turn recommends another, and so forth. Using the same interview protocol for experts and novices, both groups were consulted during interviews that spanned from the summer of 1997 to the fall of 1999. Interviews consisted of a semi-structured interview containing open-ended questions, free-listing, and a sociodemographic survey. To begin the interview, consultants were casually queried about their personal experience with local flora. Questions included “how did you come to know about wild plants?” and “what do you find meaningful about using wild plants?”. The first section of the survey included a free-list task (Weller and Romney 1988, Bernard 1994), an effective elicitation tool for ethnobotanists (Martin 1995, Cotton 1996). Respondents were asked to write down the names of as many kinds of locally available, useful wild plants as they could think of, using their own judgment of what is considered useful. Respondents were then asked to indicate how each plant is used (e.g., medicinal, edible, ornamental, etc.), the specific application for the plant (e.g., pie filling, heartburn remedy, etc.), the part of the plant that is used (e.g., stem, root, etc.), and the mode of preparation (e.g., air-dried, boiled in water, etc.). This data collection process, known as successive free-listing (Ryan et al. 2000), provides a rich, descriptive database for examining plant use patterns, and has been used in a number of ethnobotanical surveys. There is reason to believe that experts and novices exhibit different expressive and aesthetic evaluations of the constituents of semantic domains! (e.g., Chick and Roberts 1987), which may in turn effect how domains are organized cognitively (Nolan and Robbins 2001). To explore these differences, a rating exercise was administered with the free-list task in which respondents of both groups were asked to assign a number between one and five to each named plant based on the evaluation of four different variables: overall appeal, usefulness, ecological value, and beauty. The mean ranks were calculated on all four variables for the most commonly mentioned plants, and a multiple correlation analysis was per- formed on these ranks to determine how the two groups compare in their con- ceptual evaluation of salient species. RESULTS Analysis of the Free-Lists.—Of the 187 plant names collected from both groups, experts listed a total of 160 plants, comprising 85.6% of the composite list. For the experts, list lengths ranged from 12 to 61 plant names, with a median of 25.5. The mean list length was 26.4 plant names, with a standard deviation of 13.3 and a coefficient of relative variation (CRV) of .504 (see Table 1 for a quantitative 34 NOLAN Vol. 21, No. 2 TABLE 1.—Number of wild plants and applications reported by experts and novices. Number of plants mentioned Number of applications listed Experts Novices Experts Novices Mean 26.7 9.1 37.4 11.1 Median a3 8.5 36 10.5 no 2 20 3.8 18.9 4.9 Maximum 61 17 88 21 Minimum 12 5 14 2 summary of free-list results, and Appendix 1 for an inventory of all listed species and uses). The total number of applications for wild plants listed by experts was 749, representing 77.2% of the total. The number of applications listed ranged from 14 to 88, with a median of 36. On average, experts listed 37.4 applications with a standard deviation of 18.9 and a CRV of .505. Novices listed a total of 79 wild plant names, constituting 42.2% of the com- posite plant listing. The length of the novices’ plant lists ranged from 5 to 17, with a median of 10.5. The mean list length was 11.4 with a standard deviation of 3.8 and a CRV of .333. Novices listed a total of 221 applications for wild plants, or 22.8% of the total inventory. These applications ranged in number from 5 to 21, with a median of 10.5. The mean number of listed applications for novices was 11.1, with a standard deviation of 4.9 and a CRV of .441. A comparison of the two groups reveals, as expected, a higher mean number of plants free-listed by the expert consultants. The difference in means, 26.4 plants listed by the ex- perts and 11.4 for the novices, is statistically significant (t = 5.4, p < .001). Sta- tistical significance was also found for the difference in the mean number of ap- plications reported, 37.4 for experts and 11.1 for novices (t = 6.02, p < .001). Figure 2 graphically displays the positive correlation between the number of plants and the number of applications reported by both groups. As shown in Figure 2, knowledge of plant utilization rises incrementally with an increase in plant-naming knowledge for both consultant groups. The number of plants named and the number of applications reported are significantly correlated for novices (r = .87, p < .001) and experts (r = .91, p < .001). While there is some overlap between the level of ethnobotanical knowledge demonstrated by the two groups, the expert-novice distinction is reasonably clear, as indicated by the dis- persal of data points on Figure 2. The Salience of Listed Plants.—The B values given in Table 2 measure free-list sa- lience, or the proportional precedence of a listed plant over others. B is computed as follows: ROTA IY 2 Dn) oe ea eee B where 1 is the number design designated subset items and designated subset items ( for each plant free- ated subset items, fi is the number of complement r(n) is the sum of the free list ordered ranks of the é Robbins and Nolan 1997). Here, a B value was computed listed by experts and novices. To calculate individual salience Winter 2001 JOURNAL OF ETHNOBIOLOGY 35 Number of Plant Uses Reported wn o R? = 0.7554 0 10 20 30 40 50 60 70 Number of Plants Reported FIGURE 2.—Correlation of number of plants reported to number of plant uses reported in free-lists for experts and novices. values for a given plant on a free-list, n = 1 and fi = (the total number of listed items) — 1. Ranging between 0 and 1, the B value for a given item reflects the relative proportion of other items it precedes on the list. The B value for each species was summed across all lists and divided by the number of respondents listing the plant to generate a composite B value. To calculate a measure of overall cultural significance, the composite B value for each listed species was added to the proportion of respondents listing the plant and divided by 2. As seen in Table 2, there are more plants with higher frequencies of mention on the experts’ inventories than among the novices’. Consider, for example, the three plants mentioned most frequently by experts—blackberry, dandelion, and walnut, which were listed by 18, 15, and 14 experts, respectively. These frequen- cies are high compared to the three plants mentioned most commonly by nov- ices—raspberry, dandelion, and blackberry, which were listed by only 12, 12, and 11 novices, respectively. Interestingly, three of the five most frequently mentioned species (blackberry, dandelion, and walnut) are the same for experts and novices. All three of these plants can be used in a number of practical ways. For instance, walnut is a valu- able source of food, medicine, lumber, and dyes. Blackberry is also highly ven- erated for its edible berries, known locally and in the Ozark Mountains to the south as “black gold,” and for the food value of its young shoots and its medicinal roots that are often brewed into healing tonics to treat colds, fevers, and colic. 36 NOLAN Vol. 21, No. 2 TABLE 2.—Frequency and salience of plants commonly listed by experts and novices. Experts Novices Rank Plantname Freq. % B Plant name Freq. % B 1 Blackberry 7 OD 0.579 Raspberry a SD 0.35 2 Dandelion 15 0.75 0.434 Dandelion 12 60.6 0.498 3 Walnut 14 0.7 £0.345 Blackberry 11 0.55 0.404 4 Gooseberry 13, 0.65. 0379: Walnut ii OR: Oe 5 Sassafras 13. 0.65 0.377 Mulberry 10... O48 0.241 6 Lamb’s quarters 12 0.6 0.338 Sunflower 1005 0.25 7 Hickory 12%.06 0.33 Pine 9 045 . 02S 8 Pokeweed 7. oe Ore Cattell 5 0.187 9 Plantain a2). 050°" Q,315 s 6: 08 0.136 10 Persimmon 10 05 0.302 Wild onion 6 °\03 0.17 11 Wild mint 10. 205: 0241 a 6. De 0.185 12 Dewberry 10... 05 0.29 orel 5 0.25 | O08 13 Sunflower 9 045 0.212 Wild apple 5 U2 7 0ie 14 ak 9 045° 06S a 5 02 O12 15 Burdock 9 045 0.265 Black-eyed Susan 4 0.2 0.093 16 Raspberry 9 045 0.324 Wild strawberry a 702 0.112 17 Morel 8. 04. 0.138..«. Paw 4 0.2 0.101 18 Wild onion 8 04 0.21 Marijuana 4 0.2 0.128 19 ulbe BUS. O14) | Sassafras 4 0.2 0.084 20 Wild grape Se Goldenseal o 'O.49) OMe 21 Cedar 8 04 0.154 ~—Hicko 3° 0.15 0.074 22 Wild plum 8 04 0.232 Wild cherry 3 0.15 0.033 23 Wild strawberry 7 0.35 0.177 Wild rose 3. 0.15... 0.114 24 Paw paw 7 0.35 0.221 Honeysuckle 3 0.15 0.088 The dandelion is similarly edible; its young leaves and flowers are eaten by both humans and animals, and like the others, it is used regionally in medicinal tonics to treat chills and fevers. Well-known even by those with minimal interest in local flora, it is no surprise to find these species at the top of the list for the novices as well as the experts. Most interesting, however, are the differences between the two sets of re- spondents. As seen in Table group or the other. Among thos by experts, are pine, cattail, daisy, several plants appear exclusively ters, gooseberry, dewberry, plantain, persimmon, and burdock. One explanation for this pattern is the novic and ecological salience (e.g., Turner 1988). Plants that are morphologically dis- tinct, bearing obvious physic frequently among the untra available in the ambient environment. For the most part , the perceptual distinctiveness and ecological abun- dance of these species prob among novice consultants. On the other hand, species with relatively higher free- list frequency among the experts (e.g., lambsquarters, plantain, burdock) lack the i ‘ easily distinguishable Winter 2001 JOURNAL OF ETHNOBIOLOGY 37 70 4 @ Experts 60 - O Novices Number of species af ¢ e nee eB cs liam 10, topes 16 18 20 6 8 10 12 1 Number of respondents reporting use FIGURE 3.—Number of reports of use for all species listed by experts and novices. features that characterize species with high perceptual salience. Weed-like herbs such as these are not immediately obvious to the untrained eye. Nonetheless, they are emphasized cognitively by the experts who are knowledgeable about their practical uses*. To illustrate, the leaves of lambsquarters and burdock are prized for their flavor, edibility, and nutrient value, and plantain leaves are used exten- sively by experts as a bandage or a poultice for exterior wounds. The Diversity of Wild Plant Knowledge.—Figure 3 displays the number of reports of use for all wild plant species named by experts and novices in the free-listing task. While the overall knowledge pattern for experts and novices is similar, this abundance diagram conveys an interesting pattern that seems to characterize the plant knowledge of the two groups. That is, experts demonstrate a higher dis- persal of knowledge, which is reflected by the higher number of unique, once- mentioned species listed among them. As shown on the diagram, considerably more plants were reported by a single expert (93 species) than were mentioned by a single novice (39 species)*. There are fewer instances in which several novices listed the same plant. Alternately, experts demonstrate a higher overlap of listed items. The overall pattern suggested by the abundance diagram is one in which experts have command of a greater diversity of plant knowledge than novices, resulting in both a higher proportion of collective, commonly shared knowledge and a higher level of esoteric, idiosyncratic knowledge in the form of once-men- tioned species. 38 NOLAN Vol. 21, No. 2 Sd + e _ rs S oe * rs * . ? ° * ie Oo | On Oo o e a a a} oOo a Oo a oO Oo a DO Novices Da Oo @ Experts FIGURE 4.—Multidimensional scaling of positive matches between experts’ and novices’ free lists. From a qualitative perspective, the differences between the experts’ and nov- ices’ free-lists are also considerable. To determine the overall extent of free-list similarity, the number of positive matches between listed items was calculated for experts and novices in order to compare the two groups. The resulting coordi- nates were plotted using multidimensional scaling, or MDS, using the software package ANTHROPAC 4.95 (Borgatti 1998). MDS is a useful technique for visu- alizing the relations between points or items, whereby points that are closer to each other in two- dimensional space are thought to be more similar than points perts and one for novices—rather than a single shared system. Winter 2001 JOURNAL OF ETHNOBIOLOGY 39 Contrasting Plant Use Patterns.—After each respondent was asked to list as many useful wild plants as they could think of, he or she was prompted to name as many uses for each plant as possible. A review of the collected applications yield- ed a total of seven different use categories for the named plants: food, medicine, wood /lumber, ornamental, wildlife forage, handicrafts, and other. All wild plant applications on each free-list were coded with their corresponding use categories”. On occasions when consultants offered several categories of use for the same plant, each category was recorded. The number of applications that fell into each category was summed and converted into percentages by dividing by the total number of applications reported by that group. As displayed in Figure 5a and Figure 5b, food is the most commonly named use category for the plants listed by expert and novice respondents. At 48% and 52% of the total applications cited by experts and novices respectively, food is also the most culturally fundamental use for wild flora. In Little Dixie, edible plants constitute an important part of the traditional foodways that help char- acterize the region. The custom of gathering wild fruits, berries, and nuts from the local woods is shared and enjoyed by most local people, regardless of their level of botanical expertise, which probably accounts for this shared pattern of use. The remaining use categories, however, are considerably different with re- spect to the proportion of applications cited by experts and novices. The second most commonly mentioned category for the experts is medicinal plants, compris- ing a sizeable percentage (38%) of the total reported plant uses by experts. The prevalence of edible and medicinal plants in the expert pharmacopoeia reflects the interest and knowledge in holistic living and natural healing that is pursued and practiced by a number of the expert herbalists who were consulted. The remaining uses given by experts were rather evenly distributed into the decreas- ingly smaller categories of wood/lumber, ornamental, wildlife forage, other, and crafts. Among the novices, the food category was followed by ornamental (16%) and wood /lumber (11%). The relatively high percentage of ornamentals listed by nov- ices reflects a significant pattern through the course of this project—the novice predilection toward a perceptually oriented knowledge of wild plants. Ornamen- tal plants are deemed meaningful and useful by virtue of their physical charac- teristics and visual appeal. Knowledge of ornamentals is readily available to the novice, for it requires only an aesthetic appreciation for the beauty of form—and knowledge of the name of the plant—but not experience with use and function. Comprising only 6.5% of the total uses reported, the medicinal use category ranked fifth in frequency for the novices, after wood /lumber (11%) and wildlife forage (7%). To compare the overall diversity of the plant use categories for experts and novices, the index of qualitative variation (IQV) was applied to the plant appli- cation data. Ranging between 0 and 1, the IQV measures the degree of evenness in the proportional distribution of a sample. The higher the IQV value, the more uniform or balanced the distribution is deemed to be. The IQV is computed as 40 NOLAN Vol. 21, No. 2 Sten aS ctf baie ine soc aa a poeie ste ae es FIGURE 5a.—Distribution of expert uses for plants. FIGURE 5b.—Distribution of novice uses for plants. b+ > pp Ask where Pi is the Proportion of plant reports represented by each category and k is the number of use categories. For the experts, the IQV yields a value of .78, and for the novices the IOV is .79. These results indicate that, for each group, the Winter 2001 JOURNAL OF ETHNOBIOLOGY 41 relative degree of evenness in the distribution of plant applications is extremely similar. That is, the seven use categories show a moderately balanced represen- tation for each group. While the IQV measures distribution or evenness, the index of dissimilarity (D,) is useful for assessing quantitatively the differences in overall use patterns. D, is calculated as D,=5D1P.- Pal where P, is the proportion of expert plant applications in each category and P,, is the proportion of novice applications in each category. The index of dissimilarity also generates a value between 0 and 1, where 1 indicates perfect dissimilarity and 0 indicates perfect similarity between the groups’ categorical distribution. Calculating the index of dissimilarity generates a D, value of 24%, which means that 24% of either group’s distribution would have to change in order to match the other group’s distribution. So where are these differences coming from? While the proportion of appli- cations listed as food is very similar for the two groups, experts know consider- ably more about medicinal plants than novices, who report far more plants as ornamentally useful. Experts are also more intimately involved and experienced with plants in general, and have acquired through time a more extensive under- standing of the cultural uses of plants—particularly the therapeutic aspects. While it takes an expert to understand how to use plants medicinally, anyone can ap- preciate the beauty of a given species and deem it worthy of ornamental display. This very fact may explain why novices report a much higher number of plants in the ornamental category. Novices know less of the esoteric medicinal functions of wild flora, which requires a level of botanical knowledge and interest more characteristic of expert respondents. The Expressive Evaluation of Wild Plants.—In descending order, the correlations be- tween the rating scores for experts and novices are: ecological value = .70 (p < .001), usefulness = .49 (p < .05), preference = .46 (p < .05), and beauty = .36 (p > .05). These r-values reflect the similarity with which experts and novices rated the plants, especially with regard to ecological value. It is noteworthy, however, that the groups do not correlate significantly when rating the plants according to beauty. These findings agree with those by Chick and Roberts (1987), who deter- mined that machinists and non-machinists rated lathe parts very similarly with respect to complexity, but very differently with regard to beauty. Like the dis- covery by Chick and Roberts, these results show that the two groups agree most on the highly denotative variable, ecological value, and least on the most conno- tative variable, beauty. Table 3 lists the intercorrelations among the four rating variables for experts and novices. For both groups, personal preference appears to be the most impor- tant underlying dimension in the evaluation of the wild plant domain. That is, plants that are preferred are also considered useful, ecologically valuable, and beautiful. One interesting expert-novice distinction is clear, however: the corre- lation values between usefulness and beauty. For the experts, there is a low cor- 42 NOLAN Vol. 21, No. 2 TABLE 3.—Multiple correlation of mean ranks of wild plants on four variables (experts’ values shown to the left, novices’ values in parentheses). Variable Preference Usefulness Ecological value Beauty Preference 1 Usefulness 0.72"** (0:68)""7 Ecological value 0.74*** (0.78)*** 0.55* (0.44)* 1 Beauty 0.62** (0.66)** 0.39 (0.92)*** 0.68** (0.57)** 1 **p < 001, “p < 01, *p < .05. relation for the two variables (.39), yet for the novices, the correlation is very high (.92). The difference between these r-square values was tested and found to be significant (z = 3.31, p < .001). In fact, the difference in r-square values between usefulness and beauty is the only significant disparity between the two groups. This difference, taken in concert with the low rating correlation on the beauty variable, indicates that novices emphasize beauty as an organizational factor in the conceptualization of wild plants. Novices are restricted to purely visual stim- uli when abstracting an emotional and/or cognitive impression of a given plant. It follows that a plant’s usefulness is a function of its overall perceptual appeal, or beauty. The salience of beauty in wild plant evaluation would also explain the high proportion of ornamental plants free-listed by novices. On the other hand, beauty is significantly de-emphasized in the determination of usefulness in the mind of the expert. Experts have more criteria for usefulness at their disposal (e.g., nutritional value, medical efficacy, etc.). Any of these esoteric factors are most likely used in concert by experts when evaluating the usefulness of different plants. Thus, it is evident that the accumulation of expertise entails a shift in domain appreciation, or how the domain is evaluated and organized from an expressive point of view. The rating patterns by the two groups indicates that experts and novices have contrasting standards for appreciating wild plants, which appears to be linked to underlying differences in how the domain is organized conceptually. SUMMARY AND CONCLUSION It has been shown, as predicted, that experts and novi ili ifferent referential features in their P ices utilize di perts’ and novices’ expressive pl j als t : : plant judgements reve _ Ae ee eonpnasize beauty while experts prioritize cultural value when rank- ing the species. These findings reaffirm that experts are influenced most by use- Winter 2001 JOURNAL OF ETHNOBIOLOGY 43 fulness and practicality, while novices are affected more by aesthetic variables in their organization of plant knowledge. Taken together, the results suggest that the acquisition of ethnobotanical knowledge entails a cognitive shift from mor- phological factors and sensory perceptions to a more complex comprehension of plants based on abstract, culturally acquired utilitarian factors. This information can be applied in a number of ways to understand how cultural experience shapes our comprehension and appreciation of our natural worlds. NOTES ' For example, Chick and Roberts (1987) examined the evaluation of lathe parts by machin- ists and non-machinists. The authors discovered that the machinists display more agree- ment regarding the expressive aspects of lathe parts than the non-machinists, due to the experts’ better understanding of how the parts are manufactured. 2 However, these plants are not absent altogether from the experts’ wild plant inventory— they appear further down on the composite list. 3 Again, the species discussed here do appear on the novices’ inventory, but with consid- erably lower rankings in frequency and salience. ‘Similar use report patterns by plant experts appear throughout the ethnobotanical liter- ature. For example, in a study of Mestizo plant use in rural Mexico by Benz and his col- leagues, many unique or once-mentioned species were listed by expert consultants (Benz et al. 1994). Accordingly, Nolan (1998) found that wild plant experts of the Ozark-Ouachita Highlands listed relatively high proportions of idiosyncratic species. Cognitive anthropol- ogists have found considerable knowledge variation to exist among expert respondents (e.g,. Boster and Johnson 1989, Nolan 2001). These studies offer something of a challenge to cultural consensus theory, which is built on the proposition that agreement or consensus among respondents is indicative of cultural expertise. 5 The boundaries between certain use categories are often “fuzzy,” particularly with respect to food and medicine. For this reason, it was necessary to code a number of plants into multiple categories, such as those used in spring tonics (e.g., sassafras, burdock, may ap- ple). For insightful information on the categorical overlap of food and medicine in people- plant interactions, see Johns (1996, 1994). ACKNOWLEDGMENTS the Society of Ethnobiology, Durango, Colorado. Funding for this research was generously granted by the National Science Foundation’s Doctoral Dissertation Improvement Award (BCS-9903983). 44 NOLAN Vol. 21, No. 2 REFERENCES CITED BENZ, BRUCE EF, FRANCISCO SANTANA M., Ri NEDA L., JUDITH CEVALLOS E., LUIS ROBLES H., and NIZ L. 1994. Character- ization of Mestizo plant use in the Si- erra De Manantlan, Jalisco-Colima, Mexico. Journal of Ethnobiology 14(1): 23-42. BERLIN, BRENT. 1992. Ethnobiological Classification: Principles of Categoriza- tion of Plants and Animals in Tradi- tional Societies. Princeton University Press, Princeton. BERNARD, H. RUSSELL. 1994. Research Methods in Anthropology. Sage Publi- , STEPHEN P. 1998. Anthropac 4.95. Analytic Technologies, Columbia, South Carolina. BOSTER, JAMES S. and JEFFREY C. JOHN- SON. 1989. Form or function: a com- parison of expert and novice judgments of similarity among fish. American An- thropologist 91:866-889. BROWN, CECIL. 1984. Language and Liv- ing Things: Uniformities in Folk Clas- sification and Naming. Rutgers Univer- sity Press, New Brunswick, New : CHI, MICHELENE. T. H., P. J. FEL’ ICH, and R. GLASER. 1981. Categori- zation and representation of physics problems by experts and novices, Cog- nitive Science 5:121-152, CHICK, GARRY and JOHN M. ROBERTS. 1987. Lathe craft: a study in “part” ap- is i Human Organization 46(4): COTTON, C. M. 1996, Ethnobotany: Prin- ciples and Applications. John Wiley and Sons, Chichester, England. CRISLER, ROBERT M. 1948. Missouri’s Lit- e Dixie. Missouri Historical Review 42:130-139. FORD, RICHARD (editor). 2001. Ethno- opti = the Millennium: Past Prom- ise and Future Prospects. Anthr. - fos Number 91. anes ogy, University of Michi Ann Pe Ay 4 pi HEKKERT, PAUL and PIET C. W VAN WIERINGEN. 1997. Beauty in the eye of expert and non-expert beholders: a study in the appraisal of art. American Journal of Psychology 109(3):389-407, HUNN, EUGENE. 1982. The utilitarian fac- tor in folk biological classification sys- tems. American Anthropologist 84:830- HURT, R. DOUGLAS. 1992. Agriculture and Slavery in Missouri's Little Dixie. University of Missouri Press, Columbia. JOHNS, TIMOTHY. 1994. Ambivalence to the palatability factors in wild food plants. Pp. 46-61 in Eating on the Wild Side, Nina L. Etkin, (editor). University of Arizona Press, Tucson. ———.. 1996. The Origins of Human Diet and Medicine. University of Arizona Press, Tucson. KEMPTON, WILLETT. 1981. The Folk Clas- sification of Ceramics: A Study of Cog- nitive Prototypes. Academic Press, New York. LESGOLD, ALAN, HARRIET RUBINSON, PAUL FELTOVICH, ROBERT GLASER, DALE KLOPFER, and YEN WA ates, Hillsdale, New ; MARSHALL, HOWARD WIGHT. 1979. Meat preservation on the farm in Mis- souri's “Little Dixie.” Journal of Amer- ican Folklore 92:400-417. - 1981. Folk Architecture in Little Dixie. University of Missouri Press, Co- mbia. MARTIN, GARY J. 1995. Ethnobotany: A hods Manual. Chapman and Hall, London. MEDIN, DOUGLAS L. and SCOTT ATRAN (editors). 1999. Folkbiology. MIT Press, Cambridge. ——— ELIZABETH B. LYNCH, and JOHN D. COLEY. 1997. Categorization and reasoning among tree experts: do all roads lead to Rome? Cognitive Psy- chology 32:49-96. NOLAN, JUSTIN M. 1998. The roots of tra- - In press. Wild plant classification in Missouri’s Little Dixie: Variation in a re- Bee oth eee ere sp a nee re Winter 2001 gional culture. Journal of Ecological An- thropology. , and MICHAEL C. ROBBINS. 2001. Emotional meaning and the cognitive organization of ethnozoological do- mains. Journal of Linguistic Anthropol- ogy. 11(2):240-249. ROBBINS, MICHAEL C. and JUSTIN M. NOLAN. 1997. A measure of dichoto- mous category bias in free-listing tasks. Cultural Ayie Ging Methods Jour- nal 9(3):8- RYAN, GERY Ww. JUSTIN M. NOLAN, and STANLEY YODER. 2000. Successive we -listing: a new technique for gener- ting sb aoe models. Field Meth- cis 12(2):83-— SKILLMAN, — a ape No smoke? no fire: contemporary hammin e ol’ po fashioned way. Pp. 125-136 in We Gath- JOURNAL OF ETHNOBIOLOGY 45 er Together: Food and Festival in Amer- ican Life, Theodore C. Humphrey and Lin T. Humphrey, (editors). University of Michigan Research Press, Ann Arbor. SOLOMON, GREGG. 1997. Conceptual change and wine expertise. The Journal of a rose: evaluating the cultural signif- icance of plants in Thompson and Lil- looet Interior Salish. American Anthro- MBALL ROMNEY. 1988. Systematic Data Col- lection. Sage slaareaatias Thousand Oaks, Californ YATSKIEVYCH, GEORGE. 1999. Steyer- mark’s Flora of Missouri (Volume 1). Missouri Department of Conservation, Jefferson City. Vol. 21, No. 2 NOLAN synu SIBMO}J pray paas “TPIS “J0}S}OOI SAARO| suaaid ‘SaAra] juryd “amoy ajoyM “stoMo}j SaAvg] ‘S}INIJ “SJOOI “SAl1I0q poom Salad ‘SAARa] poom SIaMOYy spaas ‘saavay ‘yuetd ajoym yuny dar] ajoyM ‘SHINY SPaas ‘SAALI] $}001 Poss auDIpaut ‘ea} aATepas yuoueey} a8emas “yeyUDURUIO ‘pooj ITU} 4LI IOJ pooy ‘autpaut ISvIOJ aJI[P]IM ‘Suiapse’s JOyEM asesoy AJI[P[IM asevsog aft[PIIM yaytind pooyg ‘euDrpout “pooj aulorpaul aseioy aytTp[iM suapies JaMmoy ‘[e}uoUTeUIO vay ‘aSvioj ayI[PTIM ‘ouTstpowt “Poo} [ios ur uaSoxru xy “Seroj aytTPEM a8vioj ayI[pyM ‘TeyUoweUIO sajpuvs Supyeur ‘sjjeso Jaquiny jeyueureUso Bd} ‘POOj “~upog ee vyojuap vauvysv>y + s1]vUutp4v9 vifaqo'] "| vsosaqny svidajasy “gas (°q) a le SHAOPOLUYIF | vyary viyoaqpny ‘dds *] snqny “| wiavov-opnasd vIuigoy "Y wnyofiunsd unuangi, NN (J) vsouaovs vsnfi2UtD "| vpavuivgjnp WinuvjOS ‘dds -y vjnjag “URUI}TA, Mpavsad uoSodoapuy aD vywuiyoa vjnddyy "| wuvoiaue ‘ULL ‘dds “| sajsy “| syppunyjo snsvivdsy “| DUBILAUY VAIYINAF{ ynuyseay7) iejs Suizelg uesng paso-peig jaamssonig ya waysenig Sig ay sies3ag pasn juejd jo weg quryd 310} sas ‘sasn paysodas 1194} pure se aurleu SQUIIDS OUIeU IP[NIPUIIA ads paysy-aery jo Azoyuaaut aytsodwop— | XIGNaAddV 47 JOURNAL OF ETHNOBIOLOGY =o a8esoy aytpyM ‘dds ~ oXvpyos posuapjoy jue weq ysy ‘Saad (J) vunnaSara vrsosyday ani syeor $7001 quepnwys “uLDIpaw “| snyojanbuinb xvung Suasurry Saara] “SjOOI auDIpaw “| vyofanbuinb pumyuas urquary Sacra] ‘SIaMOY aubDrpau | vaindind syvp3iq PAOTSXOy SAARQ] aubpipew “‘yuag ("]) winuayjand uinuayjuvshayd MapaAay Sacra] ‘SUa}s aL JOJ Poo; ‘dds ~] vonjsay ssei8 anosay suMOID [eyUaWTRWIO “pooy ‘dds ~] (uno) wntpodfjag sua SalLaq SUPIPIUW “Pooy “| SISuapyUun) snONquiVS Auaquepl| re uoupajoid ogenbe ‘dds uapraryps vjapoards paamyonq ode [equeureuso “| vployy snuso> poom3oq Surppid “poo; "] Suajoaams winyjauy a sna ‘saraq aSesoy az[|PLM “poo} PIM sHvja8vyf snqny AL1ag. MAC] SIaMOY eyuaureUO “| myn{ sijpvo0sauiazy AyyAeqg SaAea] ‘SIBMOY aBesoy agI[P[IM ‘OUDIPaUt “pooj yaqam auuryjo wnovxvayy uoTepueg sIaMoy yeyuaueW0 | winuayjunoma, unwayjuvshayD Asieq poom Jaquiny “UPR (7) Mamyorystp winipoxey ssaidA>y Salaq pooj ‘TPURM Minywsopo sagry quenN) re JAOD punois ‘dds *ys1apy vrawypdsiq sseiSqeiyy syinay pooy ‘dds “] snahg aiddeqeid SaAva| poo} “XUPIA, Manwyuy] wina]Iv4a}] drusied mo> yung Jaquiny "ysaryl saprojjap snyndag PoomUo}oD SIaMOlJ yeyuaureUsO "| snuvho vaanvyuay JaMO[UIO> syuryd suapie3 Jamo "WON V140j;9U1) sisdoaso> sisdoai0) SIMO JeyueUTeUIO “| sisuapyuv vi8ajinby auIquinjo> Ways JRO] auPpIpaw "| SMpLighy saqisvjad JOO}s}OD jyueyd ajoym Surxy ua8ontu ‘a8es0y ayt~pyLM "| suadas uinyofiay JAC) saarg| “A[P}S auTorpaut "| autavdy uinyyy SIBAPAT) Spaas advo} ayIypyM "| snquajnosa snaadhy ey yAeq ‘saLiieg auIpaw “| vuriuisata snunig Axraypayoyuy SIAMOY ‘SAARAT| ‘S}OOI Pa} ‘Pooy “| sngAja wniaoyny Asoo SWOsso]q ‘sudeI3 ‘sUa}s ‘SAARaT Pooy ‘ouldtpeu “YT DIpa vIAvI}EIS peam yoru) pasn jueyd jo jeg juejd Joy sascq AWPU IWUIIS aureu JeyNoeUIaA, Winter 2001 (panuyuod) | xrpuaddy Vol. 21, No. 2 NOLAN syinay spnq Jee] yuna ‘des ‘9013 ajoym “poom yTeIs ‘SuIa}s SALI] SOARV ‘SOLLIaq ‘TTR yazeg yun “poom ‘syjnu siaqy ‘spnq ‘sy]e}s ‘Saavey SIARD] synu ‘yuetd afoym Ss}OO1 Seavey ‘{]R}S Saliiaq ‘syndy jueld ‘saavay ’s}001 pooy ‘auttpaut aultpew ‘Surjouws “Bury}o]9 apeys “pooy ‘[eyusuTeUIO ‘“IaquIn] aSeIOJ IFTPPTIM yeyueweUIO pooy [los ut ua8ouytu xy [los ul uaso TU xy [eyUaWeUIO jayund ‘susai3 ‘pooj eyusUTeUIO earqpurm ‘poo ‘JeyUUTeUIO ‘aUTDTpoU o1u0} Sutids “aurpeur Aat uostod ‘eursrpeut JeyusWIRUIO [eyusWIeUIO “O8eIO} aFITP[IM [los ur ua8ojIu xy “s}yeID suapie8 Jamo ‘[eyuaUTeUIO aSvsoy FTP Suruevsyp> Pooy syusuM.ysut yeorsnur ‘sped 8urmoos S}JeID ‘JoquNy] ‘ose10j “POO} sjnpoid Jaded ’s}je19 “auisrpout uostod poo} ‘TeyuaUTeUIO “oSeIOJ aJI[PTIM dupe uortsoie dos ‘o8eioj ajt]ppIM aseioj ‘pooy Jayiind pooyq ‘eurrpeut “J mnywyjad wn hydopog “| valyys siquuuv> “| uinavyoous 4aay ‘dds -| uoSodoupuy "| siuv8jna v8uiuhs ysing (NN) vjopiday vzysshohpy ‘dds -xypryp vzapadsa] “Yysing suadsauvd vydiomy ‘dds 7 wniuiydjaq "| uingjv unipodouay) ‘dds 4 wnipadiudé5 “| vpuruisaia snaadiun{ “| unaindand unisowdnz "| upyjud suatpduy ‘WoUps (°q) uinpAydisy viuavsiay ‘dds “7 stuy ‘dds -yua,_ visiydvg ‘Suaids ~y (7) vau19909 vlayjiysv> poy “y (Sue) vyw2I0g vidvssnjhvo "JT asuaaav uinjasinbZz "| aavdjna wniqnsvyy aeaovog jo saidads snoriea ‘HNN asuatinossiu saqiy "| sisuapvuvd sijsvipAyy siayrenb s quie’y raddysApey Jodrun{ paam adq a0 paamjamof ydjnd-ay}-ur-yporf St] o8rpuy ysniqjured uerpuy ssei3 uvIpuy [easuaploy pasn jueyd jo y1eg jyuetd Joy sasQ JUILU ITUIIIS aueU JepNIVUIaA, (panuyuos) [ xtpueddy JOURNAL OF ETHNOBIOLOGY Winter 2001 SdAKV] ‘SIOMOTS aSesOJ AJI[PTIM ‘syasur Suyoesqye "| vjosva snoned aor] Seuuy useNnd) SUuId}S [eyueuTeULIO "YsIeyy SUMUNY X1]VS MmoTpLMAssn gf saarg] ‘SuaeI3 pooj "| vaIvAI]O VIVINJAO aurysing [[e ‘Siamoy ‘s}OOr ‘SaARe| aBvioj afI]P[IM “ouDTpaut ‘youaoyy (J) vaindind vaovuryaq Jamoyauoo afding [ro Gueyd aULIpeut “pooj ‘suapses IaMOT “| siuuatq vsayjouac asorWL SIIMOT} ‘SHIM ‘SAARI] pooj ‘yey (Jey) vsnfuuny vyundo read Appiig jyueyd asesO} afI[P[IM Aerry (Avis) sidajosajay snjoqosods peasdorp aired juryd yueqg wreadys azitiqeys “yury vywuijsed vurjavds sseispiO) oelg spaos pooj WaUIOP] vioyiqiy auomasiy Addog suaai3 ‘Sala ‘SAAR2] aUIIpaUT ‘s}JeII “Pooj | vuvriuauy voovjophyd paamoax0g [Ie ‘SIaMoTy ‘S}OOI ‘SBARVT Poo; ‘ouTTpouw 71 solv o8vyuvjg UTeyUR] sa[paau ‘sauod yun.) ‘Poom pooy ‘apeys ‘[eJueuTeUIO “JaquIn] J] vypurysa snuig aug jueyd aBev1OJ IJITP[IM ‘Suaspse3 J3}eM "| DIYs vLAapayUog paaM JaIoPIq spaas ‘s}INIy pooy "7 vuviuis41a soshdsoiq uOUTWISaq SdAPI] JUDIPoul ’V3} "sia (") sapioi8ajnd vuloapay{ jeAosAuuag yuryd pios ul usSosy1u xy ‘ASG (1) viopfiq sayjsuvsophys JOMOY fous poom ‘s}nu poom ‘pooy ‘y~poy “yy (‘SueM) sisuaourp vAawvD uedeg s}Inay pooy “| s1unmiuos snshg ieag sy pooy "7 voisuad snunig ypeag s}INAy pooy yeung (J) vgoji4y vurmusy Med meg SAARI] duUDIpew ‘| vyousvour vsopfissyg JOMOPUOIsSSeg yinsy ‘poom juayjadar yjout ‘poomay “prauyps (Jey) wiafituod vanjovyp a8ursio a8esc— poom ‘synu adaid ypnyt poos "7 uinuvjsvo0ddiy snjnosay ahayong oryO synu 4yuNI] ‘suIOde “‘poom = apeys ‘poomaly ‘a3vI1O} ‘S}JLID “Jaquiny ‘dds *4 snauan?) AeoO yreq Jouur ‘ued ajoym auUDIpeut “JUL UTRaI}s 9ZI[IqQe}s "7 snyofiyndo sndavooshyg yareqouln susai3 ‘yy ‘SAARI| pooj ‘auTstpouwt ‘s}je19 ‘dds "J vain SITHON spaas pooy dds “| voissusg paeysnyy saAvgy ‘yuetd ajoym jaded japio} ‘outpour ‘TeyusUTeUIO | snsdpyy winasvgsa, Ulan] SaLitog ‘S}INAy aprys “ouTSIpout “pooy | vigna sno Axsaquny\ SIAR] juatjedas ydasut “| SUVs ]NA VISUAL WOMBnyy sdo} ‘wroorysnur afoymM DUIDIPSU “Ppoojy ""] vyUaNISa VIJAYIAOWW joroyy suaei3 ‘SaAea| pooy ‘dds -y vonjovy] aon}YaJ_S JOU saavgy ‘pod “yyrur aBeIOJ aJI[P]IM ‘ouIDIpout “7 voviihs surdajasy pram pasn jueyd jo jeg jueyd 10} sasq auTeU IYYUIIIS OWeU TeTNOPUIOA, (panuyguod) , x1pueddy Vol. 21, No. 2 NOLAN swossoy]q ‘Sude13 ‘sIoMOy ‘SdaARe, = LUN J.Jod poo} ‘ouTDTpowt “[eyWOWIeUIO ‘dds "1 vjo1, POA re yeyuouTeUIO “wiaag (J) suvoipys sisduv> auta yodumniy, wla}s ‘peay yeyuouTeUIO “spnyy sizjsaajis snovsdiq Jeska], Saka] ‘JAMO yuatjedai ydasut | auv3jna wnjasvuyy, Asue], yun} Jaquin| "| sujpjuapiz90 snuvyw]d arowedAS yueyd JAZIIGe}S BoAg] ‘ABeIOJ ajI[P]IM J winwsiia wnoiung ssevi3 Yp}IMS jueyd ajoym JeyUSWIRUIO T vywoisvaip XojYd WIRTTIM 28MG reyau a8esoy ast[pyIM “pay Vgiy snjojtjaw IBAO]D IMS jueid “1amoy efouM ‘spaes ymay y1eq “SaltIaq jueyd ajoym yINJJ ‘SAARI ‘S}OOL ‘SoLL19q SoARV] “SJOOL aSvsO] aFITP[M ‘TeyUaUTeUIO ‘Ppoo} dtu} Surids ‘ousrpeur oubDIpeul yeyuauTeUIO d}iqeyeus JOJ JUSWyRAaI] “PUTPTpPeUt aBesOJ aJIPPTIM yeyusWeUIO aulorpour Jaquuiny{ ‘ea} ‘eUTIpeu “pooy pecs suapie3 Jamoly apeys ‘Fe}USWIeUIO auItpaur ‘asesoy ayt[P[IM sero [los ut uaSo.jtu xy aUIDIpaUt “pooj auPIpout "| snnuuv snyjuvijazy ‘dds 7 snyy day viadspqns vijUuvIsapvsl "| vywords vyyuay ‘dds *y xauny ‘dds “qr uinqwuoshjog ‘ysang vypursivu vigaoydny ‘Nop winsosn. uintsopwdny ‘dds *7 uinuoShjog TUN wigns snuyf| "| pipvaut uoayywoapod ‘Ipayy (J) siozsvd-vsang vpjasdvy "| pyjasojaov Xauiny ‘Uay (‘J XU) vasogay salyouvjaury “| VILPULILAVUL VISSVD "SOON Sasi: uinpiqy svafossvs “XU tuntjofiaonh uni Suhsq ‘dds *} vtavjojosa “XI SMsodiys snqny “| minipofiasoqur uiniuaysavg [21105 ureyunoUr asind s praydays Jeasos daays ysnqpeys PUUdS SPIJLSSES JIAOP pey pasn jueyd jo jeg jueyd Joy sas JUTLU ITFTUBINS OWPU IP[NOPUTOA, (penuyuod) [ xtpueddy JOURNAL OF ETHNOBIOLOGY Winter 2001 SaAeveg] ‘y1eq *S}OO1 SIAMOTJ ‘UI9}S FeO] suaai3 SdARV] ‘SY]P}S “09.4 BJOYM ‘eq s}Indj ‘SaLieq yes ‘SaARa] ‘S}OOI ‘Qing urei3 SOARD] SOUTA ‘S}INAT $]OO1 qing SdARZI “SUTO}S yinay “YAR ‘Salieq 997} BJOUM Tt? SUOSso]q ‘sudaI3 ‘SBARZ] poom ‘y71eq ‘sTNy “synu aultpeut “saytind poolq auDIpeut Pooj Pooy ‘Te}USUTRUTO “s}JeID ‘OUTDTpoU jaytind pooyq ‘autstpewl ‘pooy POO} ba} OUDIPeW ‘Pooy [e}UsUTeUIO “OUTM “pooj aulDIpour poo} ysture3 “pooj Jaquiny ‘sUIDIpeU “Pooy auLDIpow [los ul usSor}TU xy apeys [eyusureuIO ‘poomary ‘uostod ‘outsrpaut ‘pooy “7 sndsiso xauiny "| manyofarp variysy UMOJG “yf Slv3/Na vasvqsvg "7 vqiy xt]vS ‘| pupusa vivsvsq ‘dds *| vsoy ‘| vpuvoidauy snunid ‘| vols vIVUILISUd Jay Winywpjajs uN "| vyofijissas vlavjnan "| sisuadav vYyJUaW ‘dds “| sip, ‘| asuapouvs wnsvsy “| asuapyuvs wuNny|y “UJOP] (“J) winijofasao snosuuyjuy “YY VUlJOsas SNUNA ‘NN vuvINA0pny visa “| suadaa winyofiay "| vatuojhqug x1jg ‘Wy vywsopo vavyduhn Ig “yy avursyfo uinipingsuN ‘dds -- suvjsn{ POP MOTI ssa1D 10} ynuyeM pasn juryd jo yg jueyd Joy sascq JUILU ITWUSINS OUeU TPTNOPUIOA, (panuyuos) , xtpueddy a inch iit olnge , ae Winter 2001 JOURNAL OF ETHNOBIOLOGY 53 Black Rice: The African Origins of Rice Cultivation in the Americas. Judith A. Carney. Harvard University Press, Cambridge. Pp. xiv, 240, photographs, maps. ISBN: 0-674-00452-3 ($37.50, cloth). African slaves introduced the rice technology that made the Carolinas great in the 18 century. This has been known for two decades, but only now has a book appeared that treats adequately the botany and technology as well as the history and food ethnography involved. Judith Carney’s work is a major achieve- ment. Not only does it complete the effort of restoring to prominence an African- American contribution to American life; it also stands as one of the best short studies of the way a particular crop and its production technology influenced history. African rice, Oryza glaberrima, was domesticated at least 2,000 and probably more than 3,500 years ago in West Africa, quite independently of the earlier do- mestication of O. satim in East Asia. To produce it, process it, and cook it, complex and sophisticated technologies developed, especially along the coasts of Senegam- bia and Guinea. The Wolof, Mandinka, Baga, Mende and Temne were among the major peoples involved. As rice developed in the Carolinas, slaves from this re- gion became more important, and eventually most blacks in the United States were from the “rice coast.” Carney does not elaborate on the cultural effects of this beyond food technology, but it is to this that we owe the distinctive quality of black culture in the United States, especially in music, folktales, folk speech, and visual art. The blues derive from Senegambian traditional music, the banjo was a Senegambian instrument, and the words “hippie” and “hipcat’”” may be Wolof loans in English (see e.g. Palmer 1981). Carolina rice was almost exclusively O. sativa, apparently derived from Mad- agascar and India, but Carney shows that O. glaberrima was locally grown there and elsewhere in the New World. Eventually, O. sativa made it back to West Africa, where—alas— it now threatens to replace O. glaberrima, including many wonderful varieties developed over the centuries. The Carolina rice industry was thus built on the skills of the African slaves— as well as on the horrific exploitation of their labor, death from sheer exhaustion being common and routine in the 18 century. Thus, a brilliant and successful industry developed in Africa and America, but its developers got little beyond torture and death for their contributions. This book will surely become a classic in the literature on history seen through particular crops. It reminds one of the longer and more comprehensive works of Salaman (The History and Social Influence of the Potato, 1985), Mintz (Sweet- ness and Power, 1985) and the Coes (The True History of Chocolate, 1996). A small irony says it all. On page 72, we meet Captain John Newton, who in 1750 “bought nearly eight tons of rice for feeding 200 slaves” on his ship. Captain Newton was later to repent of his horrible trade, and spend years in deep de- pression and guilt. Finally finding solace in religion, he wrote the song “Amazing Grace.” This song, often sung in thoroughly Senegambian-derived style, remains vitally important in African-American communities today. Human achievement is a strange, ironic, often cruel thing, but sometimes it can—in the words of another spiritual—‘‘outshine the sun.” 54 BOOK REVIEWS Vol. 21, No. 2 This book adds to the many that document the African Diaspora’s contribu- tions to the New World. Until recently, African contributions were widely thought to be minimal. Pioneers in research in this field, such as Melville Herskovits and Harold Courlander, were ignored or depreciated. Apologists for the plantations and for racism denied that Africans could contribute; worse, many well-meaning writers were so anxious to show blacks as ‘victims’ that they ignored or dismissed Black cultural legacies. Today, many ethnobiologists, as well as musicologists, art historians, and others, have documented a great range of contributions. E. N. Anderson Department of Anthropology University of California Riverside, CA 92521-0418 LITERATURE CITED COE, SOPHIE, and MICHAEL COE. 1996. The True History of Chocolate. Thames and Hudson, London. Journal of Ethnobiology 21(2): 55-72 Winter 2001 CAVES, URSIDS, AND ARTIFACTS: A NATURAL-TRAP HYPOTHESIS STEVE WOLVERTON Grand Valley State University Dept. Anthropology Allendale, MI 49401 ABSTRACT.—European cave deposits often contain the remains of extinct cave bears (Ursus spelaeus and U. deningeri) and artifacts or human remains. Two twen- tieth-century explanations for the apparent association of the remains and artifacts are: 1) late Pleistocene hominids preyed upon the bears; and 2) late Pleistocene hominids and bears occupied the caves at different times thus making the remains and artifacts appear behaviorally associated when they are not. The former option is dismissed in most cases based on taphonomic criteria and ursid mortality data. In caves with multiple entrances—particularly cases where at least one entrance is a vertical shaft comprising a natural trap—another option serves to better ex- plain the presence of ursid remains and artifacts in the same deposits. Ursid-bone assemblages created by accidental entrapment of bears in vertical shafts result in a distinctive mortality pattern. This pattern reveals proportionally more prime adult individuals than expected in a living population. A consideration of North American black bear (U. americanus) physiology and behavior reveals that this distinctive mortality pattern should be expected from natural trap assemblages. Thus, in assemblages from caves with horizontal and vertical entrances, mortality data can be used to decipher whether ursids died from natural hibernation deaths, human predation, or accidental falls through vertical shafts. Key words: ursids, mortality, natural trap, cave. RESUMEN.—Los depésitos en las cuevas europeas frecuentemente contienen res- tos de osos extintos (Ursus spelaeus y U. deningeri) y artefactos. Dos explicaciones a la aparente asociaci6n de restos y artefactos son: 1) que los osos fueron victimas de los hominidos del pleistoceno tardio; y 2) que estas cuevas fueron ocupadas en diferentes momentos tanto por hominidos como por los Osos; permitiendo que temporalmente asociados, Fsta ultima situacién, que ocurrié poco frecuen- restos de osos y artefactos; particularmente en es vertical y profunda en la cual el oso cae en tamiento del oso negro de norte américa patrones distintivos en la mort restos encontrados en cuevas con entradas vertic empleados para decifrar si los osos murieron por 56 WOLVERTON Vol. 21, No. 2 humana, 0 por caidas accidentales en las entradas verticales y profundas de las vas. RESUME.—Les dépéts sédimentaires de cavernes européennes contiennent sou- vent des restes humains ou des objets faconnés associés A des restes osseux d’espéces éteintes d’ours des cavernes (Ursus spelaeus et U. deningeri). Il y a deux scénarios actuels pour expliquer cette apparente association de restes osseux et d’objets fagonnés: 1) les hominidés de la fin du pléistocéne chassaient les ours; et 2) les hominidés de la fin du pléistocéne et les ours ont occupé les cavernes a des moments différents, donnant la fausse impression que les ossements et les objets fagonné t iés culturell ans la plupart des cas, le premier scénario, qui semble s’étre réalisé rarement ou pas du tout, peut étre rejeté en utilisant des criteres taphonomiques et des données de mortalité ursidée. Pour les cavernes avec entrées multiples (particuliérement celles oi: au moins une entrée a des pa- roies verticales constituant un piége naturel), un autre scénario explique mieux la présence dans les mémes dépéts de restes ursidés et d’objets fagonnés. Les assemblages d’ossements ursidés crées par la chutte accidentelle des ours dans des entrées a paroies verticales semblent présenter un profil de mortalité distinct. Ce profil révéle une proportion d’ours d’age adulte plus importante que l’on pourrait le prévoir d’aprés une population vivante. Une revue de la physiologie et du comportement de l’ours noir nord américain (U. americanus) revéle que les piéges naturels devraient produire des profils de mortalité ursidée distincts. Pour les assemblages trouvés dans des cavernes avec entrées horizontales et verticales, les données de mortalité peuvent donc étre utilisées pour déterminer si les ursidés sont morts naturellement pendant I’hibernation, suite a la prédation humaine, ou suite a une chutte accidentelle dans une entrée a Pparoies verticales. INTRODUCTION The co-occurrence of cave-bear (Ursus spelaeus and U. deningeri) remains and artifacts in European caves has been interpreted during the twentieth century to Stiner 1998). Arguments that humans hunted cave bears [Abel and Kyrle 1931; Bachler 1940, 1957 (cited in Kurtén 1976, Stiner 1998)], though popular, have been dispelled by mortality and taphonomic data [Koby 1953 (cited in Kurtén 1976); Kurtén 1976; Stiner 1998; Webb 1988)]. The routine interpretation of apparently ities at different times, but another interpretation supported by a unique ursid mortality pattern merits consideration. Taphonomic histories of cave assemblages are complex (Arsuaga et al. 1997; Oliver 1989; Stiner et al. 1996, 1998; Wolverton 1996), and though taphonomic and mortality data indicate that humans did not regularly hunt cave bears, another accumulation agent (other than alternate use of the caves by ursids and humans) might explain the co-occurrence of cave-bear and human remains in some Eu- ropean caves. In particular, caves that have or had in the past horizontal and vertical entrances may have served as shelter to humans and traps to cave bears. Winter 2001 JOURNAL OF ETHNOBIOLOGY 57 Vertical shafts comprising natural traps act as accumulation agents that produce mortality patterns distinguishable from ursid hibernation-death assemblages. In some cases—one of which is discussed here—the taphonomic histories of cave- bear/artifact assemblages should incorporate natural entrapment of ursids as a possible accumulation agent because cave structures change through time (eg., Arsuaga et al. 1997)—entrances open and close. Stiner (1998, see also Webb 1988) provides detailed discussion of expected mortality effects of hibernation deaths and human predation. Hibernation-death assemblages should produce mortality patterns that are biased toward young and old adults—U-shaped mortality—because those individuals are more susceptible to attritional death agents such as disease or starvation (Stiner 1998; see also Lyman 1994a). Human predation on hibernating bears, on the other hand should reflect L-shaped mortality or “affect prime adults, old adults, infants, and ado- lescents randomly, emulating their natural proportions in the living population sequestered in dens each year” (Stiner 1998:309; see also Lyman 1994a). Natural traps attract young-adult bears; it follows that ursid mortality pat- terns from natural-trap deposits are biased toward high representation of young- adult remains compared to their representation in stable living populations. Two cave assemblages are discussed herein to demonstrate that this unique mortality pattern offers valuable taphonomic insight into the co-occurrence of cave-bear remains and artifacts/human remains in European caves with horizontal and vertical entrances. The first is a paleontological assemblage of North American black bear remains (Ursus americanus) from the Midwestern United States (central Missouri) that dates to the late Holocene (AA38931, 233 + 39; AA38932, 207 + 34; CAMS-27141, 170 + 60 C14 yr B.P.). The second assemblage comprises cave- bear remains (Ursus deningeri) from Sima de los Huesos (Spain), a cave with a vertical shaft and possibly buried horizontal entrances, the deposit of which con- tained cave-bear and human remains (Arsuaga et al. 1997). Lawson Cave-—Lawson Cave is located in central Missouri. The cave, in profile, is a bottle-shaped solution fissure formed through long-term dissolution of lime- stone parent material. The modern entrance is a 178 by 79 cm opening located along the top of a forested ridge (long axis oriented approximately east to west); this entrance drops 11% m straight to the cavern floor. The upper 3 m of the shaft are wet and mossy; the chimney opens into the southeastern portion of the cavern ceiling. The shaft widens as it extends down toward the cavern. A collapsed horizontal entrance conjoins the vertical shaft 4% m above the cave floor and runs west to east. When open the horizontal entrance would not have provided an exit from the trap because the lower cavern walls are steeply inverted. Lawson Cave's structure suggests it is unlikely that it served as a bear den. : Today the cave is moist with dripstone flowing from the ceiling. Portions of the cave floor were excavated during the 1950s, though the bedrock floor is COv- ered with as much as 1 m of sediment; the identified mammalian remains recov- ered from the cave are listed in Table 1. Visibility of the modern cave entrance is poor; the opening cannot be seen by humans in daylight from outside of 5 m in all four cardinal directions (Wolverton 1996). Because the sample of bear remains is small (10 individuals), I postulate two explanations for the preponderance of 58 WOLVERTON Vol. 21, No. 2 TABLE 1.—Taxonomic abundances at Lawson Cave. Taxon Abundance (NISP) Ursus americanus 445 Sylvilagus floridanus 238 Sus scrofa 170 Marmota monax 66 Canis sp. 66 Didelphis marsupialis 42 Neotoma sp. oo Microtus ochrogaster 19 Peromyscus sp. 18 Mephitis mephitis 12 ; Zz Odocoileus virginianus 5 Scalopus aquaticus Procyon lotor 1 Caster canadensis 1 omys bursarius 1 young-adult bears: (1) the Lawson Cave mortality pattern is the result of random capture of black bears from the (historically extirpated) central Missouri living population or (2) the pattern is not the result of random capture, but young-adult bears are more susceptible to natural-trap mortality than bears of other ages. As demonstrated below, the mortality pattern appears unlikely to be the result of random accumulation of ursids in Lawson Cave. Taphonomy of Lawson Cave.—Quantitative units used to discuss taphonomic vari- ables include: number of identified specimens (NISP), minimum number of ele- ments (MNE), and minimum animal units (MAU). NISP is the number of bone or tooth specimens (fragmentary and complete) identified to element and taxon. MNE is the “minimum number of complete skeletal elements necessary to account for observed specimens” (Lyman 1994b:290), or the number of elements repre- sented by the identified complete and fragmentary specimens. MNEs are calcu- lated by determining whether or not two or more specimens overlap; if two frag- ments overlap—e.g., one distal right humerus overlaps one complete right hu- merus—then the specimens must be from two separate bones, which equals an MNE of two. If the specimens do not overlap, then they could be fragments from the same element, hence the MNE would equal one. If two or more fragments (0F unfused Parts) refit, they equal an MNE of one. MAU is similar to MNE except it accounts for some elements occurring more or less frequently in one skeleton than others (e.g., one cranium vs twenty 1* phalanges in the same skeleton) by rile MNE by the number of times the element occurs in the skeleton (Lyman Intensity of fragmentation, calculated as an NISP:MNE ratio, monitors how many fragments (NISP) occur per distinguishable element (MNE). If Lawson Cave served as a bear den, then high NISP:MNE is expected because trampling i _ fragmentation intensity (Lyman 1994a; Stiner et al. 1995). NISP:MNE 14 se calculated for black bear (U. americanus) and cottontail (Sylvilagus floridan- us) long-bone remains from Lawson Cave; the ratios incorporate only frag™ Winter 2001 JOURNAL OF ETHNOBIOLOGY 59 TABLE 2.—NISP:MNE and %Whole for black-bear and cottontail long bones. Fragmentation intensity Extent of fragmentation NISP:MNE le Sylvilagus floridanus Humeri 10:8 = 1.25 15/23 = 65.2% Ulnae 12:12 = 1.0 2/14 = 14.3% Femora 21:14 = 1.50 11/25 = 44% Tibiofibulae 35:22 = Loe 12/34 = 32.3% Total 78:56 = 1.39 40/96 = 41.7% Ursus americanus Humeri 22:14 = 1S? 2/16 = 12.5% Ulnae 14:12 = 1.17 1/13 = 7.7% Femora 22:14 = 1.57 3/17 = 17.6% Tibiae 10:8 = 1.25 5/13 = 38.5% Total 68:48 = 1.42 11/59 = 18.6% as the purpose of the ratio is to measure the degree of fracture of broken specimens (complete elements are unfractured). Extent of fragmentation—calculated as %Whole—incorporates fragmented and complete black bear and cottontail long bones. It measures what proportion of the bones (MNE) are complete (Lyman 1994b). The Lawson Cave ursid and cottontail limb bones are extensively fragmented; the abundance of complete elements is low indicating most long-bones were frac- tured at least once (Table 2). Intensity of fragmentation for the rabbit and bear limb bones, however, is low (Table 2). Each broken identifiable ursid limb element is represented by 1.42 fragments; for cottontails the ratio is 1.39 NISP per MNE. An intensely fragmented assemblage results in several NISP per MNE (Lyman, 1994b:292); such is not the case here. Low intensity of fragmentation suggests that post-depositional processes (including carnivore damage and trampling) were limited likely because the deposit was well sheltered within the natural trap from weathering and other attritional agents. Extensive fragmentation—that is, the fact that most of the specimens are incomplete—suggests that individuals fell into the cave breaking their bones from the fall. : Evidence of carnivore damage is present on remains from Lawson Cave (Table 3); however, substantial gnawing results in density-mediated destruction of bone. The structure of low-density elements leads to their destruction by carnivores, thus low-density elements should be rare or absent in ravaged assemblages. Whether or not density-mediated destruction has occurred can be monitored by comparing the abundances of distal (dense) ends to those of the above (rela- tively less dense) ends of long bones (Binford 1981). Ratio values (R s) — the abundance of high and low density ends are calculated by determining t e MNE for the proximal end and for the distal end of each bone, and then eee all four values ([e.g.,] proximal humerus, distal humerus, weer ot b ista tibia) by the largest of the four values” (Lyman 1994a:400). Binford’s ( ) “zone of destruction’ and “zone of no destruction” in Figure 1 are derived from em- Pirical observation of carnivore ravaged and non-ravaged faunal assemblages (see WOLVERTON TABLE 3.—Carnivore damage on ursid remains. Vol. 21, No. 2 Element MNE Carnivore gnawed Zygomatics i 0 Mandibles 17 0 Scapulae 14 3 Humeri 16 10 Inae 13 6 Radii 11 3 Innominates 14 9 Femora 17 7 Tibiae 13 5 Total MNE 127 43 (33.9%) Lyman 1994a:398-402). Carnivore ravaged assemblages produce RVs that fall within the zone of destruction. Density mediated destruction of the Lawson Cave ursid and cottontail re- mains is monitored using tibia/tibiofibula and humerus RVs (Table 4). The re- sulting graph (Figure 1) illustrates that little or no density-mediated destruction has occurred; that is, low-density proximal tibiae and humeri occur at about the same frequency as high-density distal ends. Ursids undoubtedly temporarily sur- RV Prox. Tibia and Humerus FIGURE 1.—Destruction and humeri illustrate that low-density within or near Binford’s (1981) observed ’ ® Tibiofibula 0.8 4 0.6 4 Zone of No Destructio 0.4 4 0.2 4 Zone of Destruction 0 ee ore T ot 0 0.2 0.4 0.6 0.8 1 RV Distal Tibia and Humerus ————_ : | & Ursus americanus 8 Ai floridanus i gtaph: The ratio values of ursid and 80% of molar specimens are prime aged (Figure 4)]. Following Garcia et al. (1997), 83% of the M2s (MNE = 12) are from prime adults (Table 5). Graphic comparison to the presumed hibernation-death pattern from Yarimburgaz Cave (Stiner 1998) highlights the distinctiveness of the Lawson Cave pattern (Figure 4). Validity of the Lawson Cave Ursid Mortality Pattern.—Samples of the model ec lations yield no mortality patterns as prime-dominated as that from = S ave (Table 7). Eight of the fifty samples drawn from Population A were ‘og paateond inated (the samples contain more prime adults than expected from the mo vi population); two of those samples included 60% prime adults. The fifty samples from Population B, which consisted of more prime adults than on included ~ prime-dominated samples. Two of those samples comprised 60% . : . and one consisted of 70% prime adults. The fifty che semis : ine eg a eae d thirteen prime-dominated samples. One of those contained 60% prime : ults ha another contained 70% prime adults. Based on the samples reel sii ese model populations it is reasonable to conclude that eRe cage -— can be produced randomly from stable-age structure living ee a it appears unlikely that assemblages as heavily “igo eaan nae pe . ena Lawson Cave regularly result from random sampling of stable living eo sane Given these results it is more reasonable to conclude that the prime-dominate WOLVERTON Vol. 21, No. 2 60 N Individuals Ww & Model Living Populations| A = .? Pit 50% Juvenile 30% Prime 20% Old N Individuals Ww SS = = ra 50% Juvenile 35% Prime 15% Old N Individuals Ww a or L || 45% Juvenile 40% Prime 15% Old FIGURE 3.—Age structures of modeled living-structure populations. Winter 2001 JOURNAL OF ETHNOBIOLOGY 67 Old Adults 100% 2 Lawson Cave O Yarimburgaz Cave M " / a “Shaped L-Shaped ‘ N ay Prime Juveniles \& re 7 Adults \ / LM' 100% > ‘ oo 0% FIGURE 4.—Three-pole graph illustrating ursid mortality patterns from Yarimburgaz Cave (U. deningeri) and Lawson Cave (U. americanus). pattern at Lawson Cave is the result of physiological and related behavioral char- acteristics of young prime-adult bears. DISCUSSION One can argue that Lawson Cave and Sima de los Huesos acted as similar faunal accumulation agents based on cave structure alone; Lawson Cave is 11.5 m deep and Sima de los Huesos is 13 m deep (Arsuaga et al. 1997). The mortality patterns of these two assemblages are likely more similar than appears because of a minor recovery bias at Lawson Cave; smaller elements occur In lower nen expected frequencies in the collection (Wolverton 1996). Neonate and juvenile TABLE 7.—Results of model-population random sampling. Population Population Population A B . 50 samples (n = 10) drawn per population % prime-dominated samples drawn at random 1 ey is ean ae containing 60% any ee Uy Re tok) 6% 14% 26% Number of samples containing 70% — = 1(2%) n= 1 (2%) prime adults Number of samples containing 80% a ao a =f n=0 prime adults ee oe ee 68 WOLVERTON Vol. 21, No. 2 TABLE 8.—Condition of black bears in western Washington (Poelker and Hartwell, 1973). Health status Females (n = 12) Males (n = 13) M & F (n = 25) Poor to fair condition *Y = =1 Y=2 SA = SA = SA =5 A=2 A=1 A=3 Good to excellent condition Y = 1 (anestrus) Y=1 Y=2 SA = 1 (estrus) SA = 0 SA =1 A = 5 (2 estrus) A=7 A=12 *Y = Yearling, SA = Subadult, and A = Adult. teeth from friable mandibles and crania might not have been recovered because they were not recognized or they did not preserve. The presence of one neonate mandible (with an inset deciduous premolar) might lead to the inference that Lawson Cave served as a den; however, such an interpretation contradicts all indications, such as cave structure and taphonomic data, that Lawson Cave was a natural trap. Garcia et al. (1997) report no visible recovery bias in the Sima de los Huesos assemblage. Should the Sima de los Huesos data ever be converted to Stiner’s three phases, a more clear young-adult dominated pattern than that shown in Figure 2A might emerge because the two aging schemes operate on separate ordinal scales. As stated by Garcia et al. (1997:172) “the most likely scenario compatible with the structure of the Sima de los Huesos carnivore assemblage is a natural trap (very likely the current pitfall) attracting carnivores to accidental deaths.’ A sim- ple physiological analogy—implicating a carrion attractant—is useful to under- stand the proximate reason that ursids, given their keen senses of smell (Brown 1993; Schullery 1992), were attracted to both pits. Modern bear-bait trapping uti- lizes meat or carrion to draw ursids into barrels or other enclosures (Conover 1983; Craighead et al. 1995; McLaughlin and Smith 1990; Oliver 1995). Clearly U-shaped mortality patterns cannot be expected in natural-trap ursid assemblag- es. Why, then, do there appear to be proportionally more prime adults in the Lawson Cave and Sima de los Huesos assemblages than might be expected in a stable living population? Why wouldn't bears of all ages be equally susceptible to natural-trap deaths, which would result in L-shaped mortality patterns? Again, a modern analogy provides a plausible answer. There appear to be — relative abundances of young prime adults in the Lawson Cave assemblage; - can be argued for Sima de los Huesos based on the proportion of in- ivi uals in stages I and II. Most of the prime-adult-phase molars from Lawson ae fall in the two early, prime-adult cohorts (4 and 5). Sixty percent of the left se es and “hiss Alpe percent of the right molars were aged to cohorts four ase ra within the prime-adult phase (Table 5). It is arguable, then, that these aa aa sub-adults or young adults within the prime-adult phase. Pe aman due artwell (1973:121) demonstrate that Washington-state subadult paren ose individuals no longer with their mother and in the process of pone pe a. ranges (Powell et al. 1997)—are the least healthy of all age pn : retin was gauged in terms of disease and parasite loads and twell 1973) *s neg only one subadult was considered healthy (Poelker and Har- - Yowell et al. (1997) discuss two limiting resources that shape beat a ne Winter 2001 JOURNAL OF ETHNOBIOLOGY 69 home ranges: food and prospective mates [see Craighead et al. 1995 for discussion related to grizzly bears (U. arctos)]. Adult male black bears tend to have large home ranges and do little immediate sharing of resources with conspecifics (Pow- ell et al. 1997; see also Beecham and Rohlman 1994; Boileau et al. 1994; Craighead et al. 1995; Klenner 1987; Klenner and Kroeker 1990). If home ranges overlap among males it is because the areas are large, not because of cooperation. Male home ranges only tend to shift in response to movement of potential mates, but not in response to food availability. Adult females use overlapping home ranges that change relative to food availability (Powell et al. 1997). Within this matrix are young prime-adult bears establishing home ranges. It is likely that access to both limiting resources (food and mates) is unpredictable; thus, young prime adults are less healthy (e.g., more in need of food) than older prime-aged adults with established ranges. For example, among 56 black bears studied by Garshelis and Hellgren (1994:180) in Minnesota, the relatively young males tended to “be underrepresented as breeders. However, wounds incurred from aggressive encounters with other bears’’ were common. Prime-aged males with established home ranges tended to have higher serum-testosterone levels early in the breeding season—they had early access to mates. McLellan et al. (1999:917) report that young male grizzly bears (U. arctos) in the Pacific Northwest have higher mortality rates than well-established adult bears; “perhaps due to their large ranges and inexperience, young males are more prone to encounter human attractants and be killed as problem bears than [members of] other sex- age classes.” Adult males and females with established home ranges have better access to preferred food resources and mates; as a result they have lower mortality rates. Given the argument presented here, subadult and young adult black bears are under greater nutritional stress than adult bears; they lack access to limiting resources (food and mates). It follows that subadult and young adult bears are susceptible to carrion attractants in natural traps. The apparent preponderance of young prime-adults in the Lawson Cave assemblage supports this notion because young-adult bears undergo considerable stress during their attempts to establish home ranges (Garshelis and Hellgren 1994; McLellan et al. 1999; Powell et al. 1997). CONCLUSIONS the results here are best cast in the form As the title to this paper suggests, et a d mortality. This is so for two reasons: of a natural-trap hypothesis regarding ursi see 1) only two natural-trap assemblages are examined here using di et (es methods, and 2) one of those assemblages (Lawson Cave) is small. Nevert —— the high proportional abundance of young adult ursids in , i natura trap assemblages is markedly distinct from their low proportional abun weg in win- ter-death, U-shaped mortality profiles. Further, the documented natural-trap mor- tality patterns contrast with those expec vulnerable, hibernating bears. There is a P sids are attracted to natural traps; in particular ceptible to death in natural traps. hysiological/ behavioral reason that ur- young-adult ursids are most sus- 70 WOLVERTON Vol. 21, No. 2 Ursid mortality data from sites such as Sima de los Huesos provide another line of evidence with which to understand accumulation histories of palimpsest assemblages, such as those from caves—whether archaeological, paleontological, or mixed. An important component of the argument presented here is that the Lawson Cave assemblage is a non-cultural assemblage; it can be used to ferret out expected characteristics of remains deposited via natural entrapment. In ae ticular, mortality data can be useful for understanding accumulation histories 0 assemblages that contain artifacts/human remains and ursid remains. This is particularly relevant for faunal assemblages from caves with multiple entrances. ACKNOWLEDGMENTS I thank R. Lee Lyman, Mary C. Stiner, Travis Pickering, Mitchell Sullivan, Sergio Her- rera, Alejandra Gudifio, Michelle Drapeau, and Blaine Schubert for assistance with parts of this paper and anonymous reviewers of earlier drafts for constructive comments. Pre- viously unreported radiocarbon dates were funded by the National Science Foundation (grant # SBR-9912118) LITERATURE CITED ABEL, O. and G. KYRLE, eds. 1931. Die Drachenhéhle bei Mixnitz. Spelaolog. Monogr. Vols. 7-8. Vienna. ARSUAGA, J. L, I. MARTINEZ, A. GRA- CIA, J. M. CARRETERO, C. LORENZO, N. GARCIA, and A. I. ORTEGA. 1997. Sima de los Huesos (Sierra de Atapuer- ca, Spain): The site. Journal of Human _ Evolution 33:109-127. BACHLER, E. 1940. Das Alpine Palaoli- thikum der Schweiz. Monographien zur Ur- und Friihgeschichte der Schweiz. Basel. ARE 3 fee 8 Altersfliederung der Hohlenbarenreste im Wildkirchli, Wil- denmannisloch un Drachenloch. Quar- tar 9:131-146. BEECHAM, J. J. and J. ROHLMAN. 1994. A Shadow in the Forest: Idaho's Black Bear. University of Idaho Press, Mos- IW. cow. BEGON, M., M. MORTIMER, and D. 5 - 1995. Population Ecology. Blackwell Science Ltd., Oxford. BINFORD, L. R. 1981. Bones: Ancient Men and tape Myths. Academic Press, ew BOCHERENS, H., M. FIZET, and A. MAR- IOTTI. 1994. Diet, physiology and ecol- en isotope bio- geochemistry: implications for Pleisto- cene bears. Pa aeogeography, Palaeocli- matology, Palaeoecology 107:213-225, BOILEAU, E, M. CRETE, and J. HUOT. 1994. Food habits of the black bear, Uur- sus americanus, and habitat use in Gas- pésie Park, Eastern Québec. The Cana- dian Field-Naturalist 108:162-169. BROWN, G. 1993. The Great Bear Almanac. Lyons and Burford Publishers, New ork. CONOVER, A. 1983. Getting to know black bears—right on their own home ground. Smithsonian 14:87-96. CRAIGHEAD, J. J., J. S. SUMNER, and J. A. MITCHELL. 1995. The Grizzly Bears of i, GARCIA, N., J. L. ARSUAGA, and T. TOR- RES. 1997. The carnivore remains from the Sima de los Huesos Middle Pleis- tocene site (Sierra de Atapuerca, Spain). Journal of Human Evolution 33:15 174 GARGETT, R. H. 1996. Cave bears and modern human origins: the spatial ta- phonomy of Pod Hradem Cave, Czech Republic. University Press of America, Lanham, MD. GARSHELIS, D. L. and E. C. HELLGREN. 1994. Variation in reproductive biology of male black bears. Journal of Mam- malogy 75:75-188. t GORDON, K. R. 1977. Molar measuremen ; as a taxonomic tool in Ursus. Journal 0 Mammalogy 58:247-248. Perreriemuinrersessssintomeieapecr Winter 2001 . 1986. Insular valaniseiegy body size trends in Ursus. Journal of Mam- malogy 67:395-399. and G. V. MOREJOHN. 1975. Sex- ing ‘black bear skulls using lower canine measurements: Journal of Wildlife Man- agement GRAHAM, R. W. 1991. Variability in the Bill Neff Cave, Virginia. Illinois State useum Scientific Papers 23:238-250 GRANDAL-D’ANGLADE, A. and J. R. VI- DAL-ROMANI. 1996. A population study on the cave bear (Ursus spelaeus Ros.-Hein.) from Cova Eirés (Triacas- 08 Galicia, Spain). Geobios 30:723- ot KLENNER, W. 1987. Seasonal movements and home range utilization Str . the black bear, Ursus americanu estern Manitoba. The Canadian Field. Naturalist 101:558-568. , and D. W. KROEKER. 1990. Den- ning behavior of black bears, Ursus americanus, in Western Manitoba. The Canadian Field-Naturalist 104:540-544. KOBY, FE. E. 1953. Les Paléolithiques ont-ils Chassé L’ours des Cavernes? Actes Jurass. Emul. 1954:1-48. KURTEN, B. 1958. Life and death of the Pleistocene cave bear: a study in paleo- re Acta Zoologica Fennica 95:4- . 1976. The Cave Bear Story. Colum- bia University Press, New York. LYMAN, R. L. 1987. On the analysis of ver- tebrate mortality profiles: sample size, mortality type, and hunting ean ae American Antiquity 52:125- ——.. 1994a. Vertebrate Saree Te University Press, Cam- ri ge. —.. 1994b. Relative abundances of skel- etal specimens and taphonomic analy- sis of vertebrate remains. Palaios 9:288- 298. . 1994c. Quantitative units and ter- minology in zooarchaeology. American Antiquity 59:36-71. MARKS, S. A. and A. W. ERICKSON. 1966. Age determination in the black bear. Journal of Wildlife Management 30: 389-410 McLALGHLIN, C. R. and H. L. SMITH. JOURNAL OF ETHNOBIOLOGY 71 1990. Baiting black bears: hunting tech- niques and management issues. Eastern Workshop on Black Bear Research and Management 10:110-119 McLELLAN, B., FE. W. HOVEY, R. D. MACE, S, D. W. CARNEY, M. L. GI- BEAU, W. L. WAKKINEN, and W. F KASWORM. 1999. reat and causes of grizzly bear mortality in the interior mountains of British Gokabie: Alber- ta, Montana, Washington, and Idaho. opened of Wildlife Management 63: 911-92 NELSON, D E., A. ANGERBORN, K. LI- DEN, and I. TURK. 1998. Stable iso- topes and the metabolism of the Euro- pean cave bear. Oecologia 116:177-181. OLIVER, D. K. 1995. The bears of August. Arizona Wildlife News July:2-4. OLIVER, J. S. 1989. Analogues and site con- text: bone damages from Shield Trap Cav neem Carbon County, Mon- Pp. 73-98 in Bone Modifi- vation oR. ocisachater at d M. H. Sorg, (editors). Center for the Study of the First Americans, Oreno, ME. POELKER, R. J. and H. D. HARTWELL. 1973. Black Bear of Washington. Wash- ington State Game Department, Biolog- ical Bulletin, No. 14. POWELL, R. A., J. W. ZIMMERMAN, and D. E. SEAMAN. 1997. Ecology and Be- havior of North American Black Bears: Home Ranges, Habitat and Social Or- ganization. Wildlife papal and Be- ries, No. 4: Chapman and Hall, London RAUSCH, R. L. 1961. Notes on the black bear, Ursus americanus, Pallas, in Alas- ka, with particular reference to denti- tion and growth. Zeitschrift fur Saiige- tierkunde 26: oo ¢ SCHULLERY, P. 1992. The Bears of Yellow- stone. High oe Publishing Compa- ny, aaa MITH, _ K. A. STROTHER, J. C. ROSE, dl M. SAVELLE. 1994. Chem- ical ultrastructure of cementum growth-layers of teeth of black bears. Journal of Mammalogy 75:406—409. STINER, M. C. 1990. The use of mortality patterns in archaeological studies of hominid predatory adaptations. Journal of es Sa Archaeology 9:305- ___ 4994, Honor Among Thieves: A 72 WOLVERTON Zooarchaeological Study of Neandertal Princeton University Press, . 1998. Mortality analysis of Pleisto- cene bears and its paleoanthropological relevance. Journal of Human Evolution 34:303-326. ——, S. L. KUHN, S. WEINER, and O. BAR-YOSEE. 1995. Differential burning, recrystalization, and fragmentation of archaeological bone. Journal of Archae- ological Science 22:223-237. , G. ARSEBUK, and FE. C. HOWELL. 1996. Cave bears and Paleolithic arti- facts in Yarimburgaz Cave, Turkey: dis- 112707 a palimpsest. Geoarchaeology 279-3 ——, H ACHYUTHAN. G. ARSEBUK, FE C. HOWELL, S. JOSEPHSON, K. JUELL, J. PIGATI, and J. QUADE. 1998. Reconstructing cave bear paleoecology from skeletons: a cross-disciplinary study of Middle Pleistocene bears from Vol. 21, No. 2 Yarimburgaz Cave, Turkey. Paleobiolo- gy 24:74-98. TUCKER, T. G. 1984. A study of Quater- nary black bears (Ursus americanus) from Missouri, with special reference to the extinct subspecies, Ursus americanus amplidens Leidy. Missouri Speleology 24:36-51. WEBB, R. E. 1988. Interpreting the faunal debris found in central European sites occupied by neandertals. Pp. 79-104 in Recent Developments in Environmental Analysis in Old World and New World Archaeology, R. E. Webb (editor). Bar International Series, No. 416. WOLVERTON, S. 1996. Morphometry and taphonomy of the Lawson Cave ursids. M.A. thesis (Anthropology), University of Missouri-Columbia, Columbia , and R. L. LYMAN. 1998. Measuring late Quaternary ursid diminution in the Midwest Quaternary Research 49:322- 329. Winter 2001 JOURNAL OF ETHNOBIOLOGY 73 People, Plants, and Justice: The Politics of Nature Conservation. Charles Zerner, ed. New York: Columbia University Press. 2000. Pp. 416. $49.50 (cloth). People, Plants, and Justice: The Politics of Nature Conservation makes three major contributions debates on the ethics of nature conservation. First, the theoretical and methodological approaches presented by the contributing authors to this vol- ume advance political ecology scholarship. Second, the book suggests alternative models, principles, and perspectives that, if adopted, will enable conservation organizations to improve rather than abuse human rights. Third, it boldly exposes potential and actual human rights abuses caused by conservation projects by cri- tiquing specific guiding models, principles, and perspectives. The data that are presented in this book support the proposition that the conservation of biodiver- sity does not necessarily coincide with the protection of human rights. In other words, the objectives of conservationists are not always the same as the objectives of indigenous and/or local people. It is possible to reach common ground, how- ever, if conservationists can become less ethnocentric, learn from past mistakes, and—as is repeatedly emphasized throughout the book—relinquish control to local people. The contributors to People, Plants, and Justice contextualize conservation pro- grams in the political economic struggles that characterize the contemporary world. Their substantial evidence includes data on historical and contemporary social relations involving natural resource management in Africa, Latin America, Oceania, and Southeast Asia. Many of the authors juxtapose the perceptions and practices of local communities to those of conservation organizations, examining social relations between actors who are internal and external to environments that house valuable resources. Rather than romanticizing indigenous peoples or de- monizing the multiple other groups of actors who are subjects of analysis, the book presents a non-essentialized, empirically-based analysis of the social rela- tions of conservation. Nonetheless, either because of the realities of our contem- porary world or the bias of the authors, the “scales of justice” (Zerner 2000:17) weigh heavily against conservation organizations. on The book is divided into two parts. Part One, “ACTOS the atin consists of three chapters that define the subject matter, establish the book s approach, set the prevailing tone for the writings, and review other chapters in the book (some- times applaudingly and other times harshly). The on of stage quite unique among edited volumes since three chapters an = . Oo 2 form tasks that are typically accomplished in just one chapter. ar oe 3 & mere summaries of the chapters in Part Iwo, “On Location, Aa ‘ : co chapters in Part One combine critical commentary with rich data from the aut veal own research in insightful comparisons. In the official Snneepeal ‘peti e “Toward a Broader Vision of Justice and Nature Conservation ) ake es ae the editor, leads the reader through the volume’s major Issues of sos ce ion formations, community dynamics, culturally constructed images 0 omy a commodification and global circulation of nature, and alee a se pled source management regimes. Zerner challenges his reat oO i ham critical research, design ‘‘better’’ social-ecological-political-econo : and re-align the political economy with human rights. In Chapter 1, ‘Contested 74 BOOK REVIEWS Vol. 21, No. 2 Communities, Malignant Markets, and Gilded Governance: Justice, Resource Ex- tractions, and Conservation in the Tropics,” Michael Watts uses his review of the case studies in Part Two to scrutinize ‘community’ as a concept that is trendy and instrumental yet unsound and even dangerous in terms of increasing the risk of social injustices. In Chapter 2, ‘Beyond Distributive Justice: Resource Extraction and Environmental Justice in the Tropics,” Richard Schroeder questions the stan- dard theories of conservationists suggesting that they are culturally-bound, mar- ket-driven, and insufficient agendas and challenging the merits of ‘distributive justice,” a common model for sustainable development that the World Wildlife Fund and other organizations follow in their conservation projects. Together Schroeder, Watts, and Zerner preface the prevailing temper of all the book’s con- tributors by simultaneously deconstructing and expanding—through the insertion of democracy, cultural relativity, and local control—conceptions of conservation. Part Two consists of 13 case studies that produce mixed emotions in the reader because they are at once revealing, shocking, discouraging, and inspiring. The authors confront us with the oppressive potential of imaginations including our own conceptualizations of local communities and those of institutions such as the state, development organizations, and conservation projects. For example, in Chapter 3, “Justice for Whom? Contemporary Images of Amazonia,” Candace Slater describes typical American views (shared by many of the people who will read her article) of the Amazon and Amazonian residents. As she describes the historical development of our conceptualizations, Slater gives us contradictory demographic and ethnographic evidence to demonstrate that the reality of the Amazon does not coincide with the fantasy. The reader should beware: this article may cause painful self-reflections. Another illustration of the danger of inaccurate imaginations is found in Chapter 9, “Global Markets, Local Injustice in Southeast Asian Seas: The Live Fish Trade and Local Fishers in the Togean Islands of Sulawesi,” by Celia Lowe. Lowe criticizes the routine paradigm that is used in designing regulatory policies that identifies local people as the cause of unsustainable resource use. This is a restricted, hence inaccurate / incomplete, conceptualization of environmental deg- radation because it does not consider the influence of external political economic forces, or the innumerous local-global links in commodity chains. woro, and Patrice Lew, the Sustainability of | rattan production in Winter 2001 JOURNAL OF ETHNOBIOLOGY 75 policies of Indonesia deny the validity of indigenous agroforestry, in the process de-legitimizing the basis of local property rights. The official ignorance of the state enables the appropriation of vast amounts of acreage and the establishment of less-rational forms of commodity production that are not based on local en- vironmental knowledge. While the objective of conservation organizations is to solve ecological prob- lems, they often cause social problems. In Chapter 13, “A Tale of Two Villages: Culture, Conservation, and Ecocolonialism in Samoa,” Paul Cox compares the operational procedures of a democratic, community-controlled successful envi- ronmental project in one Samoan village to a top-heavy, externally-designed un- successful environmental project in another Samoan village. Cox goes so far as to label ethnocentric conservation projects that are insensitive to indigenous per- spectives “ecocolonialism.” To avoid contributing to global imperialism, Cox (2000:343) suggests three aspects that ought to be incorporated into environmental projects in indigenous communities: ‘“Consent of the indigenous people, respect for their culture, and submission to indigenous political control.” Jill Belsky docu- ments the ways that conservation programs in central Belize exacerbated social and ecological dysfunction. In Chapter 11, “The Meaning of the Manatee: An Examination of Community-Based Ecotourism Discourse and Practice im Gales Point, Belize,’ Belsky deconstructs the “community-based conservation model and ecotourism as a solution to ecological degradation. ; Although the three chapters in Part One are all reviews of the 13 chapters in Part Two, there is little if any redundancy. Zerner, Watts, and Schroeder notice different themes, contextualize the case studies in variant yet overlapping bodies of scholarship, and take off from them in personalistic directions. Moreover, or- ganizing the volume so that there are three introductory chapters In the beginning eliminates the need to have a concluding chapter following the case studies. Thus, there is no summary chapter at the end of Part Two. Instead, the book abe a case study that Zerner (2000:9) describes as an “analytical tour de — ms eed, the audience enjoys a fiery finalé as Bronwyn Parry guides us epee : cou of plant collecting providing a chronology of the changing eed enh ss beginning with the early era of exploration, when the value of botanicals ss based on novelty and exotic-ness, through the current bio-techno = in whi botanical value is determined by efficiency in communicating knowle a . This book speaks to environmental and social advocates, policy ma si an scholars. It is a call to action. Conservationists should not read this book as ae attack on their views and goals. Instead, they should use this esas as a sane for becoming more culturally aware of the particular geograp + ee . groups, and natural resources with whom they work. Through 28 ~ a conser- vationists can consider the growing literature on the sae rela ou “aed vation as attempts to improve, not dismantle, their pore © "f as a ii nice to combine the need for environmental protection with socia es oe x also can use this volume to improve upon their work in a sel us ve dea cal instance, the writings in this volume enhance reflexivity, pagel ccsial ques- concepts, pose important questions, and provide ans th models tions as well. Most importantly, the book’s articles suppty Ss 76 BOOK REVIEWS Vol. 21, No. 2 for conducting more ethical research and tools for improving social conditions cross-culturally. Since the publication of People, Plants, and Justice: The Politics of Nature Conser- vation, it will never again be possible for the assortment of people who manage plants—ranging from herbarium collectors to biotechnicians—to claim that their endeavors are benign. As suggested by the subtitle of this book, their activities are embedded in global politics. Participants in nature conservation—from con- sumers of “‘rainforest” candybars to ecotourists—can no longer assume that their learned perceptions of ‘other’ ecosystems or their contributions to “save” the en- vironment are cross-culturally true or socially just. Cynthia T. Fowler Assistant Researcher Division of Ecology and Health John A. Burns School of Medicine University of Hawaii Honolulu, Hawaii 96822 LITERATURE CITED ZERNER, CHARLES (editor). 2000. People, Plants, and Justice: The Politics of Nature Con- servation, Charles Zerner (editor). Columbia University Press, New York. - 2000. Toward a broader vision of justice and nature conservation. Pp. 3-20 in Peo- ple, Plants, and Justice: The Politics of Nature Conservation, Charles Zerner (editor). olumbia University Press, New York. COX, PAUL. 2000. A tale of two villages: culture, conservation, and ecocolonialism in Sa- moa. Pp. 330-344 in People, Plants, and Justice: The Politics of Nature Conservation, Charles Zerner (editor). Columbia University Press, New York. Journal of Ethnobiology 21(2): 77-90 Winter 2001 THE USE OF CATTAIL (Typha latifolia L.) DOWN AS A SACRED SUBSTANCE BY THE INTERIOR AND COAST SALISH OF BRITISH COLUMBIA JOANNA OSTAPKOWICZ Liverpool Museum, William Brown St. Liverpool, UK L3 8EN DANA LEPOFSKY Department of Archaeology Simon Fraser University Burnaby, B.C, Canada V5A 186 RICK SCHULTING School of Archaeology and Paleontology Queens University Belfast Belfast, UK BT7 1NN ALBERT (SONNY) McHALSIE 1-1201 Vedder Road St6:lo Nation Chilliwack, B.C., Canada V2R 4G5 ABSTRACT.—The economic uses of plants are often more ible to researchers working with actual material remains from early ethnographic and archaeological sources than are ritual uses. Nevertheless, it is clear from the ethnographic liter- ature of the Northwest of North America that plants also served many important ritual and ceremonial functions. During the examination of two Salish wooden mortuary figures currently housed at the Museum of Anthropology, University of British Columbia, a compact, fibrous white mass was observed lodged in the back of the mouth of one of the figures. A sample of the material was identified the ritual uses of cattail down, particu- larly with regards to funerary customs, among the Coast and Interior Salish of the Northwest of North America. Key words: cattail, Salish, mortuary rituals. RESUMEN.—Las aplicaciones econdmicas de las plantas siguen siendo, a menudo, mas accesibles a los investigadores que trabajan con material real de fuentes eth- s aplicaciones rituales. Sin embargo, nogrdaphicas y arqueolégicas tempranas que la mi esta claro que en la literatura ethnographica del noroeste de Norte América las Durante la examinacién de dos escultura tenidas actualmente en el Museo de la Antropologia, 78 OSTAPKOWICZ et al. Vol. 21, No. 2 Britanica, se observ6 una masa blanca fibrosa alojada en la parte posterior de la boca de una de las figuras. Una muestra del material fue identificada como Typha latifolia L. Este papel pone el resultado en contexto, en que discute las aplicaciones rituales de la pelusa de la espadafia, particularmente con respecto a las costum- bres funerarias, entre las populaciones Salish de la costa y el interior de la costa noroeste de Norte América. RESUME.—U utilisation économique des végétaux est souvent plus accessible que leur usage rituel aux chercheurs travaillant sur les restes matériels provenant de sources ethnohistoriques ou archéologiques. La littérature ethnographique con- cernant le Nord-Ouest de l'Amérique du Nord montre néanmoins clairement que les plantes ont également eu de nombreuses fonctions rituelles et cérémonielles. Au cours de l’examen de deux figurines mortuaires en bois, actuellement conser- vées au Musée d’Anthropologie de l'Université de Colombie Britannique, une mas- se blanche, compacte et fibreuse, fut observée a l’arriére de la cavité buccale d’une des deux figurines. Un échantillon de cette substance a été identifié comme Typha latifolia L. Le but du présent article est de re-situer cette trouvaille dans son con- texte, en discutant des usages rituels de chaton, particuligrement dans le cadre de coutumes funéraires, chez les populations Salish du littoral et de l’intérieur des terres du Nord-Ouest de l‘Amérique du Nord. INTRODUCTION Though there is some record of the ritual uses of plants in the Northwest (Turner 1982; Compton 1991), the record for their economic uses is relatively more complete (e.g. Compton 1993; Turner 1995, 1997, 1998). This is in part due to the fact that though First Nations people may have described rituals in general terms for the early ethnographers, there was a reluctance on some occasions to share knowledge about the rituals associated with specific plants. This would have been particularly true for knowledge that was owned and guarded by individual households. Later in the historic era, when ceremonial life was disrupted by dras- tic depopulation (Boyd 1990, Carlson 1997a) and the performing of traditional ceremonies was suppressed or prohibited outright (Carlson 1997b; Cole and Chai- kin 1990; Fisher 1992), some details about the ritual roles of individual plants were lost. __ The identification of plants used to make ritual artifacts or those found in ritually important contexts (cf. Carlson 1999) is an avenue for understanding cer- emonial uses of plants in the past. In particular, the identification of such plants provides information on cultural Prescriptions for the appropriate plant for spe- cific ritual contexts. Such information, in turn, provides a broader understanding of traditional ceremonial life and of the larger worldview, and may furthermore suggest new lines of interpretation and investigation. fot liye ae Nba auarin's the identification of cattail (Typha latifolia L.) down vie outh of one of a pair of Salish wooden mortuary figures.’ We begin 4 brief overview of the Salish, followed by an account of the figures that provides the context for the cattail down. A review of the ethnographic and eth- nobotanical information for the Interior and Coastal Salish reveals that cattail served a variety of economic needs, but was also an important element in several Winter 2001 JOURNAL OF ETHNOBIOLOGY 79 “ Chilcotin—..” Sqtatl'imx i : pper Lillooet) “Squam %, Lilwat ish’ - (Lower Lillooet) - Bost , ton Bar : FRASERCANYON — { Nicola anes Yale ; Sto:lo ® Sardis Cee an bee oe a lor ae eee hoe fs ee 2 WA F nay 1.—Map of southwestern British Columbia, showing locations of selected Salish groups and places discussed in the text. ga of Salish ritual life. In particular, cattail down was strongly associated onre traditional funerary rites among the Salish. The ritual significance of the Scan, ey be in part associated with the symbolic importance of the color white e worldview of the Coast and Interior Salish. THE SALISH In British Columbia, traditional Salish territory extends across much of the southern part of the province (Figure 1). The most basic division of this territory is that between the Coast and Interior Salish, reflecting 4 major language division as well as cultural differences. According to late nineteenth century and early twentieth century ethnographic and ethnohistoric information, Salish social and economic organization was based on several of which ne up a household. On the coast, family units lived in large shed-roof plank ouses, while in the interior, smaller plank houses and semi-subterranean pit- houses were used. In some areas, villages were quite large, with several hundred inhabitants. The subsistence economy was based on the collection and manage- fy mals, root foods, and berries. Political authority was largely hereditary and invested in the heads of high rank- Ing families, especially among coastal groups. In the interior, social and political organization was more flexible, although still showing a strong hereditary com- ponent (Barnett 1955; Teit 1900, 1906). Both Coast and Interior Salish societies were semi-sedentary, with highly complex material culture and ceremonial life 80 OSTAPKOWICZ et al. Vol. 21, No. 2 FIGURE 2a.—Salish mortuary figure (UBC Museum of Anthropology, A1780). FIGURE 2b.—Close up of the cattail down in the mouth of the mortuary figure (Photo- graphs by J. Ostapkowicz). based largely around the acquisition of personal spirit power (Kew 1990; Suttles 1987, 1990a). THE GRAVE FIGURES The grave figure with the cattail down in its mouth is one of a pair of carvings currently housed at the Museum of Anthropology, University of British Columbia provenience, and thus specific group affiliation, is uncertain. The museum attri- 0—the Coast Salish of the central and upper Fraser Valley— however, others have attributed the figures to the N’lakapamux (Thompson), an Interior Salish group? (Figure 1). . Freestanding, fully sculptural depictions of the human figure, such as the pair under discussion, were typical of Coast and Interior Salish mortuary art. Mor- tuary figures—depicting men, women, and sxwayxwey® dancers—were carved as oe Winter 2001 JOURNAL OF ETHNOBIOLOGY 81 representatives of the deceased, and were erected for commemorative purposes at Salish grave sites. The practice of erecting these figures in front of family grave houses and box burials spanned much of Coast and Interior Salish territory from at least the beginning of the nineteenth century to the early twentieth century in British Columbia (Ostapkowicz in press). Upwards of eighty figures are known from museum collections and archival photographs (Ostapkowicz 1992), although the one under discussion here is the only example known to be associated with cattail down. Serving as memorials, the figures were painted and dressed in the deceased’s clothing (Teit 1906:273). Those responsible for commissioning the carving of the figures were likely the heads of families of high status. Teit (1900:330), in writing about the N’lakapamux, commented: ‘The Indians state that the only reasons for placing these figures near graves were to keep the dead relative fresh in the memory of the living; to show that the person respected the dead relative; and to let people know who was buried there, and that the dead had living relatives who were above the common people as to wealth and able always to renew the clothes of the figure.” The prominence of the figures and the ceremonies surrounding their erection and subsequent reclothing, were a means through which the living ex- pressed their status, wealth, and close link to their ancestors. The mortuary figures discussed here display marked similarities to one an- other and are clearly intended as a matched pair, perhaps carved by the same artist. The larger of the two figures contains the cattail down in the back of its mouth (Figure 2b). This figure is 168 cm in height and according to the museum accession records is carved from cedar (probably western red cedar, Thuja plicata Donn). Strips of leather have been nailed to the head and groin area. Only traces of white paint are visible today on the chin and cheeks, but red and black pigment were observed on the figure some decades ago (Wingert 1949:136). These figures are distinguished from most other examples of Salish mortuary art by their unusual facial carving, which invokes the Tal mask. Like the facial features of the figures, Tal masks are characterized by large, deeply sunken cheeks and eyes, an open, down-turned mouth, and bent nose. Such masks were repre- sentations of a legendary female giant (a Coast Salish version of the Kwakwa- ka’wakw Tsonoqua, an ‘ogress’ who was also the provider of great wealth), and their ownership was a hereditary privilege as well as a mark of wealth and pres- tige (Barnett 1955:170-171; Lévi-Strauss 1988:66). The masks were used during winter dances and life crises rites, including commemorative ceremonies. Barnett (1955:236) notes that the “... appearance of a Tal mask at a ceremony honouring a deceased father signalized the transference of that mask to his heir. Effigies of the dead were made for these ceremonies and their faces covered with the mask. om expand upon the significance of the association of the Tal and the cattail down elow. ECONOMIC USES OF CATTAIL AMONG THE SALISH Common cattail is a perennial that thrives in shallow marshes, ponds, wet ditches, and lakeshores. The familiar ‘cat's tail/—the brown, velvety spike located at the tip of the main stem—bears the flowers which turn into a white, cottony 82 OSTAPKOWICZ et al. Vol. 21, No. 2 fluff in the late summer and fall. The plant is harvested for its leaves in late summer, and then left to air dry (Turner and Efrat 1982:58; Turner 1998:121-123). The seed down found in the figure’s mouth was likely collected in the late sum- mer/early fall and may have been used immediately or stored for future use. Though the rootstock and pollen of cattail were collected for food by several Interior Salish groups (Turner et al 1990; Parish et al. 1996; Turner 1997), the plant's leaves were most highly valued on the coast and the interior as weaving material (Steedman 1930:496; Turner and Bell 1971; Pojar and MacKinnon 1994; Turner 1988, 1998; Turner et al. 1990). Indeed, among the Island Salish, cattail is considered “... probably the most important basket and mat weaving material” (Turner and Bell 1971:77). Baskets, bags, clothing, twine, cradles, nets, canoe sails, and mats were woven from the leaves and stems (see Teit 1900:188—190, for an overview of the weaving process). Woven cattail mats, for example, were used in various ways, such as for wall insulators and temporary summer shelters (Turner et al. 1990:145; Turner 1998:122-123). Cattail weavings would also be used as clothing (cloaks, robes, hats, headdresses) and would occasionally be combined with dog hair for added warmth (Barnett 1955; Curtis 1970; Turner and Bell 1971: 77; Turner 1988). Teit (1900:256) also notes the use of rafts made of cattail bundles among the Nicola. Based on these utilitarian uses, Turner (1988) ranks the plant in the ‘High Significance’ category for the Lillooet in her Index of Cultural Sig- nificance (ICS). Elder Rosaleen George notes that cattail has the same significance to the St6:l0 of the Fraser Valley as the cedar tree (pers. comm. to A. McHalsie). Cattail down, because it was absorbent and soft, also served a variety of everyday needs. In particular, the Coast and Interior Salish used the down as stuffing for pillows, mattresses, for wound dressing, and for infant diapers (Steed- man 1930:498; Pojar and MacKinnon 1994:338; Parish et al. 1996:359; Turner 1998: 123). Cattail down was also woven into mountain goat wool blankets—a point we will return to below. RITUAL USES OF CATTAIL AMONG THE SALISH Although cattail had several mundane uses, it also served more esoteric pur poses. Among the Saanich, of Vancouver Island, cattail charcoal was used for tattooing (Jenness in Turner and Bell 1971; Turner 1998:123), a practice reserved for the wealthy (Barnett 1955:74). Tattooing and face and body painting were also practiced by the Interior Salish N’lakapamux (Teit 1930). The Songish, again of Vancouver Island, offered a mixture of burned cattail root with Lomatium sp. and red paint in First Salmon Rites (Turner and Bell 1971:77). Among the N’lakapamux, cattail leaves were incorporated into shamans’ headdresses (Turner et. al. 1990:145), and the stalks were used to weave burial shrouds in the Nicola Valley (Smith 1900:405). There is a particularly strong association between cattail down and burial pieces ~ Coast and Interior Salish. Hill-Tout (1905:137) writes that among the ieee Se he | Lillooet), “[t]he body was customarily washed all over, the hair pease ed back, the face painted, and the head sprinkled with the down se - es [cattails], which was potent in checking the evil influences attending tpses."* This was done by a special funerary shaman, immune to the dangers Winter 2001 JOURNAL OF ETHNOBIOLOGY 83 involved in dealing with the corpse. Among the Chilliwack, a St6:16 group of the central Fraser Valley, Hill-Tout (1978:54) noted that, “‘After the body of the dead person has been taken from the house the ‘olia’ [‘the soothsayer’] would take quantities of the down of bulrushes [cattails] and spread it all over the bed on which the deceased had lain.” The connection between cattail down and the dead is further demonstrated in the protohistoric burial of an infant found near the modern town of Yale, at the northern boundary of traditional St6:16 territory and the southern limit of N'‘lakapamux territory. The infant had been interred inside a copper trade pot which, together with the other copper grave offerings, led to remarkable preser- vation conditions resulting in the preservation of soft tissues and plant fibers. Among the plant fibres was a downy white material which had been placed, together with red ochre, around the infant's anterior fontanelle. Red ochre was also placed inside the infant's mouth. As elsewhere, red ochre is a sacred sub- stance among the Salish, and is often found in burial contexts among the Coast and Interior Salish (Schulting 1995). The white material has been examined mi- croscopically and is consistent with cattail down (Schulting 1992), although the absence of attached seeds precludes a definite identification.® If the material is indeed cattail down, this and the fluff in the mortuary figure, represent the only known examples of ritual use of cattail down outside of ethnographic sources. The association of the ochre and the down with the head of the infant is significant given the spiritual importance of the head in Northwest Coast societies (e.g., Cybulski 1978). In Northwest Coast rock art, for instance, the head is almost always larger and more detailed than representations of the body (Lundy 1983) and modification to the head, through head deformation, facial tattooing, and the Wearing of labrets were used to mark membership in social groups yong 1990b). Specifically, Barnett (1955:221-222) notes that among the Coast Salish, 4 3 soul was “taken to be the vital quality of the heart or head ve and makes teference to a Saanich shaman retrieving a lost soul and placing sliabooe the pa- tient’s head. That similar concepts prevailed among the Interior Salish te casi from Teit’s (1900:363) comment concerning the N’lakapamux belief that t ets was supposed to leave the body through the frontal fontanelle. It is age i Suggest, then, a scenario in which the spirit of the eS e ceed through the fontanelle, was purified by passing through materials su ae ld down and ochre. The placement of the red ochre in the Yale infant's ce ns be viewed similarly, since this is where the breath—or life igo ag eatgae vd and may provide another parallel to the placement of cattail down in of the mortuary figure. The ritual importance of cattail is further Particular places which are considered sacred Halkomelem place name Xatsuq’ (Xaxa, sacred, | for a lake in the Fraser Valley (Hatzic Lake) which cattail. The association of cattail and — . tad 8eneral to the fact that cattail is used in sacred co : are dan- at least two other locations within Sté:16 a _ Zearet nit used by 8erous and off-limits to those who are spiritually unprepa sia chenlind cay Indian doctors on spirit power quests), suggests pat sone Ps : gn : th highlighted by its connection wit ore clearly illustrated by the 84 OSTAPKOWICZ et al. Vol. 21, No. 2 be sacred (Keith Carlson, pers. comm. to D. Lepofsky, 2000). Cattail collected from such locations may have residual spiritual power in them, and may have been the source of down which was used in mortuary and other rituals, while cattail des- tined for more prosaic uses could have been gathered from other, less dangerous, locations. DISCUSSION Among the Coast and Interior Salish, purification, or the ‘cleansing’ of the deceased was an important aspect of mortuary rituals. Sxwayxwey dancers, for example, would be called in to ‘wash’ the corpse (Barnett 1955:217), and normally a year after burial, they would again be hired to ‘wash’ the mortuary figure (Barnett 1955:220). Barnett (1955:217) notes that the cleansing rituals involving the sxwayxwey and surrounding the burial “... did not differ from that employed for ‘washing’ a pubescent, a newly named adult, an infant, a dancer initiate, or any other individual assuming a new social position.’’ Occasionally, figures bear- ing the sxwayxwey mask would be permanently erected at the burial sites of families who had rights to the masks—a long term, public affirmation of the family’s good standing and their accordance with the proper ritual observances. The use of skowmidgeons—supernatural creatures most akin to fishers (a large member of the weasel family)—was another important aspect of mortuary cleans- ing rights, and representations of these creatures would often appear on mortuary figures and posts (Ostapkowicz 1992).6 __ The various Salish groups considered several plants to be important in cleans- ing ceremonies and used them in rituals surrounding an individual’s death. Ac- cording to Hill-Tout (1978:34), Squamish and Lillooet “... burnt cedar (Thuya gigantea) [stet] as well as salal-berry (Gaultheria shallon) branches and whip the whole dwelling with boughs, particularly that part where the body lay, to drive away the presence of death, sickness and ghosts, all of which are supposed to linger there.” Spruce boughs (Picea sp.) are placed both at the head and under the bed of the husband or wife of the deceased as a protective measure against Sickness and death, and food is eaten off these boughs for a month after the funeral (Hill-Tout 1978:35). Among the Coast Salish, the body of each participant ina bereavement ceremony was cleansed by smoking branches, while those who were in direct contact with the body (undertaker, coffin-maker, pallbearer) washed with various herbs after the completion of the ceremony (Barnett 1955: 219). The N’lakapamux also used Douglas-fir (Pseudotsuga menziesii [Mirbel] Fran- co) during rituals for the bereaved (Turner et al. 1990:58). In sum, various plants were vital to the fulfillment of a number of important, highly ritualised events, and through the associated actions natural materials were transformed into spir- Though the Salish used sever to have held particular importan in several aspects of mortua deceased had lain prior to b in preparation for burial, an al plants in funerary rituals, cattail down seems ce. It is now apparent that the down was used ry rituals: it was strewn over the place where the urial, it was sprinkled on the head of the deceased d it was placed in the mouth of grave figures. The Winter 2001 JOURNAL OF ETHNOBIOLOGY 85 recurrent association of down with the head, a focal point of the human body, mind, and spirit in Salish belief systems underscores its ritual value. Another significance of the cattail down in this particular case is seen in its conspicuous placement in the mouth of a Tal. Both the Tal and the down are instruments of cleansing, yet the mask would be ineffective in cleansing rituals without the use of what Suttles (1987:104) calls the ‘ritual word’. Indeed, the ritual word was at the heart of cleansing rites, charging the instruments used during these ceremonies with efficacy. Wearers of the Tal masks presumably had asso- ciated power songs, or specific ritual words, that were private and used only during important events. Hence, the placement of the cattail down in the mouth of the mortuary figure may be interpreted as emphasising the power of the spo- ken word. The underlying theme linking cattail down and concepts of death, the after- world and spiritual cleansing may be the symbolic potency of the colour white. Tepper (1994:75), in outlining the importance of colour among the N’lakapamux, points out that different colours are associated with “. . . abstract concepts usually linked to a system of religious beliefs’. Teit (1930:419) recognized the symbolic importance of colour among the N’lakapamux, and drew attention to the white as a ‘spirit’ colour, linked to “. . . ghost, spirit world, dead people, skeletons, bones, sickness, coming from the dead”. The white down of cattail, associated as it is with burials and burial figures, hints at such a symbolic association.” In addition to its links with the ritual aspects of death, the colour symbolism of white has overarching associations with status and spirituality. In Salish society, items made of white wool were often highly valued elite and ceremonial objects. For example, white blankets made from mountain goat wool or wool shorn from dogs actively bred for their white pelage, were highly treasured items (Schulting 1994). Cattail down was sometimes woven into these blankets as well (Barnett corporated into shamans’ or ritua ing rituals. Initiates of certain secret socie ing mountain goat wool (see Kew 1990: Fig. 1), sae? : i their new names (see Suttles 1990a: Fig. 10). Ritualists attending ceremonies would weave wool into their hair (Barnett 1955:153), eclusion following her first menses (Barnett covered with white swan feathers and down.* ee Cattail down seems to differ from the other white, sacred gee On = it was relatively abundant and easy to acquire. However, we know a ‘ g aris collection of cattail that was intended for ritual purposes. Je as 7 ee at ° material may have been restricted to ritual specialists, collec : : aia Particular ways, possibly only from specific stands _ sete . ee titual use (such as Xatsuq’). Alternatively, any cattail dow y it i d substance. Propriate, and only its inclusion in rituals transformed it into a sacted' s 86 OSTAPKOWICZ et al. Vol. 21, No. 2 Redcedar boughs used in ritual cleansings might be another example of a com- mon plant that is transformed during ritual performances. Materials such as white cattail down were, at least in part, visual indicators of the status of the ritual practitioner (mediating between the spirits and the realm of the living) or of the supernaturally vulnerable initiate or patient. The link be- tween spirituality and higher moral, social, and economic status is a prominent feature of Salish society (Hayden and Schulting 1997; Suttles 1987). For the Salish, as with many cultures, the ability to out-perform ordinary community members in the observances of what are regarded as the proper rituals both confers and justifies the high standing of certain families (cf. Owens and Hayden 1997). In the case of cattail down, the connection with the elite was made in several ways: it was part of a larger ceremony that included the carving and erection of a large mortuary figure, the hereditary right to carve a Tal mask (with its wealth con- notations), the clothing (and periodic reclothing) of the figure, and presumably a relatively elaborate graveside ritual. down marked and helped to facilitate the change. It was a means of cleansing the individual, thereby preparing their spirit for the journey ahead; in addition, it helped to protect the living from the uncontrolled and dangerous influences at- tendant upon the corpse. This example of an economic/ symbolic dichotomy is not unique for (or to) the Salish. Many plants, in many societies, present the same complex relation- ea Our task is to try to understand and appreciate both ways of viewing the world. NOTES ' The identification was based on the morphology of the fluff as well as the attached seeds. In an account by J.S. Matthews—the founder of the Vancouver Archives and an archi- vist there until 1970—the figures were found at a grave site between Boston Bar and Lytton, Winter 2001 JOURNAL OF ETHNOBIOLOGY 87 couver], and set them out to photograph before returning them back to the shed. This would indicate that the figures may not have been in the Raley collection at this time, and suggests yet another version of events. The University of British Columbia Museum purchased the mortuary figures in 1948 (UBC Museum accession records). >The term sxwayxwey has come to refer to both the characteristic Salish mask with pro- truding cylindrical eyes as well as the associated dances and ceremonies that feature this mask. Different names are recorded for various sxwayxwey masks (see Suttles 1987:109- 111), which are distinguished by additions of bird or animal heads in addition to the frequently seen round collar and/or a crest of feathers. Such masks and dances function as instruments of cleansing. ‘The common name “bulrush” is often used interchangeably with “cattail”. True bulrush (Scirpus sp.) does not produce a fluffy seed head. *The morphology of the down alone is insufficient to distinguish among the many species of plants which produce seed fluff (Cathy D’Andrea, pers. comm.). ° Jenness (1934:73) notes the myth related to skowmidgeons, and how they had the power to ‘wash away the tears’ of the bereaved: “Later Khaals changed some sarees hn group into fishers, and said to Seleeptim: ‘These animals will comfort you in orenyee to come. They shall be your ¢xwte'n, a solace to drive away your tears. When a chil a or some dear kinsman, you shall kill two, four, or even six fishers, dry their ais ee store them in safety. Then you shall utter the prayer that I will now teach you, ana they shall wash away your tears.” ‘The white berries of snowberry (Symphoricarpos albus L.) are often sae ripe . dead. In several languages on the coast and interior the berries are given names like a4 : berry’. For instance, the berries are referred to as ‘the saskatoon berries of the people 0 the Land of the Dead’ in one Stl’atl’imx story (Pojar and MacKinnon 1994:70). ate eagle or swan down into var- 8 : Numerous northern Northwest Coast peoples ale aed wi dewn'and when thie ious ceremonies. During certain dances, the headdress i ; dancer tilts his head, hss aes the down to fall to the ground. Holm vegetated Notes that the down incorporated into such headdresses was = shaken a dene and over y sharp movements of the dancer’s head, then swirled and drifted aroun ar covered the assembled watchers. Following headdress dances, the floor of the house Pie cnloe with drifts of white down’. Again, there is the connection between the sacred, white, and the head. ACKNOWLEDGMENTS : i infant burial, We appreciate Cathy D’Andrea’s efforts to identify the down found in the in i ‘cal. Many thanks also to Deeley oss Cin Coon ad 0 jepson et their insightful 88 OSTAPKOWICZ et al. Vol. 21, No. 2 LITERATURE CITED BARNETT, HOMER G. 1955. The Coast Sa- lish of British Columbia. University Press, Eugene, Oregon. BOYD, ROBERT T. 1990. Demographic his- tory, 1774-1874. Pp. 135-148 in Hand- book of North American Indians, vol. 7, Northwest Coast. Wayne Suttles (edi- tor). Smithsonian Institution, Washing- ton, D.C. CARLSON, KEITH, T. 1997a. Early encoun- ters. Pp. 27-40 in You are Asked to Wit- ness: the St6:l6 in Canada’s Pacific Coast History, Keith T. Carlson (editor). St6:16 Heritage Trust, Chilliwack, Brit- ish Columbia. . 1997b. Early nineteenth century Sto:lo social structures and government assimilation policy. Pp. 87-108 in You are Asked to Witness: The St6:l6 in Canada’s Pacific Coast History, Keith T. Carlson (editor). St6:10 Heritage Trust, Chilliwack, British Columbia. CARLSON, ROY L. 1999. Sacred sites on the Northwest Coast of North America. Pp. 39-46 in Bog Bodies, Sacred Sites Jorgensen (editors). WARP Occasional Paper, Exeter, UK. COLE, DOUGLAS and IRA CHAIKIN. 1 Intyre and University of Washington Press, Vancouver and Seattle. COMPTON, BRIAN. 1991. ‘It pulls every- thing to you’: North Wakashan herbal talismans. Pp. 33-70 in Collected Papers of the 26th International Conference on Salish and Neighbouring Languages. University of British Columbia, Vancou- ver. - Ph.D. dissertation (Botany), University of British Columbia, Vancou- Vv er. URTIS, EDWARD §. 1970 [originally pub- lished 1913]. The North American In- dian, Vol. 9. Johnson Reprint Co., New ork. CYBULSKI, JEROME S. 1978. Modified hu- man bones and skulls from Prince Ru- pert Harbour, British Columbia. Cana- dian Journal of Archaeology 2:15-32. FISHER, ROBIN. 1992. Contact and Con- flict. UBC Press, Vancouver. GUSTAFSON, PAULA. 1980. Coast Salish Weaving. University of Washington Press, Seattle. HAYDEN, BRIAN and RICK J. SCHULT Sphere and Late Prehistoric cultural complexity. American Antiquity 62(1): 51-8 1-85. HILL-TOUT, CHARLES. 1905. Report on the ethnology of the StlatlumH of Brit- ish Columbia. Journal of the Royal An- thropological Institute 35:126-218. ——.. 1978 [Originally published 1902]. The Salish People: the local contribu- tions of Charles Hill-Tout, Vol. I, The Squamish and the Lillooet. Talonbooks, ncouver. HOLM, BILL. 1990. Soft Gold: The Fur Trade and Cultural Exchange on the Northwest Coast of America, 2nd edi- tion. Oregon Historical Society Press, Portland. JENNESS, DIAMOND. 1934. The Indians of Canada. Anthropological Series 15. Na- tional Museum of Canada Bulletin 65, LEVL-STRAUSS, CLAUDE. 1988 [Original ly published 1975]. The Way of r Masks. Douglas and Mcintyre, Vancou: ver. LUNDY, DORIS. 1983. Styles of coastal rock art. Pp. 88-97 in Indian Art Traditions of the Northwest Coast, Roy Carlson (editor). Archaeology Press, Simon Fra Visible Ghosts: the human figure in a lish Mortuary Art. B.A. honors thes Winter 2001 (Archaeology). Simon Fraser University, Burnaby. : ress. The human figure in In- terior Salish mortuary art. In Burial Practices of the Interior Salish and their comparative study of transegalitarian hunter-gatherers. Journal of Anthropo- logical Archaeology 16:121-161. PARISH, ROBERTA, RAY COUPE, and DENNIS LLOYD. 1996. Plants of South- ern Interior British Columbia. Lone Pine Press, Vancouver. POJAR, JIM and ANDY MACKINNON. 1994. Plants of Coastal British Colum- bia. Lone Pine Publishing, Vancouver. SCHULTING, RICK J. 1992. Preservation of soft tissues by copper in the Interior Plateau of British Columbia, Canada. Paper presented at the First Internation- al Congress on Mummy Studies, Puerto de la Cruz, Tenerife. ———-. 1994. Hair of the Dog: the identi- fication of a Coast Salish dog-hair blan- ket from Yale, British Columbia. Cana- Fraser Plateau. Archaeology Press, Si- mon Fraser University, Burnaby. SMITH, HARLAN I. 1900. Archaeology of um of Natural History, Vol. 1, Part VI, New York. STEEDMAN, ELSIE V. (editor). 1930. The Ethnobotany of the Thompson Indians of British Columbia, based on Field Otes by James A. Teit. Pp. 447-522 in Bureau of American Ethnology, 45th am Report, 1927-28. Washington, SUTTLES, WAYNE. 1987. Productivity and its Constraints: a Coast Salish Case. Pp. 100-133 in Coast Salish Essays. Univer- isidecanorsa 1990a. Central Coast Salish. Pp. Oast, Wayne Suttles (editor). Smith- Se Institution Press, Washington, JOURNAL OF ETHNOBIOLOGY 89 SUTTLES, WAYNE. 1990b. Introduction. . 1-15 in The Handbook of North American Indians, vol. 7, Northwest Coast, Wayne Suttles (editor). Smith- sonian Institution Press, Washington, D rag TEIT, JAMES A. 1900. The Thompson In- dians of British Columbia. Memoirs of the American Museum of Natural His- tory 2(4):163-392. New York. 1906. The Lillooet Indians. Mem- oirs of the American Museum of Nat- ural History 2(5):195-292. New York. —_—. 1930. Tattooing and face and body painting of the Thompson Indians, Brit- ish Columbia. Pp. 397-439 in Bureau of American Ethnology, 45th Annual Re- port, 1927-28. Washington, D.C TEPPER, LESLIE H. 1994. Earth Line and Morning Star: Nlaka’pamux Clothing Traditions. Canadian Museum of Civi- lization, Ottawa. TURNER, NANCY J. 1982. Traditional use of Devil's Club (Oplopanax horridus) by native peoples in western North Amer- ica. Journal of Ethnobiology 2:1:17-38. _____ 1988. “The Importance of a Rose”: Evaluating the cultural significance of plants in Thompson and Lillooet Inte- tior Salish. American Anthropologist 90:272-290. _ 1995. Food Plants of Coastal First Peoples. UBC Press, Vancouver. " 1997. Food Plants of Interior First Peoples. UBC Press, Vancouver. _____ 1998. Plants in Technology of First Nations of British Columbia. UBC Press, Vancouver. _ and BARBARA S. EFRAT. 1982. Ethnobotany of the Hesquiat People of the West Coast of Vancouver Island. British Columbia Provincial Museum, Cultural Recovery Papers 2, Victoria. _and M. A. M. BELL. 1971. The Eth- nobotany of the Coast Salish Indians of Vancouver Island. Economic Botany 27: 57-310. A N THOMAS, BARRY F. CARL- SON and ROBERT T. OGILVIE. 1983. the Nitinaht Indians of Vancouver Island, British Columbia Provincial Museum Occasional Papers, No. 24, British Columbia Provincial Mu- seum, Victoria. Ethnobotany of 90 OSTAPKOWICZ et al. Vol. 21, No. 2 , LAWRENCE, C. THOMPSON, M. Royal British Columbia Museum, Vic- TERRY THOMPSON and ANNIE Z. oria. YORK. 1990. Thompson Ethnobotany: | WINGERT, PAUL S. 1949. American Indian Knowledge and Usage of Plants by the Sculpture. J. J. Augustin Publisher, New Thompson Indians of British Columbia. York. Journal of Ethnobiology 21(2): 91-127 Winter 2001 CONTRIBUTIONS TO THE ETHNOBOTANY OF THE CUP’IT ESKIMO, NUNIVAK ISLAND, ALASKA DENNIS GRIFFIN Archaeological Frontiers 295 East 33rd Eugene, Oregon 97405 ABSTRACT.—Ethnobotanical information on the Native use of 47 species of in- digenous plants on Nunivak Island, Alaska is presented. Changes in subsistence use among the Cup’it Eskimo of Nunivak, throughout the twentieth century, have resulted in the loss of traditional ethnobotanical knowledge. While previous stud- ies have presented limited information on the importance of particular plant spe- cies to the local diet, additional data regarding the role of indigenous plants and subsequent changes in plant use have recently been recorded. They are discussed here in light of the adoption of western foods and medicines and increased contact of the Cup’it with mainland peoples. Current knowledge of traditional plant use and the importance of plants to local dietary, medicinal and utilitarian uses are summarized. Key words: ethnobotany, Cup’it Eskimo, indigenous plant use, Nunivak Island, Alaska RESUMEN.—Presentacién de datos ethnobotanicos de 47 especies de plantas in- digenas y las maneras de uso por la gente indigena de la Isla Nunivak en el estado de Alaska. Los cambios en los usos de estas plantas para la subsistencia por los Esquimales Cup’it de Nunivak a través del siglo XX han ocasionado la pérdida de conocimientos ethnobotanicos tradicionales. Mientras que los estudios anteriores han presentado datos limitados sobre la importancia de ciertas especies de plantas comestibles en la dieta local, recientemente se ha documentado ~ macién adicional respeto al papel de las plantas indigenas y los cambios en si usos de éstas. Esta nueva informacién se discute en este trabajo en vista de la adoptacién de alimentos y medicinas occidentales cial entre los Cup’it y los habitantes del continente. : los usos Basticion kes de estas plantas nativas y su importancia en Y ore ri asi que los usos utilitarios y medicinales de éstas, S¢ resumen en este trabajo. RESUME.—Des informations éthno-botaniques sut pomene e —~ _ rmis les Esquimaux Cup’it 92 GRIFFIN Vol. 21, No. 2 indigénes traditionelles et l’importance des plantes dans les usages locaux quant au régime alimentaire et 4 la médecine et dans d’autres emplois utilitaires sont présentées ici en résumé. INTRODUCTION The Yukon-Kuskokwim Delta, a geographic and cultural area historically oc- cupied by Central-Alaskan-Yup’ik speaking Eskimos in southwestern Alaska, en- compasses an area of almost 81 million kilometers (31,250 square miles) or 8.1 million hectares (20 million acres). This region consists of a vast and largely road- less expanse of low lying tundra that has attracted limited attention from eth- nographers in the past. Native villages are located along the area’s major water- ways with development largely limited to commercial fishing. The degree of con- tact between subcultural groups within the Delta cannot accurately be determined due to conflicting early historic data and later movements of peoples throughout the region, but villages are known to have been linked by extensive trade net- works, intermarriage among village residents, and village alliances during times of warfare (VanStone 1984:224). Knowledge of the Native use of indigenous flora in the Yukon-Kuskokwim Delta remains quite limited. Early ethnobotanical stud- ies in the region are limited to research on Nunivak Island (Fries 1977; Lantis 1946, 1959), Nelson Island (Ager and Ager 1980) and the village of Napaskiak (Oswalt 1957) located along the Kuskokwim River (see Figure 1). Nunivak Island, located approximately 37 kilometers (23 miles) west of the Alaskan mainland and 209 kilometers (130 miles) west of Bethel, the largest town in the Delta, has tra- ditionally remained the most isolated area in southwestern Alaska. Nunivak is the only major off-shore island inhabited by Central-Yup’ik speaking people, the Cup’it! or Nunivarrmiut (VanStone 1989), who maintained their isolation until after World War II when an airstrip linked the island to the mainland. The present study summarizes the known traditional use of indigenous plants on Nunivak Island in addition to changes in plant use during the twentieth century, and provides comparisons of plant use with that of mainland Yukon-Kuskokwim Es- kimo peoples. This information was obtained from Cup’it elders during a four year (ca. 1995-1998) collaborative anthropological project between the author and the community of Mekoryuk. Community members participated in all facets of the project, including archaeological excavations, oral interviews and artifact and plant identification, and were monetarily reimbursed for sharing their expertise. REGIONAL SETTING : Nunivak Island is located in the Bering Sea off the western coast of Alaska etween 165°30’ and 167°30' West longitude and 59°45’ and 60°30’ North latitude. raphy of the island is highly diverse. The west coast cliffs, reaching over 122 meters (400 feet) in elevation, Winter 2001 JOURNAL OF ETHNOBIOLOGY 93 Bering Sea Bering Sea oH Nash Harbor Mebosyel 3 25 scale in miles FIGURE 1.—Map of Yukon-Kuskokwim Delta showing villages discussed in text. Which provide a spectacular bird sanctuary for southern coastline contains miles of sand beache north and east coastlines are comprised of relatively tocky beaches and numerous coves and protective contains an upland plateau-like area rising in elevation from 152 to 244 meters above sea level (498 to 800 feet), culminating in a mountainous area of volcanic origin. The lowland areas are generally well-watered and contain numerous lakes and ponds, while the mountainous areas have fewer lakes and ponds although Most of the larger lakes are located within this latter region. Nunivak inna is subject to a Subarctic maritime climate, influenced by the Surrounding sea which produces a relatively stable temperature. soe ae are generally cool and windy, with some areas experiencing frequent fog; winters are cold with both wet and dry periods. The island’s mean annual temperature ; - Centigrade (C) (20°F) with mean daily temperatures ranging from —25 94 GRIFFIN Vol. 21, No. 2 (—13° F) in January and February to 10° C (49.9° F) in August (Swanson et al. 1986). Rain and snowfall is heavier than on the adjacent mainland, resulting in frequently overcast days with dense fogs. This difference from the mainland delta regions is due to the greater effect of the Bering Sea on the island environment. Precipitation is moderate with a mean annual rainfall of 40.6 cm (16 inches) and snowfall of 137 cm (54 inches). The present flora of Nunivak has been intensively studied by Bos (1967), who built upon the earlier work of Palmer and Rouse (1945). The island’s vegetation is predominantly comprised of Arctic tundra containing a variety of lichens, grasses, sedges, flowers, and shrubs. It is similar to coastal and coastal-upland vegetation found throughout western and northwestern Alaska. The tallest island plants are shrubby willows which can reach up to eight feet in height along some of the island's river courses. Major vegetational types (Figure 2) are comprised of wet tundra, dry tundra, and grass-browse (i.e., grass hummock and beach grass- forb). Wet tundra covers approximately 57% of the island and is most prevalent on the north side of the island between the villages of Mekoryuk and Nash Har- bor, extending southward. Dry tundra covers most of the interior portions of Nunivak (13.6%) and includes two recognized subtypes: dry tundra found on areas of sloping terrain having good drainage, and alpine tundra found at higher elevations on hills and mountains. Grass-browse covers approximately 23.4% of the island and is found interspersed with the dry tundra subtype and along edges of streams and rivers adapted to periodic flooding. PREVIOUS ETHNOBOTANICAL RESEARCH Previous investigations of the Native use of Nunivak Island flora are limited to the works of Margaret Lantis and Janet Fries. Margaret Lantis spent a year on Nunivak (ca. 1939-1940) studying the social dynamics of the Cup’ it people (Lantis 1946), with subsequent research efforts focusing on the development of children, local genealogies, the psycho-dynamics of Cup’it society, and community politics. A brief summary of local plant use was later published by Lantis (1959) along with comparisons to the Native use of plants throughout Alaska. In 1977, Janet Fries (1977) completed a senior honor’s paper on the vascular flora of Nunivak which addressed the flora she found to be in current use at the time of her study. My investigation of the use and importance of island flora stems from my 1995- 1998 Ph.D. anthropology research on Nunivak where I was able to work closely with Cup’it elders from the village of Mekoryuk, the only village remaining on Nunivak, and build upon these earlier studies (Griffin 1999). While my research focus was based on reconstructing changes in Native lifeways over time at the village of Nash Harbor, located approximately 43 kilometers (27 miles) west of Mekoryuk, I was also able to discuss traditional use of indigenous plants with island elders. This paper presents a summary of Cup’it plant use derived from elder interviews both in their homes and during collecting activities. Indigenous plants were an integral part of the year-round diet of Eskimo people in addition to their incorporation in other facets of their life. Contrary to the popular perception of Eskimo people surviving solely on fish and meat, the Cup’it utilized a large number of local plants for food, medicinal, and utilitarian Winter 2001 JOURNAL OF ETHNOBIOLOGY 95 Cape Etolin Bering Sea Se. Cape Manning Cape Mohican, ee ¥ aA fy) = \ * “g rd RS oo Rac | / NUNIVAK ISLAND Etolin Strait ue ct ey Le \ * Cape Corwin i a S Bering Sea 2 BGF = Legend 15? ” Cape Mendenhall | Wet Tundra 0 15 DT Dry Tundra Grass ‘ : BR scale in miles eer tsteieeereeeeemeateeeeeeeeeeeees FIGURE 2.—Distribution of major vegetation types (adapted from Bos 1967). Purposes. An earlier Alaskan study estimated that up to 15% of the diet of West- em Eskimo people (Kotzebue to Alaska Peninsula) is made up of vegetable re- sources (Young and Hall 1969:43). While plant resources remained sparse on some off-shore islands such as St. Lawrence Island (Young and Hall 1969), on Nunivak they Provided a significant addition to the Cup’it’s year-round diet. Table 1 pro- Vides a list of the seasonal use of indigenous plants by the Cup’it- A complete list of all utilized species (including subspecies, variations and synonyms), au- wad scientific names and voucher specimen numbers is included in the ix. Appendix NATIVE PLANT TAXONOMY f 4 dictionary of Cup’ig terms and their roots is in draft form and an analysis Y son Toot systems is not yet possible. However, an examination of general “Pik terms (Jacobson 1984) provides comparative data useful in distinguishing Mints basic plant terminology distinctions among the Cup’tt. Yup’ik speakers (in- Ww “id ~ Cup’it) tend to divide plants into basic groups based on how plants % traditionally used, their similarity in appearance or physical characteristics. Ny tail the Cup’ig plant name ciwassit translates to ‘wild greens that can : ey and is used to denote several distinct species that are prepared ina ni ar manner (i.e., Rumex arcticus (sour dock), Polygonum bistorta (bistort) sm rhlans) e (alpine bistort). Kumarutet is used to denote all moss species (eg. Pohlia light). ased on the traditional use of moss as a wick in lamps (kuman ' amp. 1) elg Xamples of plants grouped by similarity in appearance or setting include: Mat—term used to designate several varieties of seaweed (e.g., Palmaria pal- Vol. 21, No. 2 GRIFFIN sale M ‘1S Svjnyvsy 4vsyw Ausaqpnoy> snaowmavuiavys snqry SOLLIOG ns ansvavinnd Arraquoo3en, snj1jo4v snqny SUId}S ‘SOARI] ns ‘Ss dno1ayyng seed usvjjvd snjnounuvy awioziys ns ‘S qissvm19 _qieqnuyy PIM, 403s1g ourdpy uinavdiaia wnuoshjog SIARI] ng ‘a qussvm19 sawnyg YUIg ‘}10}s1g vj40}81q UinuosAjog jueyd S JajnAvUINy sso| suvjnu vIpYyog sIaMOy S yomasno’y AJOOM VID]PIINJ4A2 SlAvINIIPad SdALI] M ‘nS ‘S qibnsui _Alapea /a3eqqe> PpIlM,, synonpnu vhs yueyd M 4 ‘1S qunbya asinq “paameras awuyod viAvUl]vg SdALI] M ‘nS ‘S 4vysynnb [aL10g uTeyUNOPy vuhdsip m4hxoO SaLlieq 4 ‘ns Arraqueiy 30g sndavoodU SNIIONRXE SIARD] é qwodanyaaunig yeapia}sAQ IULIPLAVUL VISUALAaV SUId}s ‘SAARA] ‘S}OOI ns ‘S qoasavynia ‘qnhiv,yny _Agsieg /drusieg plim,, ‘e8eao7] yoeag MINI1JOIS UNIISNSTT SAARI] ns ‘S qn fiv vay, Jopeiqe’y] aajsnjod winpaT SUI9}S ‘SBALI] M ‘ns ‘S qwsayjnyny WOMpues UpRaqeas ‘sudaI5) yoRag sapiojdad vhuayouopy SUI9}S ‘SBARI| M4’S jnavvhvy [rey Soreyy SLAVS]NA *H io vpjhydvsjay standdipy ued M4 ‘NS ‘S qonbyja soRimsiappelg ‘dds snonq wa}s JO aseq ans sauyad ssei3u0}07 [[eL unyofiysnsuv unaoydoiaq SAARI] ns paaMally wnuyofiysnSuv wunigopidz Sal1Iaq M4 saunvd ‘ywsunvd AaqMory uinaSiu wnajaduz spuoyj 4 ‘NS aavobji9 ua PPTs vyoywyip stsajdohaq SAAR2] M ‘NS ‘S wbnsur ,P20}2'] PIIM,, vasoquadhy “q JO stjvasog vqvuiq 3001 ns ‘Ss Aa]sieg-YIOTWAP{ U9}SA asUuaUIy? WiNUIasoluo> wo) S jidjn 98304 PIIM,, Ajneag-3utids snosaqny, vsosagny miuojshvj> juejd arjue S susypr'] dds miuopyjD ways ‘jO01 d jasauyad sa3pas dds xaiv> juejd argue M‘S gnjaim pjosueyy ysiepy suujsnjud vyyjv> saLliaq 4 ‘ns Svjavy Ausaqieag auidry vuidjv sojhydvjsojaay yIeIS ‘SAARI] ns ‘S dnd Yt} Al] PIM, vpion] voyasuy syue][g pooy yied jury uOsRaS saueu 31.dn> saureu UOUTUIO> GURU ITWUAIIS ‘purys] YearuNN uo sjyueyd jeuDtpaul pue pooy snouaSrpul jo asn yeuoseas— | FIGVL 97 JOURNAL OF ETHNOBIOLOGY Winter 2001 SolIaq 4S = ywywyrpSvuny 4y8vu0H4 Aizequeiy ysng-Mo'] ‘Arraquosury] VAVPI-S1J1A2. WUNIUIIIUA, SARI] 4 ‘ns yissvant9 _’peutds piim,, oq ros SNINJIAD XIUNY yueyd argue qns’s jajnavuiny sso SUBINU VITYOd sule}s q ‘ns ’s JadUuay [re}JesIOFY UOUTUIO asuansy unjasinby SIAL] ans‘s sseiry aAQ PTIM syjou snuh,y yueyd Mans ’s SSOP] JoapuTay ‘SudYoT] puLlafIsuvd vIUOpY]> SARI] 4 jasauyad sa3pas ‘dds xasv> umouyun é Sausajasa pooysyuUop] wuintyofiurydjap wnju0ay syue[d pesf) UeneW IA SARI] S qoibau ywysau do1399U0}45 “JOO.18SOY vasod Wnpas SIARB] q ‘ns ‘s qonsnhsnb MOT[IM Jee;puoweiqg, vaygnd x1vs SUD}O ‘SIARB] rns 's quavavysnunb MOTIIM, suaosaosnf X1VS SaLLeq q ‘ns Svjnyvsyy 4vsqv Aazaqpnoy> SnAOWAaVULDYD SNqny SUI9}S ‘SAARZT A ‘ns ‘s yn fiv vay, Jopeiqe’] aajsnyod uinpay SIMO] nes qajnibjam ‘nbsnu ssei3u0}0> dds wnasoydoaq SdARI] ns poamalty uinijofigsnsuv wnigopidy spuoly ime's ADYANJUII usa preys vooiajsny sisajdohiq SdARI] é yorlg JreMq syjixa vjnjag SdARI] q ‘ns ‘s POOMULION ‘PaaMyxUyS Issa]l] VISUAL syue[d [PUDIpey SdLLIoq M4‘S — yoqwy1psvuny aysvuHy Aaiaqueia ysng-mo7 ‘Arrequoosurqy VAVPI-S1J10. LANIUIIIYA SaLlaq ns qwaana Auragang 80g ‘Arragentg autdry UINSOULSI]N UNIULIIVA SaLLaq ns Snjpsvsyv YIRIS peystm |, snyofixajdu sndojdasys SU9}S ‘SAARI] M ‘nS auegealy ‘IOMSey vaudy-opnasd o12aUuas SIOMOT} S quoibau qwysau. do199u0}5 ‘JOO1aSOY wuinasod UNpas SdARI] S qossynnb SASPIJIXeS dds vSvsfixus SAARI] ‘SUDO nee qonsnhsnb MOTIM Jeaypuouleicg] vaygnd x1j1vS suDy}eo ns:‘s qunsnhsnb MOTIIM PYSETV SISUAXYIY XI]VS SUIB}S ‘SOARZT M ‘nS ‘S jissvani9 ypo0g Anos *ypo0q SNIYIAVD XAUNY yied jueyg uOSseas soureu 31,dn>5 saureu UOUTUIO7) aureUu SYWUaIIS (panuquod) | equ. 98 GRIFFIN Vol. 21, No. 2 mata [dulse], Fucus spp. [bladderwrack]), and 2) agyam an’a(i)—used for all puff- ball species (Lycoperdon spp. and Calvatia spp.). In Yup’ik, agyam ana translates to meteor and meteors are traditionally said to turn into puffballs when they land (Jacobson 1984:48). Still other plant names highlight distinctions within a genus such as qugyuguat which is used to refer to all Salix (willow) species except those exhibiting catkins which are referred to as gimugkararat. Further analysis is need- ed in order to fully understand the Cup’it’s concept and categorization of local flora. A similarity of plant use and some Native plant names between the Cup’it of Nunivak Island and the Inuit on the Seward Peninsula to the north were iden- tified during the study. The Eskimo linguistic branch consists of two clearly dif- ferentiated sub-groups, Yup’ik and Inuit-Inupiaq (Woodbury 1984). Yup‘ik was spoken aboriginally on the coast of the Chuckchi Peninsula in Siberia and in Alaska from Norton Sound south to the Alaska Peninsula and east to Prince William Sound. The Cup’it speak a sub-dialect of Yup’ik known locally as Cup’ig which is the most divergent dialect within the Yup’ik branch. The Inuit of North- ern Alaska and Canada speak Inupiaq which is spoken by Inuit peoples from the Seward Peninsula in Alaska across Arctic Canada. Similarities between some Cup’ig, Yup’ik and Inupiaq plant names (e.g., kavlag—kavlak—kavlag [Arcto- staphylos alpina], paunrat—paunraq—paungag [empetrum nigrum], pekner—pek- neq—pikneq (Eriophorum angustifolium]) and food preparations (e.g., akutar—ak- utaq—akutug [Eskimo ice-cream comprised of berries, seal oil, reindeer tallow (Crisco), snow and sometimes salmon eggs]) highlight extended contact between western Alaskan peoples over time. Further research is needed to evaluate the degree of sharing between these language branches with regard to the recognition and use of indigenous plants. PLANT HARVEST, PREPARATION AND STORAGE On Nunivak, most indigenous plants were traditionally gathered by women and children when the men were harvesting other available resources (e.g., cari- bou, waterfowl, seal) (Della Boesche, personal communication September 1995; Lantis 1946). While fresh spring greens provided a welcome addition to the diet, which in winter was based largely on dried and stored foods, other greens were harvested throughout the year as they ripened, and used with some of those stored for winter use. With the melting of the island’s snow pack, local greens and berries not picked during the previous fall’s harvest, begin to appear and were added to the local diet. Depending on the time the ice pack began to break- up, Cup’it families would leave their winter villages and move to spring seal camps. Cup‘it men would journey out along the ice to harvest arriving sea mam- mals (i.e., seals, walrus) while the women would spend much of their time har- vesting available plant resources (greens and seaweeds) and shellfish. Early spring plants included: marsh marigold (Caltha palustris), sour dock (Rumex arcticus), wild celery (Angelica lucida), wild lettuce (Draba borealis or D. hyperborea), wild parsnip (Ligusticum Hultenit), wild rhubarb (Polygonum viviparum), mountain sorrel (Oxyria digyna), Pallas buttercup (Ranunculus pallasii), and Labrador tea (Ledum palustre decumbens). Winter 2001 JOURNAL OF ETHNOBIOLOGY 99 After the completion of the hunting season, families would move to summer fish camps. Fish comprised the most prolific and essential subsistence resource for many Alaskan Natives living in the Yukon-Kuskokwim Delta region and its harvest would occupy the majority of the families’ efforts for several months. Traditional plants would continue to be harvested as they ripened and were eaten fresh or placed in underground caches for temporary storage. By late summer/ early fall, several berry species (e.g., Rubus chamaemorus, R. arcticus, Empetrum nigrum) and local greens (e.g., Rumex arcticus) were ready to be harvested and women and children would spend most days on the tundra gathering plant re- sources. Most plants were available in a variety of locales and their harvest did not dictate moving the family to specific camps. Plants that grew in abundance in specific terrain, such as several varieties of cliff greens, usually offered other re- sources that could be harvested at the same time (e.g., fish, Sandhill cranes). Greens such as Rumex arcticus (sour dock) could be found throughout the island and all old camp sites are said to contain buried cache pits once used for plant storage (Williams and Williams 1995a). Still, several specific camps were high- lighted in oral interviews for their abundance of particular greens. These camps would be visited seasonally and are often marked by the location of numerous stone cache pits used to store the greens until their removal in the fall to the harvester’s winter residence. As an example, when harvesting “wild spinach’ or sour dock, elders state that they would stay in an area until they had harvested enough for their family’s long-term needs (Amos 1991; Kiokun 1995a). After picking, they would cook the spinach a little bit before placing it into a cache dug underground. Cook em half way, just for the leaves to just shrivel up and not take much space, and they would dig ditches and line it with a certain type of twigs and grass and put em’ in there until the weather gets colder, before the ground get hard, knowing that when it freezes, that Ciwassat? (Rumex arcticus) would freeze in with the earth. So before that time they would 80 Over there again, pull the Cizvassat out and this time leave em’ on top of the ground .... They would cover them with grass, probably willows too to keep them together and they would leave them until it freezes (Amos 1991:16), Before placing the spinach in the caches, the cooked leaves would be drained uice and the pit lined with woven grass mats. “Some people rolled them up € a ball and put them away. Each roll was made enough for one meal. They — the spinach ball big enough for their dinner or a snack. That's how they an them out of the ground” (Amos and Amos 1989:25). Grass was placed on pg the cache was covered with rocks to insure it would not be disturbed a ae (Kiokun 1995a). Berries were stored in much the same way, except Spinach Pits Would be lined with rocks (Kiokun 1995a; Whitman 1995) and raw vt Nach was used as an inner lining (Kiokun 1995a). The berries would have no juice when removed, since they would have dried out while being stored under- — In the fall, people would return to their seasonal caches and transport T Stored berries and greens to their winter village. Curtis (1930:36) describes of j 100 GRIFFIN Vol. 21, No. 2 oe C ra pl ng SR “4 : ;- Ss i AAS i ; i IGURE 3.—Rock-lined cache pits at Nash Harbor Village, Nunivak Island, Alaska. berry caches as ‘a small box-like structure of flat stones lined with grass and covered with sod until air-and water-tight.” Examples of such features were dis- covered during recent archaeological excavations on the island (see Figure 3). METHODS Earlier ethnobotanical studies among the Cup’it (Fries 1977; Lantis 1946, 1959) identified many of the plants in use in the 1940s and 1970s. Information within these studies do not always agree regarding the traditional use of island vegetation (i:e., Lantis (1946:172) states that no plant poisons were used by the Cup’it in hunting or fishing while Fries (1977:32-33) states Aconitum delphinifolium [aconite] was used by “old-timers” to make poison darts or arrows). My resear sought additional information and clarification on the Native use of indigenous plants and changes to this use over time. During my investigation, collecting expeditions were conducted on the local tundra near the villages of Nash Harbor and Mekoryuk in order to gather ex amples of utilized plants. At Nash Harbor, Cup’it crew members participating 1" a community archaeology project (Griffin 1999), pointed out significant plants and shared information on their harvest, preparation, use, and storage. On several occasions, I was able to join families on plant forays to gather seasonal greens OT berries. Plant specific information was shared on the use of various plants during these trips. While information on Native uses of indigenous plants was gathered infor- mally during the initial phase of this study, more detailed, plant specific infor- Winter 2001 JOURNAL OF ETHNOBIOLOGY 101 mation was obtained during subsequent interviews with Cup’it elders. Interviews took place between 1995-1998 and involved elders examining fresh and dried and pressed plant species, in addition to the identification of plants through pub- lished botanical guides (e.g., Schofield 1989). Interviews were conducted during all seasons of the year but fresh specimens were not always available during discussions. Pressed and dried specimens, collected while on the island, often proved of little use due to poor recognition resulting from color change and withered condition. In these cases, published botanical guides with large color plates were used to assist the discussion with information regarding plant iden- tification later collaborated with Muriel Amos, a Cup’it educator who has con- ducted preliminary research on local plant species during the process of compil- ing a Cup’ig dictionary (Amos and Amos 1999). Cup’it interpreters were used during all interviews to assist in gathering data on plant usage, although my limited knowledge of Cup’ig prevented me from freely conversing with most elders resulting in perhaps more abbreviated discus- sions of plant use. The majority of information was shared by Cup’it women (ages 66-85), although several Cup’it men (ages 73-95) also actively participated in these discussions. Ethnobotanical information shared by elders was generally con- sistent between interviews. However, knowledge of the use of a few plant species was known only by one or two individuals. When information was limited or contradictory, I have listed the source of my information in the following plant summaries. In cases where many elders offered data consistent with previously published sources, no new specific references have been cited. PLANTS USED ON NUNIVAK ISLAND The following species index details specific data on the Cup’it use of 47 in- digenous plant species on Nunivak Island. This list is compiled from plants that I collected on Nunivak Island during the 1995-1998 field seasons, and supple- mented with earlier reports of Native plant use (e.g., Fries 1977; Lantis 1946, 1959). In the following text, all species are arranged by alphabetical order (i.e., botanical name) with each species designated by its botanical name, common name, Cup’ig name, and any previously published Native name variation. In cases where the Spelling of the Cup’ig name has not been approved, I have included the Yup’ik Plant name for additional reference. Data regarding the location of each utilized Plant Species on Nunivak is also presented along with details regarding harvest, Native use, and storage. Previously published references on specific Cup’it plant use are included with documentation of current knowledge along with any com- Parative data with other Southwest Alaskan Eskimos. Previous ethnobotanical Studies in the Yukon-Kuskokwim Delta include studies on Nelson Island (Ager and Ager 1980) and the village of Napaskiak (Oswalt 1957), in addition to some oe data collected by Andrews (1989) and Lantis (1959) from several Lower rein River villages (e.g., Eek, Kasigluk, and Nunapitchuk) and by Fienup- — (1986) from several lower Yukon Delta and coastal villages (e.g., Alaka- tents Point, Scammon Bay). Figure 1 shows the location of Nunivak v Nin relation to each of these villages. In addition to the villages mentioned ve, comparisons of plant use are made with the Inuit from the Seward Pen- 102 GRIFFIN Vol. 21, No. 2 insula in northern Alaska (Jones 1983) due to the similarity in plant use and spelling of some Native plant names (i.e., Cup’ig—tInupiaq). The identification of plant specimens was obtained by using published guides to the flora of Alaska (Argus 1973; Barr and Barr 1983; Duddington 1971; Grout 1940; Hultén 1968; Viereck and Little 1972; Welch 1974;) with taxonomy following that of Hultén (1968), except in cases of identifying bryophytes, where I used Grout (1940) and Steere (1978), and for seaweeds, Abbot and Hollenbeck (1976) and Guiry (1974). Plant specimens were preserved in the field by drying in plant presses. Inclement weather and the general damp climate of Nunivak Island ham- pered the rapid drying of many plant specimens. In some cases, specimens de- teriorated to such a degree that they had to be discarded. Voucher specimens of the remaining ethnobotanical plants are currently in the possession of the author but will soon be deposited at the Yupiit Piciryarait Museum, Bethel, Alaska. Not all plant species listed in the index were identified during the current study. Previous collections of Nunivak Island flora have been collected by Eric Hultén (1968), Margeret Lantis (ca. 1946), Janet Fries (ca. 1976), Peter Stettenheim (ca. 1954), and Charles Utermohle (ca. 1973). The results of previous investigations have been incorporated here in order to provide a comprehensive summary of Cup’it plant use. The location of earlier Nunivak botanical collections include: Hultén (State Museum of Natural History, Stockholm), Lantis (University of Cal- ifornia Herbarium, Berkeley), Fries (Middlebury College, Vermont), Stettenheim (Michigan State University, East Lansing), and Utermohle (University of Alaska Herbarium, Fairbanks). Food Plants Angelica lucida L. “Wild Celery” Cup’ig: ik’itut Location: Common along shores, dunes, backshores, and on grassy river banks. Use: Very important food plant. Collected in abundance throughout the summer months and eaten fresh. Leaves and stalk first eaten at the end of June when only a large stem base and few leaves are present. Later, as flower stalk grows, they become very delicious. In late July and August older stalks become woody and lose their flavor. Not stored over winter. Elders state that plant turns bad when stored in barrels. Juiciest plants were found on bird cliffs along west coast (due to nutrient rich soil) and are still harvested by hanging over cliffs on ropes. References: | Andrews 1995; Fries 1977:44—45; Lantis 1946:178; Nowak 1975 Comparisons: Ager and Ager 1980:37; Andrews 1989:340; Jones 1983:17; Oswalt 1957:31. Siberian Eskimos inhale fumes of roasted root as seasick remedy and once carried root as amulet to ward off polar bears (Hultén 1968:705). The Inupiaq name for this plant (ikuusuk) is similar to that in Cup’ig (Jones 1983). Caution: Plant closely resembles the deadly Cicuta mackenzieana (Water Winter 2001 JOURNAL OF ETHNOBIOLOGY 103 Hemlock), one of the most toxic botanicals in North America (Schofield 1989:130). Arctostaphylos alpina (L.) Spreng Alpine Bearberry Cup’ig: kavlag Alternative: ga’valix® (Lantis 1959) Location: Common on peat mounds in wet tundra and on dry and alpine tundra. Use: Berries eaten fresh in 1940s. While berries are large and edible, no evidence of continued use was found on Nunivak in 1970s or 1990s. References: Fries 1977:46; Lantis 1959:61; Williams and Williams 1997. Comparisons: Andrews 1989:496; Jones 1983:108; Oswalt 1957:21. The Inupiaq name for this plant (kavlaq) is very similar to that in Cup’ig (Jones 1983). Caltha palustris L. ssp. asarifolia (DC.) Hult. Marsh Marigold Cup’ig: wivlut (leaves—arnat, bulbs—anngutet) Alternative: wi’vilux (Lantis 1959) Location: Found in marshes and along edges of creeks and rivers throughout island. Use: In spring, before flowering, stems and leaves are eaten when ten- der; cooked with seal oil or seal flippers. Whole plant rarely eaten raw. Some store over winter. References: Lantis 1959:60; Smith, Whitman and Shavings 1997a Comparisons: Similar use recorded for Nelson Island (Ager and Ager 1980:35) and lower Yukon Delta (Fienup-Riordan 1986:113) while roots were eaten in Nunapitchuk (Andrews 1989:340, 496). Caution: Plants contain irritant protoanemonin and should never be eaten raw (Turner and Szczawinski 1991:268). Carex L. spp. Sedges Cup’ig: pekneret Alternative: pa’knex (Lantis: 1959) Location: Found near coastal areas in moist, silty, sandy soils. se: Root and lower part of stem eaten raw; not stored. Leaves peeled off but not eaten; only the basal stem eaten. Picked in fall and mixed with akutar (Eskimo ice cream). References: Amos, Amos and Mike 1997; Lantis 1959:61; Smith, Whitman and Shavings 1997a; Williams and Williams 1995a, 1997 Cladonia Hill spp. Lichens Pig: Yup’ik: ciruneruat (Jacobson 1984) ee Found growing on rocks in tundra areas throughout island. Se: Used in soups with other available food items. Used often during times of starvation but “old timers” liked it other times as well. Referens No longer in use in 1990s. ces: Kiokun 1995b; Kolerok 1995 104 GRIFFIN Vol. 21, No. 2 Claytonia tuberosa Pall. Tuberous Spring-Beauty, “Wild Potato” Cup’ig: ulpit Location: Grows on bird cliffs along northwest coast of island. Use: Harvested in June. Corm eaten like potato. References: Tootkaylok 1997 Comparison: Possible use on Nelson Island (Ager and Ager 1980:35). Corm eat- en boiled or roasted by some mainland Natives while leaves are eaten in salads (Hultén 1968:405). Conioselinum chinense (L.) BSP. Western Hemlock-Parsley Cup’ig: Location: Common on back shores. Use: Roots of plant can be found by digging below last year’s dead flower stalks and are eaten in spring. Voucher specimen not col- lected in 1990s. References: Fries 1977:44 Draba borealis DC. or D. hyperborea (L.) Desv.(?) “Wild Lettuce” Cup’ig: ingugit Location: Grows quite large (>0.5m) high on bird cliffs and unconsolidated rocky slopes on the north shore. Use: Appears in early spring and people begin to eat them when they are still sprouts. Leaves are washed and relished raw, dipped in seal oil or mayonnaise. Also boiled in water for few minutes and stored for winter. Sometimes mixed with Rumex arcticus (ciwas- sat). Species identification uncertain. Voucher specimen not col- lected in 1990s. References: | Amos & Amos 1989; Fries 1977:36 Dryopteris dilatata (Hoffm.) Gray (2) Shield Fern Cup’ig: cilqaarat Alternative: ilgaarat Location: Located along stream banks and marsh areas. Use: Harvested when plant is dying; not when fresh. Used as tea. Not considered a medicine. Identification uncertain. No voucher spec- imen collected. References: Williams & Williams 1997 Empetrum nigrum L. Crowberry Cup’ig: paunrat or pauner Alternative: pa’unaxo'tax (Lantis 1959) Location: Dominant in dry and alpine tundra in addition to peat mounds in wet tundra and sand dunes. Fruit is not generally preferred but the abundant black berries are Picked in fall and eaten fresh or stored and mixed with other berries and eaten during winter in akutar (Eskimo ice cream): Berries were also added to sour dock and stored in barrels. References: Fries 1977:45-46; Lantis 1959:61; Nowak 1975:26; Smith, Whitman and Shavings 1997a; Williams and Williams 1997 Comparison: Use similar on Nelson Island (Ager and Ager 1980:37), the Kus- Use: Winter 2001 JOURNAL OF ETHNOBIOLOGY 105 kowim and Yukon Delta villages (Andrews 1989:496; Fienup-Rior- dan 1986:141), Seward Peninsula (Jones 1983:94), and Napaskiak but was not stored in the latter. The entire plant was also used to brew a tea by coastal people (Oswalt 1957:22). The Inupiaq name for this plant (paungaq) is very similar to that in Cup’ig (Jones 1983). Epilobium angustifolium L. Fireweed Cup’ig: Alternative: ci‘lkax (Lantis 1959) Location: Found in disturbed areas along coastline. Common in backdune areas and mesic tundra. Use: Leaves boiled for tea and occasionally eaten when tender. References: — Lantis 1959:5, 59 Comparison: Used as tea in both Nelson Island (Ager and Ager 1980:34) and Napaskiak (Oswalt 1957:22). Young shoots also harvested in early summer and eaten raw or blanched, with seal oil on the mainland and Seward Peninsula (Jones 1983:23-24). Eriophorum angustifolium Honck. Tall Cottongrass Cup’ig: pekner Location: Located in bogs and wet tundra areas. Use: Base of stem was eaten raw and considered to have a sweet taste in the summer. Bulbous underground stem was collected by lem- mings for winter use and caches were often found and eaten be- fore freeze up. No knowledge of plant use as a food source iden- tified in 1990s. References: Fries 1977:21-22; Smith, Whitman and Shavings 1997b Comparison: Stems were considered edible in Napaskiak (Oswalt 1957:27), plant greens were eaten in summer while roots were collected in fall along the lower Yukon Delta region, and the roots were eaten in Nunapitchuk while the reeds were dried and braided for use in construction of bags and mats (Andrews 1989:496). In the Sew- ard Peninsula, the base of the stem was collected from mice or vole caches and eaten raw or boiled after the root hairs have been removed. Also preserved in seal oil (Jones 1983:120).The Inupiaq name for this plant (pikniq) is similar to that in Cup’ig (Jones 1983). Fucus L, spp. Bladderwrack rep ig: elquat eration: Found washed up on beaches year round ig Harvested year-round but chiefly collected in late spring and early summer. Eaten raw or cooked with mussels or clams. Some people cook it by dipping in it hot water (turns green) then dipping in Ref seal oil. 8 elerences: Amos, Amos and Mike 1997; Williams and Williams 1995b, 1997 106 GRIFFIN Vol. 21, No. 2 Hippuris tetraphylla L. or Hippuris vulgaris L. Mare’s tail Cup’ig: tayaarut Alternative: taxa’xo (Lantis 1959) Location: Common in tundra ponds. Use: In autumn, stems and leaves are cooked with seal blubber and salmon eggs. One informant said plants are collected just before ponds freeze, leaves and stems are chopped up, cooked separately, then beaten with salmon eggs and blubber. In spring, when plant floats on ponds, it’s gathered and cooked in seal-meat soup. Only plant part above water used. Some stored over winter. References: Lantis 1959:61; Smith, Whitman and Shavings 1997b; Williams and Williams 1997 Comparison: Ager and Ager 1980:37; Oswalt 1957:22; roots were eaten in Nu- napitchuk (Andrews 1989:496). Honckenya peploides (L.) Ehrh. ssp. major (Hook.) Hult. (syn. Arenaria peploides var. major Hook.) Beach greens, Seabeach sandwort Cup’ig: tukullegat Altenative: tuku’lixax (Lantis 1959) Location: Common adjacent to tidal zone on beaches around island. Use: Actively harvested on Nunivak. Edible from spring to mid-August and collected before flowering. Leaves and stems are boiled and said to taste like buttered greens. Leaves are sometimes chopped and boiled with other plants such as Rumex arcticus (ciwassat) or with seal oil blubber & fish eggs. Leaves are often cooked inside of fish when baked in open fire. Greens are stored with dock leaves for winter. References: Fries 1977:31-32: Lantis 1959:60; Smith, Whitman and Shavings 1997b; Tootkaylok 1997 Comparison: Ager and Ager 1980:35; Jones 1983:43-44 Ledum palustre L. ssp. decumbens (Ait.) Hult Labrador Tea Cup’ig: ay'ut Alternative: ai’yu (Lantis 1959) mete: Abundant on dry tundra and on peat mounds in wet tundra. se: Picked in spring/early summer before plant flowers. Leaves are delicious used in tea. Recently used primarily as flavoring in black tea. References: Fries 1977:46; Kiokun 1995a; Lantis 1959:61 Comparison: Similar use in Nelson Island (Ager and Ager 1980:37-38), Nunap- itchuk (Andrews 1989:340, 496), lower Yukon Delta area (Fienup- Riordan 1986:113), Seward Peninsula (Jones 1983: 61), and Napas- kiak (Oswalt 1957:32) although the latter village also used dried Stalks in healing practices to get rid of ghosts. Plant contains andromedo toxins. Safe in weak tea solutions but should not be used too strong (Turner and Szczawinski 1991:267). Caution: Ligusticum scoticum L. ssp. /ultenii (Fern.) Calder & Taylor Beach Lovage or ‘Wild Parsnip/Parsley” Winter 2001 JOURNAL OF ETHNOBIOLOGY 107 Cup’ig: tuk’ayut, ciukarrat Alternative: tuxkai’yuk or ciuga’Xax (Lantis 1959) Location: Common along backdunes and sandy areas in addition to the in- terior. Use: First thing available in spring once snow melts. When plant first sprouts, roots eaten raw, dipped in seal oil or eaten without oil. Often eaten with dried fish in spring. Leaves and stems are eaten raw or dipped in seal oil or boiled and eaten as greens. By late summer, leaves gets large and are considered mildly poisonous. Cooked and added to akutar (Eskimo ice cream). Fresh leaves provide a good source of Vitamins A and C. References: Fries 1977:44; Kiokun 1995a, 1995c; Lantis 1959:60; Smith, Whit- man and Shavings 1997b; Williams and Williams 1997 Comparison: Ager and Ager 1980:37; Fienup-Riordan 1986:112; Jones 1983:14. e Inupiaq name for this plant (tukkaayuk) is similar to that in Cup’ig (Jones 1983). Lycoperdon Pers. spp. and Caluntia Fr. spp. Puffballs Cup’ig: agyam an’a(i) Location: Located in wet tundra near coastline. Use: Said to be eaten by mainlanders but not on Nunivak. Considered “feces of the stars.’” Matthiessen (1967:23) earlier reported harvest of “red mushrooms” on Nunivak but no knowledge of the Native use of fungi is recalled today. References: Williams & Williams 1997 Mertensia maritima (L.) $.F. Gray Oyster Leaf Cup’ig: ciunerturpat Location: Along coastal areas. Use: Leaves eaten on Nunivak long ago but harvest and preparation information no longer known. References: Williams & Williams 1997 Comparison: On Nelson Island, the long leafy stems were placed whole in cold water and brought to boil. They were cooked briefly and eaten with seal oil. No longer used today (Ager and Ager 1980:38). Oxycoccus microcarpus Turcz. (syn. Vaccinium oxycoccus L.) Bog Cranberry Te Yup’ik: uingiar (Jacobson 1984) aga Common in peat bogs. sip Berries eaten by people of Mekoryuk but not found in sufficient R quantity to constitute an important part of the berry harvest. ataigg Nowak 1975:26 *mparison: Ager and Ager 1980:37; Fienup-Riordan 1986:141 Oxyria dignya (L.) Hill Mountain Sorrel, ‘“Sourgrass”’ oP ig: quulistar Ocation: Abundant on cliffs in alpine tundra and in dry tundra near the coast. 108 GRIFFIN Vol, 21, No.2 Use: Beginning in spring, leaves are eaten raw, dipped in seal oil, or boiled. Larger leaves are relished by families that used to live at Nash Harbor where the plant grows in abundance along rocky slopes. Others prefer the leaves of the similar Rumex arcticus (ci- wassat), common near fish camps and Mekoryuk. Leaves were added to sour dock and berries and stored in barrels. References: Fries 1977:29; Lantis 1959:61; Nowak 1975; Smith, Whitman, & Shavings 1997b Comparison: Ager and Ager 1980:35; Jones 1983:65 Caution: Edible in moderation. If eaten in large quantities or over long pe- riods of time, they can cause poisoning and interfere with the bod- ies calcium metabolism (Turner and Szczawinski 1991:211). Palmaria palmata (L.) Stackhouse Seaweed, Dulse Cup’ig: elquat Location: Common on rocks in middle and upper tidal zones. Use: Collected in summer or during winter when ice cracks expose seaweed on rocks. Eaten raw or in fresh soup with fish, mussels or seal meat. Dipped in hot water (turns green), seal oil and then eaten. Elquat appears to be a generic name for seaweed species however no other varieties were seen or collected during 1990s. References: Kiokun 1995a; Lantis 1959:61; Nowak 1975:26; Williams & Wil- liams 1995a Parrya nudicaulis (L.) Regel (?) “Wild Cabbage”, ‘Wild Celery” Cup’ig: ingugit Alternative: inu’kit (Lantis 1959) Location: Found along cliffs. Use: Leaves usually eaten raw, occasionally boiled, or stored with dock leaves for winter use. Cliff greens. Species identification uncertain. No voucher specimen collected. References: — Kiokun 1995a; Lantis 1959-62 Pedicularis verticillata L. Woolly Lousewort Cup’ig Yup’ik: ulevleruyak (Jacobson 1984) — Common on island back shores, wet tundra, and mesic tundra. ™ Flowers of this genus are popularly called “Bumblebee food” and are picked and sucked for nectar. References: Fries 1977:50 Comparison: In addition to the use of its nectar, Nelson Island Natives are known to harvest the roots of some Pedicularis spp. in the early spring and eat them raw with seal oil (Ager and Ager 1980:38). Pohlia nutans (Hedw.) Lindb, (syn. Webera nutans Hedw. Descr.) Moss Cup’ig: kumarutet i i Alternative: ke’ 7 Location: Generally found in wet tundra ny are eet Use: In spring, seal meat is boiled w ith moss for soup. Moss sometimes mixed with seal oil and fish eggs. Also used as tea. No longer used in 1990s Winter 2001 JOURNAL OF ETHNOBIOLOGY 109 References: Burg 1941; Kolerok 1995; Lantis 1959:61; Williams & Williams 1995b Comparison: Ager and Ager 1980:33 Polygonum bistorta L. Bistort, Pink Plumes Cup’ig: ciwassat Location: Found on grassy hummocks in the interior. Use: Cup’ig name is similar to that given to several other plants (e.g., Polygonum viviparum, Rumex arcticus) but is not thought to have been actively sought on Nunivak Island due to scarcity. No infor- mation on use available during 1990s. References: — Fries 1977:30 Comparison: Jones 1983:19 Caution: Leaves of several polygonum spp. are phototoxic. They should not be eaten in large quantities or over prolonged periods (Turner and Szczawinski 1991:24, 211, 272). Polygonum viviparum L. Alpine Bistort, “Wild Rhubarb” Cup’ig: ciwassat Alternative: an.agocu’noax (Lantis 1959) Location: Common in many habitats particularly along the coastline. Use: In the early spring and summer the rhizome is collected and eaten raw. Cup’ig plant name similar to that given to several other local plants (e.g., Polygonum bistorta, Rumex arcticus). Not stored. References: Fries 1977:29; Lantis 1959:59 Comparison: Leaves of P alaskana were gathered and eaten in ea Nunapitchuk (Andrews 1989:340, 496). Caution: Leaves of several polygonum spp are phototoxic. ‘They should not be eaten in large quantities or over prolonged periods (Turner and Szczawinski 1991:24, 211, 272). rly summer in Pallas Buttercup pi nasgasax (mature) (Lantis 1959) Ranunculus pallassi Schlecht. Cup’ig: Alternative: agolu’noux (young), Location: Common in tundra ponds (submerged or floating). eae a Leaves and stems of plant are collected in spring and eaten bole They’re considered very tender and delicious. After boiling, sea oil poured over them or else shoots are boiled in seal meat soup. In late summer they are cooked with dock leaves. (Fries states that they are locally called ‘wivalook’” but she is probably referring . wivlut which is the same name given to Caltha palustris (mars marigold). Species not — a References: Fries 1977:33-34; Lantis 1946:178, © ee Comparison: Ager and Ager 1980:35; Andrews 1989:340, 496; Fienup-Riordan 1986:112 Caution: Ranunculus spp know blistering causing juice ered potentially poison 1991:104-105). n to contain varying quantities of an acrid, which yields protoanemonin. Plant consid- ous to humans (Turner and Szczawinski 110 GRIFFIN Vol. 21, No. 2 Rubus arcticus L. Nagoonberry, Arctic Raspberry Cup’ig: puuyaragur; bloom = puuyuraqur Location: Found in mesic tundra, backdunes and on peat mounds. Use: Not many on island. Berries picked from mid-August to Septem- ber and eaten fresh. Fries had earlier reported no evidence of har- vest in 1970s although well known in 1990s. References: Fries 1977:39-40; Kiokun 1997; Smith, Whitman & Shavings 1997b Comparison: Oswalt 1957:23; Jones 1983:103 Rubus chamaemorus L. Cloudberry, ‘““Salmonberry” Cup’ig: atsar atsakutag Alternative: a’tsax (Lantis 1959) Location: Abundant in many habitats including back shores, roadsides, peat mounds of wet tundra, and dry tundra. Use: Fruit is abundant all over island in mid to late August. It is the most sought-after berry on the island. Berries are eaten raw, frozen for winter use (alone or with Vaccinium uliginosum (currat) and Empetrum nigrum (pauner), or mixed with other berries into aku- tar. Cup’it believe that a long winter with lots of snow insures a large harvest the following summer. Berries were traditionally stored in seal-pokes without being cooked or stored in rock-lined underground pits that were lined with Rumex arcticus (sour dock) leaves, berries packed in, covered with more leaves, sod, then rocks. References: | Edwards 1995; Fries 1977:39; Nowak 1975:26; Williams & Williams 1997 Comparison: Andrews 1989:496; Fienup-Riordan 1986:141; Jones 1983:74; Os- walt 1959:23 Rumex arcticus Trautu. Sour Dock, Dock, “Wild Spinach” Cup’ig: ciwassat Alternative: ciwaSax (Lantis 1959) Location: Common in wet tundra areas including along tundra ponds, peat ridges and standing water. Use: Delicious and important edible plant for Nunivak people. Contains high amounts of Vitamins A and C. Young stems are eaten raw in spring, or chewed with juice sucked from them. Leaves are eaten raw with seal oil or boiled in summer. By late summer stalks are considered too stringy. For winter use, leaves were parboiled, juice drained off and placed underground in temporary caches. Braided grass mats were used to line caches with grass and willows placed on top for protection. Later stored in large wooden storage dishes; frozen. When removed from storage to make soup, it's cooked with salmon eggs and dried fish (fresh fish?) or salmon eggs and seal oil; or boiled with a little seal oil; or chopped and beaten Up with fish and seal oil. Most abundantly used plant except possibly Empetrum nigrum (crowberries). Leaves are often chopped an boiled until all flavor enters water with the resulting sour tasting mixture frozen for use in winter and taken with sugar as a drink Winter 2001 JOURNAL OF ETHNOBIOLOGY 111 or frozen dessert. Cup’ig plant name is similar to that given to several other plants (e.g., Polygonum bistorta, P viviparum) References: Curtis 1930:35; Fries 1977:28-29; Kiokun 1995a, 1995b; Lantis 1959: 59; Nowak 1975:26; Williams & Williams 1995a; Whitman 1995 Comparison: Ager and Ager 1980:35; Andrews 1989:340, 496; Fienup-Riordan 1986:112; Jones 1983:36; and Oswalt 1957:24. Plant also used in the Kuskokwim River area as a landmark and navigational aid in marshy areas because plant is known to always grow in the same place (Andrews 1989:340). Caution: Plant contains soluble oxalatis which can interfere with calcium uptakes (Turner and Szczawinski 1991:267) Salix alaxensis (Anderss.) Cov. Alaska Willow Cup’ig: qugyuguat (common name for willow spp.) Location: Found along slopes of stream banks and gravel bars. Use: Eskimo children strip the catkins of this shrub and chew them. They are commonly referred to as “Eskimo bubble-gum’”’ and are eaten before seeds ripen in June and July. References: Fries 1977:28; Williams & Williams 1997 Comparison: Similar use reported for Nelson Island (Ager and Ager 1980:34— 35), Napaskiak (Oswalt 1957:24-25) and the Seward Peninsula (Jones 1983:8), in addition to the tips of leaves being eaten raw with seal oil or added to meat or fish stews and soups. On Nelson Island, the shrub was also sometimes burned to produce ashes which were added to chewing tobacco or snuff. Salix pulchra Cham. Diamondleaf Willow Cup’ig: qugyuguat (common name) Alternative: ki’xmi°ax (Lantis 1959) Location: Located on wet tundra and along gravel bars and banks of rivers and streams. Use: Flowers were eaten raw. In 1927, Curtis recorded the use of this plant as a food source. In 1940, Lantis states that while most Cup’it denied ever eating willow leaves, one old woman said the leaves were once soaked in seal oil and eaten with dried fish. In 1990s, elders state that willow leaves were traditionally picked by Natives in Northern Alaska and that some Cup’it had recently adopted the practice. There is no memory of the traditional use of this plant R by the Cup’it. : eferences: Curtis 1930:35; Lantis 1959:60; Smith, Whitman & Shavings 1997b Comparison: Jones 1983:10; Oswalt 1957:24. Young leaves are eaten raw with seal oil by Siberian Eskimos (Hultén 1968:359). Saxifraga L, spp. pane "ede quulisstat ral Found in cliff areas Leaves are eaten fresh in spring. Tastes like lime. Species not pos- itively identified during 1990s interviews but believed to be S. 112 GRIFFIN Vol. 21, No. 2 punctata or S. spicata. Cup’ig name similar to Oxydria digyna (Mountain Sorrel). No voucher specimen collected. References: Williams and Williams 1997 Comparison: On the Seward Peninsula, S. punctata leaves were picked from spring through fall and eaten in seal oil with fish or meat or pre- served in seal oil (Jones 1983:22). Sedum rosea (L.) Scop. (syn. Rhodiola rosea | Roseroot, Stonecrop Cup’ig: megtat negiat Alternative: ca’klax (Lantis 1959) Location: Found along coastal cliffs and rocky slopes in addition to river banks, meadows, and peat mounds in wet and dry tundra. Use: Flowers boiled in water to make tea, not necessarily for medicine, just as a drink. Plant no longer in use in 1990s. References: Fries 1977:36-37; Lantis 1959:24, 60 Comparison: In earlier times this plant used medicinally to treat sores in mouth on Nelson Island but it is no longer used (Ager and Ager 1980: 36). The entire plant (stems, leaves, young flower buds, and roots) are picked, eaten and preserved each spring in many northern Alaskan communities (Jones 1983:55). Caution: Various species contain oxalic acid and soluble oxalates and should be used only in moderation (Turner and Szczawinski 1991:268) Senecio pseudo-Arnica Less. Ragwort Cup’ig: Alternative: ko’xoyu’xoax (Lantis 1959) Location: Found in well-drained sandy and gravelly soils on upper beaches and along crests of beach ridges. Use: Leaves and sometimes stems are boiled with fresh fish in late sum- mer. Also stored and eaten with dock leaves. References: —_ Lantis 1959-60 Comparison: On Nelson Island, in addition to above usage, the top of shoot is often peeled and eaten raw with seal oil (Ager and Ager 1980:38). The root is considered poisonous by Napaskiak residents (Oswalt 1957:34), Caution: Plants contain pyrrolizidine alkaloids which can produce liver- damaging compounds. Ingestion is not recommended (Turner and Szczawinski 1991:16). Streptopus amplexifolius (L.) DC. Twisted stalk Cup’ig: atsarllug Location: Found along river banks. Use: Berries make noise when chewed. Some are eaten but most spit out. Very bitter and seedy. References: Williams and Williams 1997 Vaccinium uliginosum L. Alpine Blueberry, Bog Blueberry Cup’ig: currat Location: Found in interior and along the coast on dry tundra slopes. Use: Berries are sought by natives in August. Winter 2001 JOURNAL OF ETHNOBIOLOGY 113 References: Fries 1977:47; Williams & Williams 1997 Comparison: Ager and Ager 1980:37; Andrews 1989:496; Jones 1983:79; Oswalt 957;25 Vaccinium vitis-idaea L. ssp. minus (Lodd.) Hult. Lingonberry, Low-bush Cranberry Cup’ig: tumaglir or tumaglikatat Location: Common in dry alpine tundra and on peat mounds of wet tundra. Use: Berries are very sour and eaten fresh in fall. Local preference is to wait until after the first frost or the next spring and eat berries that have remained under snow all winter. Islanders occasionally make wine from them. Berries are sometimes stored. Now used in akutar (Eskimo ice cream) and bread. References: Fries 1977:47; Lantis 1959:61; Smith, Whitman, & Shavings 1997a; Williams and Williams 1997; Tootkaylok 1997 Comparison: Ager and Ager:1980:37; Andrews 1989:265, 496; Jones 1983:87; Os- walt 1957:25—26 Medicinal Use of Plants Artemisia tilesii Ledeb. Stinkweed, Wormwood, ‘’Caribou Leaves” Cup’ig: neqnialngut Location: Common on coastal cliffs and'back shores. Se: Leaves are boiled and 1-2 cups of the infusion taken daily for a variety of ailments including asthma. Mostly used by “old timers.” Kolerok (1995) states use as medicine was introduced after arrival of Euro-Americans. References: Fries 1977:52; Kolerok 1995; Smith, Whitman and Shavings 1997b Comparison: On Nelson Island, tea was used as a laxative, for arthritic ailments, swollen areas, and as general tonic. Natives in both Nelson Island and Napaskiak applied leaves directly to wounds to stop bleeding, used on skin for infection, or crushed and applied to hands to remove or mask odors after cleaning fish (Ager and Ager 1980:38; Fineup-Riordan 1986:113). In Napaskiak, switches from this plant were also used during the sweatbath (Oswalt 1957:33). Betula exilis (Sukatsch.) Hult Birch, Dwarf Birch ots Alternative: cupu’yaxotet (Lantis 1959) Sta Found in dry tundra and peat mounds in wet tundra. e: Leaves boiled to make a tea. Medicine for stomach ache and in- Ref testinal discomfort. Fries found no use of birch in 1970s. “terences: Fries 1977:28; Lantis 1959:5, 61 Dryopteris austriaca (Jacq.) Woynar Shield Fern P's centurkar Alternative: sto’xkax (Lantis 1959) eae Found near stream banks. Fronds put in boiling water and boiled a long time to make tea. Used as medicine for stomach aches and intestinal discomfort. 114 GRIFFIN Vol. 21, No. 2 References: Lantis 1959:5, 61; Williams and Williams 1997 Epilobium angustifolium L. Fireweed Cup’ig: : Alternative: ci’lkax (Lantis 1959) Location: Common in backdune areas and mesic tundra; in disturbed areas along coastline. Use: Leaves boiled to make medicine for stomach ache and intestinal discomfort. References: — Lantis 1959:5, 59 Comparison: Ager and Ager 1980:36-37 Eriophorum L. spp. Cottongrass Cup’ig: musqu’ or melqiutet Location: Found near wet bogs and tundra Use: Cotton-like flowers picked in spring and summer by children and given to old women for wiping eyes. Also used for cuts to staunch bleeding. No distinction in use between available species. Known species include E. angustifolium, E. russeolum albidum, E. Scheuchzeri, and E. vaginatum. References: Lantis 1946:202; Smith, Whitman and Shavings 1997a; Williams and Williams 1997 Comparison: In Napaskiak, stems of plant were gathered in summer, dried, and woven for use as boot soles (Oswalt 1957:28). Cotton-like flowers were used in Eek to treat boils; method not reported (Lantis 1959: 17). Ledum palustre L. ssp. decumbens (Ait.) Hult Labrador tea Cup’ig: ay'ut Alternative: ai’yu (Lantis 1959) Location: Common throughout dry tundra, alpine tundra, and on peat mounds in wet tundra. Use: Stems and leaves used as medicinal tea for stomach ache and in- testinal discomfort and considered useful in curing colds. References: Fries 1977:46; Kiokun 1995a; Lantis 1959:61 Comparison: On Nelson Island the leaves were also used as treatment ‘for those that spit blood” (Ager and Ager 1980:37). Plants even collected in winter when wind exposed them from snow. Rubus chamaemorus L. Cloudberry Cup’ig: atsar atsakutag Alternative: a’tsax (Lantis 1959) Location: Abundant in many habitats including back shores, roadsides, peat mounds of wet tundra, and dry tundra. Use: Juice of berries drunk as medicine. References: | Edwards 1995; Fries 1977:39; Nowak 1975:26; Williams & Williams 97 Salix fuscescens Anderss. Willow Cup’ig: qimugkararat (common name for willow with “cottonballs” [cat- kins]) Alternative: pa’li (Lantis 1959) Winter 2001 JOURNAL OF ETHNOBIOLOGY 115 Use: Leaves chewed to treat sore mouth; not eaten. Old men known to put willow cotton or ‘Alaska cotton’ (cotton grass) in inner corner of eye, if suffering from watery eyes. References: — Lantis 1959:60 Salix pulchra Cham. Willow Cup’ig: qugyuguat (common name for willow spp.) Iternative: ki’xmi°ax (Lantis 1959) Location: Located on wet tundra and along gravel bars and banks of rivers and streams. Use: Leaves chewed to treat sore mouth. References: Curtis 1930:35; Lantis 1946:202, 1959:60; Smith, Whitman & Shav- ings 1997a Comparison: Nelson Island Eskimo used leaves from Salix alaxensis in similar manner (Ager and Ager 1980:34). Lantis (1959:5-6) reports that the inner and outer bark of willow (Salix spp.) was boiled and used as a gargle in one Kuskokwim River village while only the inner bark was used in another. Sedum rosea (L.) Scop. (syn. Rhodiola rosea) Roseroot, Stonecrop Cup’ig: megtat negiat Alternative: ca’klax (Lantis 1959) Location: Found along coastal cliffs and rocky slopes in addition to river banks, meadows, and peat mounds in wet and dry tundra. Use: Leaves were boiled and used for medicinal tea for stomach ache or intestinal discomfort. Flowers eaten raw as aid for tuberculosis. No one recognized use of the plant in the 1970s or 1990s. Referred to as “‘bee’s food.” References: Fries 1977:36-37; Lantis 1959:5, 24, 60; Williams and Williams 1997 Comparison: Nelson Island Eskimo used to chew roots raw to treat sores in mouth. The juice was then spit out and not swallowed. No longer in use (Ager and Ager 1980:36). Utilitarian Use of Plants Aconitum delphinifolium DC. Monkshood ruPig: —_esetegneg a el Ocation: Common in mesic tundra, backdunes and near old village sites. “i Fries told that “old-timers” used to make poison darts or arrows from plant. Lantis states that no plant poison was used on Nu- nivak and denies use of plant. No knowledge of traditional use Refer was recalled during the 1990s interviews. Co ces Fries 1977:32-33; Lantis 1946:172 aution: Plants considered highly toxic and potentially fatal. Contains aconitine and aconine (Turner and Szczawinski 1991:204—205) Carex L, spp. Sedges Cup’ig: pekneret Alternative: pa’knex (Lantis 1959) Location: Common in bogs and along coastline. 116 GRIFFIN Vol. 21, No. 2 Use: Grassy leaves picked in fall, cleaned, dried, and smoked a little to make thinner for mukluk lining and socks. References: Amos, Amos and Mike 1997; Lantis 1959:61; Smith, Whitman and Shavings 1997a; Williams and Williams 1997 Cladonia rangiferina (L.) Hoffm. Lichens, Reindeer Moss Cup’ig: Yup’ik: tuntut neqait (Jacobson 1984) Location: Common in bogs and tundra areas. Use: Used for applying oil to kayak frame or pottery. Dipped in seal oil and applied to object. Plant no longer in use in 1990s. References: Kiokun 1995b; Kolerok 1995 Elymus mollis Trin. Wild Rye Grass, Dune Grass Cup’ig: Yup’ik: taperrnag (Jacobson 1984) Location: Found along coastline. Use: Braided “seahorse grass” was traditionally used as menstrual pad for a girl's first menstruation. Leaves used for thread, woven mats and basket construction. References: Lantis 1946:178-181; Noatak 1986; Pratt 1990:77 Comparison: Nelson Island Eskimo use grass in construction of baskets, mats, and ropes (Ager and Ager 1980:34). In Scammon Bay (Fienup- Riordan 1986:113) the grass is used for basket weaving and for braiding to aid in the spring harvest of herring and tom cod. Equisetum arvense L. Common Horsetail Cup’ig kenret Location Found in a variety of habitats including marshy areas and tundra. Use: Not eaten. Stalks are used by children as play matches References: Smith, Whitman and Shavings 1997b Comparison: On Nelson Island, upper stem is brewed in tea to stop internal bleeding. Black edible nodules attached to roots are also collected and eaten. Roots are often ground up when green and added to akutar (Eskimo ice cream), or mixed with fish eggs into soup (Ager and Ager 1980:33). Caution: Common Horsetail is known to be toxic to livestock. Green veg- etative shoots should never be eaten (Turner 1995:24). Pohlia nutans (Hedw.) Lindb. heed Cup’ig: kumarutet Alternative: ke’agenax (Lantis 1959) Location: Generally found in wet tundra and bog areas. Use: Moss dried and used as children’s diapers and dressing for wounds, or soaked in seal oil for fire starter. Earlier wrapped around clay pottery (i.e., greenware) before being fired. Moss n° longer harvested in 1990s. References: Burg 1941; Kolerok 1995; Lantis 1959:19, 61; Williams & Williams 1995b Winter 2001 JOURNAL OF ETHNOBIOLOGY 117 Rumex arcticus Trautu. Sour Dock, Dock, ‘Wild Spinach’ Cup’ig: ciwassat Alternative: ciwaSax (Lantis 1959) Location: Common in wet tundra areas including along tundra ponds, peat ridges and standing water. Use: Leaves used for lining underground cache pits used for storing berries. e References: Kiokun 1995a Vaccinium vitis-idaea L. Lingonberry, Mountain Cranberry Cup’ig: tumaglir or tumaglikatat Location: Common in dry alpine tundra and on peat mounds of wet tundra. Use: Berries used for dyeing dog hair for seal gut parka decorations or rass for baskets. No longer in use in 1990s. References: Fries 1977:47; Lantis 1959:61; Smith, Whitman, & Shavings 1997a; Williams and Williams 1997; Tootkaylok 1997 Plants recognized by Cup’ig name but without knowledge of Native use: Common Botanical Name Name Cup’ig Name Palmaria mollis (Setch. & Gard.) Meer Dulse elqurlut or cinarassit & Bird (syn. Rhodymenia palmata (L.) rev.) Ulva L. spp. Sea lettuce cinarassit, cinarayet Alaria Greville spp. Ribbon Kelp cinarassit Petasites Pers. spp. Coltsfoot qallngaguar CHANGES IN PLANT USE While oral accounts have added extensive details to previous knowledge of Subsistence procurement and storage techniques of the Cup’it on Nunivak Island, One must keep in mind that the memories of earlier subsistence use may be af- fected by recent changes to island culture. The most obvious change in Cup’it indigenous plant use, from the time of Curtis and Lantis’ earlier studies, is the Current lack of use of many previously used plants. With the abandonment of all but two island villages by the early 1940s, and an increased reliance on western foods, fewer families rely on traditional subsistence resources (Nowak 1975). In time, information on earlier plant use may be forgotten and influences resulting from increased contact with mainland peoples can add or supplant earlier local knowledge. For example, in 1927 Curtis (1930:35) recorded the use of willow leaves (Salix spp.) as a food and medicinal item. In 1939, Lantis (1959:60) found only one elder who still recalled the earlier use of willow and today such tradi- tional use is routinely denied by Cup’it elders. Recent influence of northern Es- — on the island population has resulted in a renewed use of the plant, al- oe contemporary Cup’it elders believe that its use is only of recent innovation. Ww Similar pattern of traditional versus recent use has been noted for stinkweed / ormwood (Artemesia Tilesii). 118 GRIFFIN Vol. 21, No. 2 It is easy to assume that observed Native lifeways in the early twentieth cen- tury reflect those practiced during the late prehistoric period or before. However, in spite of the evident continuity of tool use and general subsistence practices on Nunivak throughout the past 500 years (Griffin 1999), the Cup’it’s traditional life- ways may have been different, possibly more complex than those historically recorded. Following increased contact with mainland Native peoples (i.e., trade, intermarriage) and Euro-Americans (after the island’s ‘discovery’ by Russia in 1821) during the nineteenth century, changes in the use of indigenous plants were probably an on-going process, influenced by the degree and type of contact with non-Cup’it people, as well as impacts from a serious loss in Native population resulting from the introduction of western diseases throughout the nineteenth century (Griffin 1999:205-208). The Cup’it historically maintained close ties with the people of Nelson Island to the east and may have assimilated mainland refugees from regional internecine warfare during the eighteenth century (Griffin 1999:158-163; Nelson 1877-1881: 60-61). As such, one would expect a similarity in plant use between Nunivak Island and Alaska mainland peoples based on their degree of contact in the past. Differences in recorded plant use may be due to local cultural variations, outside influence since historic contact, and/or loss of knowledge of the extent of past plant use. Another factor which may affect the comparison of Cup’it plant uses with those of other Yup’ik groups is the general lack of ethnobotanical data from the Yukon-Kuskokwim Delta. Previous research in Native communities within the Delta have focused on documenting changes to Native lifeways following the arrival of Euro-Americans to the region (e.g., Fienup-Riordan 1983, Lantis 1946) however, these studies have provided little detailed information on traditional use of indigenous plants. As with the present Cup’it study, the collection of ethnobotanical information was not the central focus of research efforts and a systematic analysis of Native plant use throughout region has yet to be undertaken. Given the incorporation of west- ern foods in Native diets and a corresponding decline in the harvest of many indigenous plants, additional efforts to collaborate with Native communities need to be undertaken before information on traditional use of area vegetation has been forgotten. CONCLUSION The Cup’it of Nunivak Island traditionally occupied an isolated portion of southwestern Alaska with limited contact between island residents and mainland peoples until the late nineteenth century. Having to primarily rely on locally available resources for their subsistence, the Cup’it incorporated many of the is- land’s indigenous plants into their year-round diet. As a result of working col- laboratively with the residents of Nunivak Island, information on the traditional use of 47 indigenous plant species was collected along with details regarding seasonality of use, plant harvest and storage. Contrary to earlier stereotypes of Arctic peoples’ heavy reliance on a meat-based diet for survival, island flora were routinely incorporated into the Cup’it’s diet in addition to Native pharmacology and utilitarian tasks. Winter 2001 JOURNAL OF ETHNOBIOLOGY 119 The present study comprises a survey of the Cup’it use of indigenous plants located along the north coast of Nunivak Island, Alaska, with focal areas around the villages of Mekoryuk and Nash Harbor. Given the general inaccessibility of the island’s interior and southern dunes region (i.e., lack of roads and prevailing dense fog during the summer months), a wide variety of additional plant species, more acclimatized to the island’s dry and alpine tundra and sand dunes may have been in common use by the Cup’it in the past but have yet to be documented. Prior to historic contact, the majority of island residents resided on the south side of the island near the Cape Mendenhall area (i.e., dune portion of the island). After 1930, a general shift in island population to the north side of the island (i.e., area dominated by low-lying wet tundra) occurred, induced by the establishment of an island trading post, school and mission (Lantis 1946). There have been no attempts to date, to document differences in variety and use of indigenous plants within Nunivak’s dune region. Extensive Native trail systems are known to have also once crisscrossed the island (Griffin 1999:333-334). Elders recall that trips through the island’s interior were quite common before the island school was moved to Mekoryuk in 1940 and the majority of Cup’it villages on Nunivak Island were forced to be aban- doned. Given the emphasis of the current Nunivak study on northern wet tundra areas, further research on indigenous plant use in other island vegetative regimes is needed to better understand traditional Cup’it plant use. Elders knowledgeable of traditional plant use on Nunivak remain few and younger generations have Not expressed an interest in preserving this data. Except for the continuing harvest of a few popular plant species (e.g., Angelica lucida [wild celery], Rumex arcticus [sour dock], Caltha palustris [marsh marigold], Rubus chamaemorus [cloudberry]), much of their knowledge is not being passed on and will likely disappear with the Passing of today’s elders. It is important that additional research efforts to record traditional use of plants in these areas occur before knowledge of such use ls forgotten. NOTES ' The Cup’it of Nunivak Island have a distinct culture and speak their own sub-dialect of Yup‘ik (Lantis 1984) known locally as Cup’ig (Drozda 1994) and by linguists as Cux (Ham- merich 1958, Woodbury 1984). It is the most distinct dialect within the Yup’ik language family and serves to highlight the isolation and uniqueness of the Cup’it people. Pa current Cup’ig spellings of all plant and proper names are taken from the Cup’ig ichonary by Amos and Amos (1999) and have been placed in bold italics. 3 ‘ . __teviously published Cup’ig names do not conform with current orthography (i.e, Amos nd Amos 1999). All instances have been underlined in text. ACKNOWLEDGMENTS Aa The fieldwork for this research was supported in part, by a Phillips Grant for Native “rican Research from the American Philosophical Society, and a National Science Foun- dation Doctoral Dissertation Improvement grant. I owe a special debt of gratitude to the 120 GRIFFIN Vol. 21, No. 2 many Cup’it elders who participated in my research and shared with me their extensive knowledge of the use of indigenous plants. Elders included: Nona Amos, Walter Amos, Bertha Andrews, Irene Davis, Richard Davis, Nancy Edwards, Nan Kiokun, the late Robert Kolerok, Harry Mike, Helen Noatak, Daisy Olrun, Susie Shavings, Mary Smith, Katie Toot- kaylok, Sophie Weston, Mildred Whitman, the late Elsie Williams, and George Williams, Sr. A special thanks to Muriel Amos who provided me with the Cup’ig spellings of many traditionally used plants. I would also like to extend my appreciation to the three anony- mous reviewers who provided many useful comments on an earlier draft of this paper. LITERATURE CITED ABBOT, ISABELLA A., AND GEORGE J. HOLLENBERG. 1976. Marine Algae of ern Stanford University Press, Californi AGER, THOMAS A., and LYNN PRICE AGER. 1980. Ethnobotany of the Eski- mos of Nelson Island, Alaska. Arctic Anthropology 17(1):27-48. AMOS, MURIEL, and HOWARD AMOS. 1999. Nunivak Island Cup’ig Language Dictionary. Manuscript in possession of authors, Mekoryuk, Alaska. AMOS, WALTER. 1991. Taped interview. Robert Drozda, interviewer. Hultman Kiokan, interpreter. Mekoryuk, Alaska. yuk Alaska (NIMA) Corporation, Me- ko uk. , and NONA AMOS. 1989. Taped in- terview. Ken Pratt, interviewer. Howar Amos, interpreters. Mekoryuk, Alaska. 2 April. Tape 89NUNO2. Copy in pos- session of interviewer. ——, and HARRY MIKE. 1997. Taped in- terview. Dennis Griffin, interviewer. Howard Amos and Muriel Amos, inter- pe Mekoryuk, Alaska. 8 June. Tape opy on file at the Nunivak teland Mekoryuk Alaska (NIMA) Cor- poration, Mekoryuk and in interview- er’s possession ANDREWS, BERTHA. 1995. Taped inter- view. Dennis Griffin, interviewer. Mar- vin Kiokun, interpreter. Mekoryuk, Alaska. 22 September. Tape 95NUN25. Copy on file at the Nunivak Island Me- Mekoryuk, American Philosophical So- ae alrromig s and in interviewer's pos ANDREWS, ELIZABETH FE 1989. The Ak- ulmiut: Territorial Dimensions of a cd ik Eskimo Society. Alaska Depart- ent of Fish and Game, Division of fcbcloleiee Technical Report No. 177, Juneau, Alaska. ARGUS, GEORGE W. 1973. The Genus Sa- lix in Alaska and the Yukon. National Museums of Canada, Publications in Botany, No. 2, National Museum of Nat- ural Sciences, Ottawa. BARR, LOU, and NANCY BARR. 1983. Un- der Alaskan Seas: The Shallow Water Marine Invertebrates. Alaska North- west Publishing Company, Anchorage, laska. BOS, Gregory N. 1967. Range types and their utilization by Muskox on Nunivak Island. M.S. thesis, University of Alas- ka, Fairbanks. BURG, AMOS. 1941. Nunivak Island: Nun- nee-wak. Unpublished ethnographic notes from 1941 trip. Records on file in the Amos Burg Manuscript Collection, Box 1, sir 18, Oregon Historical So- ciety, Portla CURTIS, “EDWARD S. 1930. The North ion of Canada, and Alaska. Vol. 20. 1978 Reprint. Johnson Reprint Corporation, New York. ! DROZDA, ROBERT. 1994. Qikertamten! Nunat Atrit Nuniwarmiuni: The names of places on our island. Draft manu- script assembled by the Mekoryuk IRA Co uncil, Nunivak Island, Alaska. Copy in possession of the author. ; DUDDINGTON, C. L. 1971. Beginners Guide to Seaweeds. Drake Publishers, New York. EDWARDS, NANCY. 1995. Taped inter- view. Dennis Griffin, interviewer; Mar- vin Kiokun, interpreter. Mekoryuk, Alaska. 23 September. Tape 95NUN27. Winter 2001 Copy on file at the Nunivak Island Me- koryuk Alaska (NIMA) Corporation, Mekoryuk; American Philosophical So- ciety, Philadelphia, and in interviewer's possession. FIENUP-RIORDAN, ANN. 1983. The Nel- son Island Eskimo: Social Structure and Ritual Distribution. Alaska Pacific Uni- versity Press, Anchorage. —. 1986. ‘‘When Our Bad Season Comes’: A Cultural Account of Subsis- tence Harvesting and Harvest Disrup- tion on the Yukon Delta. Aurora, Alas- ka Anthropological Association, Mono- graph Series #1. FRIES, JANET. 1977. The Vascular Flora of Nunivak Island, Alaska. Senior honors paper (Environmental Studies), Middle- bury College, Middlebury, Vermont. GRIFFIN, DENNIS. 1999. Portrait of Nash Harbor: Prehistory, History and Life- ways of an Alaskan Community. Ph.D. dissertation (Anthropology), University of Oregon, Eugene. University Micro- films International, Ann Arbor. GROUT, A.J. 1940. Moss Flora of North America North of Mexico, Volume IL. Nefane, Vermont. GUIRY, M. 1974. A preliminary consider- ation of the taxonomic position of Pal- maria palmata. Journal of the Marine Bi- ological Association, U.K. 54:509-528. HAMMERICH, LOUIS L. 1958. The west- ern Eskimo dialects. Pp. 632-639 in Pro- ceedings of the 32™ International Con- gress of Americanists. Munksgaard, Copenhagen. HULTEN, ERIC. 1968. Flora of Alaska and Neighboring Territories: A Manual of Vascular Plants. Stanford University Press, Stanford, California. JACOBSON, STEPHEN A. 1984. Yu'pik Es- kimo Dictionary. Alaska Native Lan- Suage Center, University of Alaska, Fairbanks. JONES, ANORE. 1983. Nauriat Nigifia- qtuat: Plants That We Eat. Maniilaq As- coed Kotzebue, Alaska. — NAN. 1995a. Taped interview. ©nnis Griffin, interviewer; Mona Da- vid, interpreter. Mekoryuk, Alaska. 12 os at the Nunivak Island Mekoryuk aska (NIMA) Corporation, Mekor- * American Philosophical Society, JOURNAL OF ETHNOBIOLOGY 121 Philadelphia, and in interviewer's pos- session. . 1995b. Taped interview. Dennis Griffin, interviewer; Mona David, inter- preter. Mekoryuk, Alaska. 22 Septem- ber. Tape 95NUN23. Copy on file at the Nunivak Island Mekoryuk Alaska (NIMA) Corporation, Mekoryuk; Amer- ican Philosophical Society, Philadel- Griffin, interviewer; Mona David, inter- preter. Mekoryuk, Alaska. 12 Septem- ber. Tape 9S5NUNO2. Copy on file at the Nunivak Island Mekoryuk Alaska (NIMA) Corporation, Mekoryuk; Amer- Society, Philadelphia Island Mekoryuk Alaska (NIMA) Cor- poration, Mekoryuk and in interview- er’s possession. KOLEROK, ROBERT. 1995. Taped inter- view. Dennis Griffin, interviewer; Mar- Mekory ciety, Philadelphia, and in interviewer's possession. LANTIS, MARGARET. 1946. The Social Culture of the Nunivak Eskimo. Trans- actions of the American Philosophical Society XXXV, Pt. 3. American Philo- sophical Society, Philadelphia. _ 1959. Folk medicine and hygiene: LC. MATHIESSEN, PETER. 1967. Oominmak: The Expedition to the Musk Ox Island in the Bering Sea. Hastings House, New York. NELSON, EDWARD W. 1877-1881. Ed- ward William Nelson’s Alaska Journals, Volume VI. Original manuscripts on file 122 GRIFFIN at the Smithsonian Institution Archives, eshioateny J NOATAK, "ANDREW. 1986. Taped inter- view. Ken Pratt and Bill Sheppard, in- NUNO3. Copy on file at Bureau of In- and Affairs ANCSA, Anchorage and lng Island Mekoryuk Alaska (NIMA) be ape Mek NOWAK, MICHAEL. 1975. Subsistence trends in a sr Eskimo community. Arctic 28(1):21-34. OSWALT, WENDELL. 1957. A Western Es- kimo ethnobotany. Anthropological Pa- pers of the University of Alaska 6(1):16— PALMER, L.J., and C.W. ROUSE. 1945. Study of the Alaskan Tundra with ref- erence to its reaction to Reindeer and other Grazing. U.S. Fish and Wildlife Research Report No. 10. U.S. Govern- social aspects of Nunivak Eskimo “‘cliff- hanging.”’ Arctic Anthropology 27(1): 75-86 SCHOFIELD, JANICE J. 1989. Discovering Wild Plants: Alaska, Western Canada, the Werte Alaska Northwest Books, Anchor SMITH, MARY, MILDRED WHITMAN and SUSIE SHAVING. 1997a. Taped in- terview. Dennis Griffin, interviewer. Mona David, interpreter. Mekoryuk, Alaska. 11 June. Tape 97NUNO8. Copy on file at the Nunivak Island Mekoryuk Alaska (NIMA) Corporation, Mekoryuk and in a src S possession. . 1997b. Taped interview. Pesala Griffin, interviewer. Mona David, interpreter. Meko Alaska. 11 June. Tape 97NUNO9. Copy on file at the Nunivak Island Mekoryuk Alaska (NIMA) Corporation, Mekoryuk and in interviewer's possession. STEERE, WILLIAM CAMPBELL. 1978. The Mosses of Arctic Canada. J. Cramer, ermany. SWANSON, J. DAVID, DEVONY LEHNER, JENNY ZIMMERMAN, and DALE PAULING. 1986. Range Survey of Nu- nivak Island, Alaska. U.S.D.A. (three volumes). Soil Conservation Service, Washington, D.C. TOOTKAYLOK, KATIE. 1997. Taped inter- Vol. 21, No. 2 view. Dennis Griffin, interviewer. Beth- el, Alaska. 7 June. Tape 97NUNO2. Copy on file at the Nunivak Island Me- koryuk Alaska (NIMA) Corporation, Mekoryuk and in interviewer's posses- sion. TURNER, NANCY J. 1995. Food Plants of Coastal First Peoples. Royal British Co- lumbia Museum Handbook. B Press, Vancouver. , and ADAM E SZCZAWINSKI. 1991. Common Poisonous Plants and Mushrooms of North America. Timber Press, Portland, ae ee VANSTONE, JAMES W. 1984. Mainland Southwest Alaska Eskimos. Pp. 224-242 in Handbook of North American Indi- ans: Arctic, Volume 5, David Dumas (editor). Smithsonian Institution, Wash- ington, D.C. : 1989. Nunivak Island Eskimo (Yuit) Technology and Material Cul- ture. mage New Series, No. 12. Field Museum of Natural History, Chicago. VIERECK, LESLIE A., and ELBERT L. LIT- TLE, JR. 1972. Alaska Trees and Shrubs. Agricultural Handbook No. 410. US. Department of Agriculture, US. Forest Service, Washington, D.C. WELCH, STANLEY L. 1974. Anderson’s Flora of Alaska and Adjacent parts of Canada. Brigham Young University Press, Provo, Uta WHITMAN, MILDRED. 1995. Taped inter- view. Dennis Griffin, interviewer; Mar- vin Kiokun, interpreter. Mekoryuk, Alaska. 14 September. Tape 95NUN11. Copy on file at the Nunivak Island Me- ciety, Philadelphia, and in interviewer's 0 WILLIAMS, GEORGE, SR., and ELSIE WILLIAMS. 1995a. Taped interview. Dennis Griffin, interviewer; Marvin Kiokun, interpreter. Mekoryuk, Alaska. 13 September. Tape 95NUNO05. Copy on file at the Nunivak Island Mekoryuk — (NIMA) Corporation, Mekor- ; American Philosophical Society, Philadelphia, and in interviewer's pos- session. ——,, and 1995b. Taped inter- view. Dennis Gr iffin, interviewer; Mar- vin Kiokun, interpreter. Mekoryuk, Alaska. 21 September. Tape 95NUNIS. Winter 2001 Copy on file at the Nunivak Island Me- koryuk Alaska (NIMA) Corporation, Mekoryuk; American Philosophical So- ciety, Philadelphia, and in interviewer's possession. , and . 1997. Taped interview. Dennis Griffin, interviewer; Mona Da- vid, interpreter. Mekoryuk, Alaska. 12 June. Tape 97NUN13. Copy on file at the Nunivak Island Mekoryuk Alaska (NIMA) Corporation, Mekoryuk and in interviewer's possession. JOURNAL OF ETHNOBIOLOGY 123 WOODBURY, ANTHONY C. 1984. Eskimo and Aleut Languages. Pp. 49-63 in Handbook of North American Indians: Arctic, Volume 5, David Dumas (edi- tor). Smithsonian Institution, Washing- ton, D.C. YOUNG, STEVEN B., and EDWIN S. HALL, JR. 1969. Contributions to the ethnobotany of the St. Lawrence Island Eskimo. Anthropological Papers of the University of Alaska 14(2):43-53. Vol. 21, No. 2 GRIFFIN 124 Z1-96NON ‘(6S6L) SHUR “(896T) USHTNH “961 ‘Z PEYS/SeMA (6961 ‘OP6L) SHURT “(Q96T) UPITNH ‘ZI ‘fF PeYS/seaty ZOL Pays /Sely €L-96NON “(6S61) SHUR ‘(8961) U9ITNH ILI PaYyS/Sel (EZ6L “B9) aTyoutsay “(S96L) USITNH “99E “FZE ‘ETE ‘967 “ESL ‘OLT ‘OFT PPYS/SAI “(Z961L) SO" 6Z ‘SE ‘BI PeY4S/setty (9P6L) SHURT “(Q961) UITNH “Gg PAYS /se{ Z6T POYS/S €7-96NON (8961) UPINH ‘6Z ‘SE ‘ST “SS ‘Z6Z PAYS /SOL FI-96NON ‘EI-S6NAN ‘STE “9€ ‘OZ PeY4S/SeLL LT-96NNON ‘PI-S6NNN ‘9F6L SHUR] “6zE POY4S/SALY (6961) sHuey PETZ ‘ZEE PEYS/SELA 90-96NON ‘80€ “S67 ‘9ZT ‘SLT “OBI “POT “7-7 POYS/SeM ypaiq prem MOTTIM MOTTIM B¥SeTV MOJ[LM JeatpuoueIg] TRIS paysim | sa3pag ssei3 [Ie] SaIep] ssei3 U0}}0> SseI3U0}OD [PL ssei3 ounp ‘ssei3 adr prim ee | Wey Pes [reJas1OF{ UOUTWIOD WOH (‘YPsyexNs) siixa vnjog (Ajrumey ypsrg) avadeynyjeg ‘ssiapuy suaosaosnf x1]VS ‘AOD (‘Ssiapuy) sisuaxv]V X1]VS ‘umeYyD wiyajnd xyVvS (Ajrurey MoT]IM) avareoTTes ‘Sq (7) snyofixajdiuv sndojdasjs (Ajrurey Aqr]) avaoerr] ‘dds “J xaiva "| winywuisva "FJ ‘AN uinpig]y eA saiijy “y wnjoassna “J addoyy mazyonayrs °F ‘dds -y wnaoydouyq ‘youOp] wnyofiysnSuv wnsoydoz (Ajrurey a8pas) aevaovredA ‘UlL], stjjowu snuAlq (A[rurey ssers) ovauturessy reudon (‘boef{) voviajsny sisajdohsgq Avid (WyJOH) vyvyypip sisajdohaq (Ajrurey utay pjarys) evaoerpidsy “| asuaaav winjasinby (Ayrurey [reJasIOP{) aeadeyastnby -siaquinu uawtDdads soweu UOUTLUO') {UOHRIYWUEp! pure aureU ITpTUITIS ‘ydn5 ayy Aq paztyn syueyd jo 8oye7ey— XIGNAddV JOURNAL OF ETHNOBIOLOGY 125 Winter 2001 e007 ‘OLT Pe4US/SAHA ZI-S6NON “(6S61 SHUR) WTEYUA}AS “(6G6L) SAULT (8961) USITNH “907 ‘SIL ‘PS ‘ST PPYS/SPMA POT PEYS/SPT (8961) Y9HNH ‘OST P9Y4S/SeHA (6S6L ‘9F6L) SHUT “(8961) USITNH ‘8 PeYyS/SeL{ OE “PZT PAYS /SAIA LO-Z6NOWN ‘S0-96NOAN “(6S6T ‘OF6L) SHUR] (8961) USIINH ‘PE PeYS/SeLA Z0-96NON “6S6L) sau] (S961) USITNH ‘79 ‘OE PeYS/SIMA poayeTjoo aydures ou (6S6L) sue] (8961) USIINH ‘O9€ ‘68T ‘O9T P9YS/SeA PTL PaYS/SseLy SO-S6NAN ‘(6S61) SAUeT (8961) UPINH ‘EZ ‘GOT “TST ‘ELL PPYUS/SPHA LO-S6NON “6S6L) SHUR] ‘(896L) USIINH “STE PeYS/SeLy aSevayixes paxids aSeijixes padrs]-ayepsloD JOOIISOY ,2204)2] PIIM,, _,Ad2]22 PIIM,, 40 ,A8eqqe> PIIM,, dnose}qnq sey[ed pooys yuo] pjosiuew ysieyy JJOMpues YDeaqeas _oyejod prim,, ‘Ayneaqg-Surids snosaqn ], yloysiq autdyy 410 ,,qreqnyl PIIM,, saumnyd yurg ’310}sI1g jetios ureyUNOy pop Anos “0d uoq ‘q vywaids °S JO “| vyjound vSv.fixvs (Ajruney a8vajixes) aveadeSesyIxes ‘dods (J) vasoa uinpas (Ayrurey dord auoys) avadejnsseiD ‘Asaq] (J) vasoquadhy ‘G 10 “Dq sijvasog vqviq] Jesey (J) sinvoipnu vAssog (Ajrurey paeysnyy) sesteyloniD ‘yyparups issvjjud snjnounuvy ‘Da wntofimydjap wnpruosy "NE (DA) vyofiavsy ‘dds | stajsnjod vyqvD (Ajrutey JOO, MOID) aeade;NOUNULY (‘yoopy sofeur Tea sapiojdad visvuasy “UAS) ‘yNET (OoH) 4olviu ‘dss "yay (") saprojdad vhuayouopy (Ajrumey YuTg) aeadeyjAydoAre> ‘Ted vsosaqny viuozhy]D (Ajrurey aue[sing) aevaedeoe[Nyog "| uinavdiaia unuoshjog "| vjsojsig uinuoShjog HH (1) viusip vishxo TyNeIL, snIyI4v XaUINY (Ayrurey yeaymyxang) avadvuosAjog -saaquinu usuttdeads saureu UOUTUO ,uoHeoOyUep! pue aureu IY~HUSIIS (penuyuos) x1pueddy Vol. 21, No. 2 126 (6S61) SHUR] (8961) UPIMH “Z11 P84S/Seny SI-S6NIIN ‘SEE “OZZ “ZZ1 P9YS/SPY 81-S6NNN (6S61) SHUR] (S961) UITNH “91 PAYS/SALy FO-S6NNN “(6S61) SHUR] “(8961) UMINH “el PeYS/sey £0-S6NNN “(6S61) SHUR] “(S961) UINH “¢ PIYS/sey $0-96NON ‘80-S6NN.N {6S61) SHUR] (8961) UDITNH ‘291 P8YS/sey (S961) UMINH ‘7S7 PAYS /S*y ZO-S6NON “(6961) SHUP] (8961) UPINH “Ze “TSE “9T PAYS /SP4 _Aajsied /drusaed pyyy,, 40 a8eao] ypeag WH (‘ppo}) snunu dds “| vavpi-sijia winiui290A, “| wnsomSyn wins, Suaids ("]) vurdqn sojhydyjsojnuy NE (Wy) suaqungap “dss “| aajsnjod wnpa'] (Ajrurey yyea}4) evaoeoiag “| minadiu uinajaduy (Ajpurey AasaqmosD) avareyadugy A0jAe]. 79 29P[e> ‘(usay) nuaynpy “dds “| wnoy09s wnaysndr7 (Ajrusey Adysaeg) aesayyjaquil (g961) Ua “| stavdyna “Hy (6661) SHUR] “(G961) URN “bo “11 PaYS/seug [ey saaeyy 40 “| oyphydosjay sii (Aprusey plojyrus zayey) avaoeSesoyepy LO-S6NON “9F61) SHUR] (8961) URN paomagy “| wnyofiysnSuv unigopdy (Ajuaey asosunad Buruang) avaoesZeuG OL-96NAN ‘ODE ‘797 ‘28 PAUS/seg Assagdses syaay ‘Assaquoo8eyy “| snou24y snqny 696NON ‘ZI-S6NAIN “{6S61) SHUR] “(G961) UPINEY “Ze PAYS /sen4 Asagpnol) “| Snaomamuinys snqny (Ajruaey asoy) avarvsoy suequinu uaurpadg saureu UOUTWIO YORROyHUaP! pur auTeU SyQUAIDS (panuyuos) xipueddy 127 JOURNAL OF ETHNOBIOLOGY : : (6C61) SHUT] SOTO} RVLYSND sLua}dohug] 404 AINYPUEYION *(FZ61 AZIND Pur 9261 BequatloH Pue HOgGY) SpaaMmvas 10} Pu (g/6] 2499S PUR (LPET W025) saydydodag Burdyquapt jo sased ur ydaoxe (g961) UIINH SMOTIOJ sapeds jo yuawaSuRe puke aINyRPUaWIOU YYHUS | (6661) Shuey] as[ng ‘paamras (aaas) ("]) vywuied viuauhpoyy “UAS) 9L-96NNIN asnoyyeeys (""]) vied viavijyd (aeBye pay) eeacvuewyed 1-96NAN ypeamsapprld ("| snsoynaisaa x Ajaxy| sour) “ds "| snong (aeBje UMoIg) aeadRONy spjeqynd ‘dds 14 vyyuqyvD syeqund ‘dds ‘sag uopsadooh] (Ajruney jpeqyyng) aeasepsadooA] (6S6L) SHUR] sso (19saq] “Mpay] suvjnu viagayy “UAS) SSO ‘qpury (MpaH) suvjnu vI]Yyod (Ajrumey ssou-peary) aeaoeAag 10-96NNN usp] "WyJOH] (J) vuriafiSuvs vivopy[> sudUpr'] ‘dds [["H viuopy]D (Ajrumey uaypr]) aeadetuopeyD S1-96NON (6961) SHUR] (8961) 493TH }IOMBeyy ‘ssa‘] voIusy-opnasd o12aUag ZIL-96NNN SAARI] ‘OI-S6NNAN (8961) Y9HNH +207 P9YS/SPA 80-96NNN ‘TI-S6NO.N (8961) U9INH ‘TOL ‘EPL ‘OZL P28Y4S/Se €0-96NON *(8961) UPHMH “(EZ6L ed) afyouayy pure (Z961) sog Aq paysoday noqurey,, ‘POOMULIOM ‘paaMYUYS jlomasnoy ATOOM jeaj1a}sAQ "Qape’] Jsali} VISIMAHY (Ajrurey aytsoduio>) aeytsoduio> "] ByO][i9490 StAv]NItpad (Ajrumey JJOMS1J) aeadeepNYydoryIsS Aer W'S ("]) vulva vIsuapayy (Ajrurey a8eiog) aevaseursel0g ssuaquinu uauDads saureu UOUTUIO-) \UOHRIYHUEP! Puke aUTeU IYHUEPS (penunuos) xipueddy ei ct Lt a i ry e 2 ' — + Sea meet: Winter 2001 JOURNAL OF ETHNOBIOLOGY 129 Flora of the Gran Desierto and Rio Colorado of Northwestern Mexico. Richard Stephen Felger. The University of Arizona Press, Tucson, Arizona. Pp. xi; 673, 2 maps, 19 B/W photographs, ca. 400 line drawings of plants, gazetteer, six appendices, bibliography, index. US$75.00 (hardcover). ISBN: 0-8165-2044-5. Richard Felger’s beautiful new flora—the latest volume in the University of Arizona's Southwest Center series—is a comprehensive and engaging account of plants and environments in the heart of the Sonoran desert and in the adjacent remnant wetlands of the Rio Colorado delta. The area the book covers stretches from the U.S. border on the north to the Gulf of California on the south, and from the delta of the Rio Colorado and the Mexican portion of the river on the west to about Mexico Highway 8 on the east. The roughly 15,000 square kilometers of desert plains, volcanic fields, granitic mountains, sand dunes, desert oases, small rivers, and wetlands within the flora area include some of the hottest and driest places on the North American continent. The area nevertheless supports a rather diverse flora of 589 species in 327 genera and 85 families. Of these, eight are pteridophytes, two are gymnosperms, and seventy-nine are non-native angio- sperms, the latter confined mainly to disturbed urban and agricultural habitats. The rest are native angiosperms, with dicot species outnumbering monocots by about five to one. Felger’s flora describes all 589 species, and provides keys and illustrations that should allow even the novice botanist to correctly identify the vast majority. The extensive and excellent line drawings by noted botanical artists, and Fel- gers highly accessible morphological descriptions and keys, are reason enough to purchase his flora and plan a “botanizing” trip to the Gran Desierto. But the book 's much more than a tool for identifying desert plants. It is instead a comprehen- sive introduction and guidebook to the plants, vegetation, and natural and human environments of a unique region that has fascinated Felger for over 25 years and which his book almost dares us to not also find compelling. The massive under- taking that produced The Flora of the Gran Desierto provided Felger the opportunity to share not only his extensive botanical expertise and genuine interest in plants, but also his interest and knowledge and enthusiasm for natural history, human history, and human-plant interactions in the Sonoran region. Readers familiar with Felger’s earlier publications (Felger and Moser 1985; Felger et al. 1992) will €xpect to find ample information related to ethnobiology, and will not be disap- Pointed. ‘ The broad context of Felger’s flora is established in a 36-page opening section (“Part I: The Environment and Human Interactions”) covering paleoclimate, pre- sent climate, major habitats, history and human influences, growth forms, and botanical history. The focus on geography, habitat diversity, and human history established in Part I continues in Part II (“The Flora”), in which entries for indi- Vidual taxa describe not just morphology but also geographical patterns in dis- tribution, characteristic habitats and vegetation associations, and where relevant, aspects of human interaction with taxa and historical information on first record- observations of introduced species. The gazetteer and six appendices that fol- tag floristic treatment offer further insight on the physical environment and an history of the flora area, as well as on the plants themselves. The gazetteer 130 BOOK REVIEWS Vol. 21, No. 2 of place names and locations (with latitude and longitude accurate to the nearest second) includes information on, for example, the depths of natural bedrock wa- terholes, the ages and compositions of lava flows and the origins of their names, the early history of Mexican settlements, and the dates for the construction and paving of different roads in the flora area. The appendices include tables on growth forms and distributions of species (Appendix A); habitats of plant species in a volcanic crater (data for Syke’s crater, but probably extrapolatable to others; Appendix B); commonly cultivated trees and shrubs, focusing on three settle- ments (Sonoyta, San Luis, Puerto Pefiasco; Appendix C); non-native plants and habitats (ruderal, disturbed, natural; Appendix D); and the relative abundance and dependence on human disturbance (Appendix E) and geographic distribu- tions (Appendix F) of grasses in the flora area. is volume is a treasure that belongs in the library of every ethnobiologist, geographer, anthropologist, botanist, and ecologist working in North American deserts. Why then, does perusing this book bring me sorrow as well as delight? For the simple reason that I wonder how much longer books like this will be written. Are we training and encouraging and rewarding students of botany to have the depth and breadth of knowledge of plants and their environments that Richard Felger brought to bear in this splendid monograph? In a recent com- mentary in Systematic Botany, Lammers (1999) wondered about the direction the systematic community is headed, with more and more of its practioners involved solely in ‘‘cladistic analysis of gene sequences.” He asked, “Will the ‘taxonomist’ of the coming century be someone who doesn’t know plants as living organisms integrated in their environment? Will a diverse community schooled in multiple disciplines give way to a cadre of lab technicians ... who know their plants only as extracts in a glass tube? Will no one be left who can write a Latin diagnosis, count chro- mosomes, perform experimental hybridizations, or use (much less write) a dichotomous key?’’ Richard Felger’s magnificent Flora of the Gran Desierto and Rio Colorado of North- western Mexico is a potent argument that we should not—must not—let this hap- pen. Buy it, read it, use it, and share it with your graduate students and with foundation and funding officers. Our understanding of biological diversity and ability to conserve and manage it depends on our ability to answer basic questions about the identity of plant species, how they differ from each other, and where they grow (Lammers, 1999). Our need for information on plants and their envi- ronments and interactions with human society will only grow in the more crowd- ed world of the future. We need more, not fewer, books like this one, and we need to be training and supporting now the students who will someday write them. Sally P. Horn Department of Geography University of Tennessee Knoxville, Tennessee USA Sturn ie Winter 2001 JOURNAL OF ETHNOBIOLOGY 131 LITERATURE CITED FELGER, R.S. and M.B. MOSER. 1985. People of the Desert and Sea: Ethnobotany of the Seri Indians. University of Arizona Press, Tucson, Arizona. , PL. WARREN, S.A. ANDERSON, and GP. NABHAN. 1992. Vascular Plants of a Desert Oasis: Flora and Ethnobotany of Quitobaquito, Organ Pipe Cactus National Monument, Arizona. Proceedings of the San Diego Society of Natural History 8:1-39. LAMMERS, T.G. 1999. Commentary: Plant Systematics Today: All Our Eggs in One Basket? Systematic Botany 24(3): 494-496. eee 08 oe ener nR aie Mee Journal of Ethnobiology 21(2): 133-144 Winter 2001 TAXONOMIC IDENTITY OF “HALLUCINOGENIC” HARVESTER ANT (Pogonomyrmex californicus) CONFIRMED KEVIN P. GROARK Department of Anthropology, University of California, Los Angeles Los Angeles, CA 90024 ABSTRACT—tThe use of California harvester ants (Pogonomyrmex californicus) for visionary and therapeutic ends was an important but poorly-documented tradi- tion in native south-central California. In this brief report, a confirmation of the taxonomic identity of the red ant species used in California is presented, and the descriptive record of its use is supplemented with additional ethnographic ac- counts. This taxonomic identification of this species is of particular importance, as visionary red ant ingestion provides the only well-documented case of the widespread use of an insect as an hallucinogenic agent. RESUMEN.—La utilizaci6n de hormigas granivoras rojas (Pogonomyrmex califor- nicus) con fines alucinégenos y terapéuticos, fue una tradicion de mucha impor- tancia pero mal documentada en el sur y centro-sur de California. Este breve articulo confirma la identidad taxonémica de dicha especie y la descripcién de su uso se hace a través de datos etnograficos adicionales. Esta identificacion taxo- nomica es de especial interés, puesto que €s el unico ejemplo etnografico debi- damente documentado de un agente alucinogeno derivado de un insecto. RESUME.—1 utilisation des fourmis moissonneuses rouges (Pogonomyrmex califor- nicus) 4 des desseins religieux et thérapeutiques était une tradition peu docu- V'identification taxonomique de la fourmi et a la description de la méthode de son utilisation s’ajoute des données ethnographiques suplémentaires. Linteret de ce sujet est considérable car il s’agit 1a du premier exemple ethnographique bien documenté d’un agent halluncinatoire que provient d’un insecte. INTRODUCTION This report supplements an article previously published in this journal under the title, “Ritual and Therapeutic Use of ‘Hallucinogenic’ Harvester Ants (Pogo- nomyrmex) in Native South-Central California’ (Groark 1996). In this earlier paper, | presented an overview of a widespread, but poorly documented, tradition of Visionary and curative red ant ingestion among native southern Californian In- dians. Building on several key ethnohistoric accounts from the unpublished field- Notes of Smithsonian ethnologist and linguist John P. Harrington (as well as a number of obscure published sources), I reconstructed the general details of this ant ingestion tradition,” outlining its cultural distribution and probable origins. The Paper closed with a discussion of ant venom bioactivity and toxicology, as 134 GROARK Vol. 21, No. 2 well as preliminary suggestions concerning likely biochemical bases for the psy- choactive effects reported in the ethnographic record. Recently, another early account written by J.P. Harrington has come to my attention. In addition to supplementing our understanding of ritual ant use with additional ethnographic details from the Luiseno-Juaneno Indians, Harrington also provides us with a precise taxonomic identity for the red ant species used in native southern California. This “new’’ account is particularly significant in its confirmation of the speculative taxonomic identification offered in Groark (1996). In addition, a set of Pogonomyrmex specimens collected by Harrington has been located in the ant collection of the Smithsonian Institution, further increasing the certainty of the identification. In the present report, I provide a brief summary of the major features and distribution of ritual and therapeutic red ant use, followed by a presentation and discussion of the aforementioned Harrington account (which is currently acces- sible only in a very rare edition), as well as a description of the newly located specimens. The paper closes with a discussion of the significance of this taxonom- ic confirmation for future toxicological studies of Pogonomyrmex species and their utilization in visionary contexts. This identification is of particular importance, as it provides the only well-documented case of the widespread use of an insect as an hallucinogenic agent. OVERVIEW OF CULTURAL DISTRIBUTION Visionary Use of Red Ants.—Ingestion of red ants for visionary and shamanic ends was most highly developed among the indigenous groups of south-central Cali- fornia, seven of which are known to have engaged in the practice. The ants were swallowed alive and unmasticated, in massive quantities (often exceeding 400 ants), in order to induce a prolonged state of unconsciousness during which tu- telary spirits (usually referred to as ‘‘dream helpers’ or ‘‘suertes’”) appeared to the aspirant, often becoming life-long supernatural allies. These visions, which often took the form of animals or personified natural forces, were highly sought after by young men—quite apart from any specific skills they might confer, dream helpers (and the power they embodied) were critically important in leading a safe, healthy, and prosperous life. In addition, men who aspired to be shamans would ingest repeatedly red ants or the potently hallucinogenic toloache (more commonly known as Jimsonweed; Datura wrightii Regel) over a period of months or years. If they were fortunate, they gradually acquired multiple or specialized dream helpers who bestowed extraordinary shamanic skills upon them. (See Groark [1996: 7-11] for detailed accounts of the ritual administration and result- ing visions.) The ingestion of red ants in visionary contexts appears to have been strongest among the Shoshonean groups occupying the southeastern edge of the south- central region of California—the Kitanemuk (Harrington 1986b:r1.98, frs.449-450), Kawaiisu (Zigmond 1977:62, 1986:405), Tiibatulabal (Voegelin 1938:5, 46, 67-68), and the various Hokan-speaking Chumash groups, particularly the Interior Chu- mash (Harrington 1986b:rl.98, frs.608-609, 648-652). In the Central Valley to the north, some of the neighboring Southern Valley Yokuts (particularly the Yawel- Winter 2001 JOURNAL OF ETHNOBIOLOGY 135 mani) and Southern and Central Foothill Yokuts (Wikchamni, Yawdanchi, Bok- ninwad, Yokod, and Palewyami) also swallowed ants in order to gain dream help- ers and shamanic power (Harrington 1986a:11.94, fr.387; Driver 1937:99), but the practice among these latter groups appears in a somewhat attenuated form. The Northern Miwok are also reported to have ingested ants “for vision or power” (Aginsky 1943:440). Collectively, these groups constitute the core of the visionary ant ingestion tradition. Based on the reported distribution, the practice appears to have devel- oped among the Shoshonean-speaking groups of the southern Sierra Nevada re- gion, spreading to the Interior Chumash to the west, then on to the various Yok- utsan groups occupying the southern end of the San Joaquin Valley. Interestingly, this distribution is largely coextensive with the Toloache-Dream Helper complex, an egalitarian religion stressing individual contact with the supernatural and the acquisition of one or more dream helpers (usually mediated through the ingestion of Datura wrightii Regel). Boys’ Ant Ordeal.—A number of groups in southern California also administered the ants externally (and on occasion, internally as well) in the “ant ordeals” of boys’ initiation ceremonies. These ordeals were ubiquitous among, the Takic- speaking Cupan groups in southern California (Gabrielino-Fernandeno, Luiseno- Juanefio, Cahuilla, Cupenio), especially those involved in the proto-historic Chin- gichngish religion.! It should be emphasized, however, that these “ordeals” lacked the visionary component that formed such an important part of ritual ant use as reported from the south-central groups. In 1852, Hugo Reid described the ant ordeal of the Gabrielino as follows: To make them hardy and endure pain without wincing (for cowardice as to corporeal suffering was considered even among the women as dis- graceful) they would lie down on the hill of the large red ant, having handfuls of them placed in the region of the stomach and about the eyes. Lastly, to ensure a full dose, they swallowed them in large quantities, alive! [Reid 1968 (1852): 36]. In a revealing comment, one of Harrington's Kitanemuk informants identified these ants as being identical to the vision-inducing red ants used by the south- central groups described above (Harrington 1986a:rl.98, fr449). A number of ethnographic accounts indicate that similar ant ordeals were found further to the north among, the Chumash (Hudson 1979:73), the Tubatu- labal (Driver 1937:98), the Northern Miwok (Aginsky 1943:440), and possibly the Onache (Driver 1937:99). Among these groups, the ordeal often lacked the for- mal initiatory function found among the groups that were integrated into the Chingichngish religion. Instead, the practice served to mark the transition from youth to adulthood. __ It is interesting to note that, although visions are not reported to have man- ifested, loss of consciousness was common during these ordeals and appears to ave been an explicit goal. Profound loss of consciousness was considered essen- tal to shamanic, visionary, and initiatory practice throughout the region, and was understood to represent a sort of “small deat ” im which the aspirant was 136 GROARK Vol. 21, No. 2 “killed” by the supernatural agents which he wished to contact. Despite the lack of associated visions, the goal of the ant ordeal was largely identical to that of visionary ant ingestion—augmentation of individual strength and fortitude, and the establishment of a personal connection with supernatural power. Both vision- ary ingestion and the ant ordeal of boys’ initiation ceremonies represent the in- dividual’s first personal contact with supernatural power—a connection which he could then draw on in daily life for vigorous health, luck in hunting or gambling, or for more esoteric purposes (see Groark [1996: 9-10, 16-17] for additional de- tails). Therapeutic Uses.—In addition to the esoteric uses outlined above, the ants played an important role in both curative and preventative medicine, treating a diverse inventory of common ailments, including: paralysis, gastrointestinal ailments, se- vere colds, pain, arthritis, and gynecological disorders (particularly those occa- sioned by childbirth). Ethnohistoric accounts indicate that initiatory and thera- peutic ant ingestion persisted through the Mission Period (in some cases, surviv- ing until at least the mid-1850’s), but these practices appear to have been aban- doned by the turn of the century (see Groark 1996: 11-16 for a detailed discussion). A NOTE ON INDIGENOUS NOMENCLATURE A brief survey of indigenous nomenclature reveals striking homogeneity in the name applied to this ant among Takic-speakers of both the Serran and Cupan branches. The ant used in these ceremonies was referred to by the Kitanemuk as ‘anaqt or ‘anoht (pl. anom). Zigmond records the Kawaiisu name as aanat (“big red ant—eat for pain’) (Unpublished 1937 fieldnotes of M.L. Zigmond; quoted in Anderton 1988:270), while the Luisefio-Juanefio term was anut (“red ant’’) (Kroe- ber 1925:672). It should be noted that this name was not a generic term for “red ant’. Rather, it applied specifically to the ‘medicinal red ant’ used in ritual and therapeutic contexts, with other local species being referred to by distinct names (see Anderton 1988: 597; Harrington 1933:164, note 128). Neighboring non-Takic groups had very different names for this ant—the Chumashan groups appear to have used the term shutilhil (Walker and Hudson 1993), while various Yokutsan speakers of the Tule-Kaweah dialects (Yawdanchi, Wikchamni, Gawia, Bokninwad, Yokod), referred to these ants as k’awk’aw, ‘‘cra- zy ants,’ possibly in reference to their intoxicating potential (Harrington 1986a: rl.94, fr.382). WAS POGONOMYRMEX THE SPECIES USED IN CALIFORNIA? Despite the surprising detail and high quality of many of the sources cited above, these early accounts provide neither the common nor scientific name for the ant species in question. As a result, I was forced to assume a somewhat speculative tone in the previously published article (Groark 1996). Based on an analysis of the biological and behavioral details provided in the ethnographic literature, I concluded that the ant was most likely a Pogonomyrmex species, but acknowledged the problems inherent in any precise identification: a Winter 2001 JOURNAL OF ETHNOBIOLOGY 137 The taxonomic status of the red ant species used in aboriginal California is uncertain. All ethnographic accounts describe them merely as “large red ants”... The accounts uniformly emphasize their large size, the fact that they build small mounded nests, and the excruciating pain of their sting... Unfortunately, no voucher specimens were collected when the ethnographic accounts were recorded, and the precise taxonomic identity of the ant species must therefore remain tentative. However, the taxonom- ic and toxicological literature strongly support the assertion that a Pogo- nomyrmex species was indeed the red ant referred to in the ethnographic accounts. Of all the ant genera present in California and the Great Basin, Pogonomyrmex is distinguished by the large size, exceptionally painful sting, and highly biodynamic venom of its representative species. [Groark 1996:3] Based on the ecological distribution of the various Pogonomyrmex species pres- ent in California, it seemed probable that the most common and conspicuous species, P. californicus, was the ant referred to in the accounts. Based on this in- ference, | proceeded to examine the ethnographic accounts in light of general biology and toxicology in order to assess possible pharmacological underpinnings for the reported visionary and therapeutic effects. While the results were far from conclusive, a survey of the toxicological lit- erature indicated that the Pogonomyrmex species present in California possess po- tently toxic venom containing a number of highly bioactive compounds, includ- ing: kinins, peptides, and neurotoxins, as well as complex alkaloids previously known only from certain higher plant taxa. In large quantities, these venom con- stituents are capable of acting on the mammalian central nervous system, trig- gering a wide range of psychophysiological reactions that includes highly altered metabolic states resembling those reported ethnographically. In addition, Harvester ants of the genus Pogonomyrmex have been shown to possess the most toxic insect venom recorded to date. Their venom has the highest own mammalian lethality of any arthropod—it is 5 times more toxic than the venom of the Oriental hornet, and 8 to 10 times more toxic than honeybee venom (Schmidt and Blum 1978a,b,c). Based on unpublished venom lethality data for P californicus provided to me by Justin Schmidt, I determined that the doses em- Ployed in visionary contexts by California Indians were clearly within the range of pharmacological activity, representing approximately 35% of a lethal dose for an individual with a body weight of 100 Ib. (45.5 kg). (See Groark [1996: 17-22] for a full discussion of venom toxicology and complete LD5o calculations). _ Despite these compelling data, my argument was weakened by the uncer- tainty of the taxonomic identity of the ant. I was therefore extremely pleased to come across a key reference which resolved this ambiguity—a footnote written °y John P. Harrington in his 1933 annotation of the Relacién Historica, Fray Ger- Onimo Boscana‘s classic Mission Period account of the Luisefto-Juanefio Indians of Southern California. ; this extensive note, Harrington clearly ‘dentifies the ant species n question 8S Pogonomyrmex cal ifornicus Buckley, and provides additional ethnographic details based on his own field research with surviving Luisefio-Juaneno individuals (the 138 GROARK Vol. 21, No. 2 bulk of which was carried out intermittently between 1919 and 1933). Due to the rarity of these accounts, I will reproduce two variant versions of Boscana’s original text as well as Harrington’s annotation in full. The “New” Accounts: Two Versions and an Annotation.—The author of these accounts, Ger6énimo Boscana, was a Franciscan friar who lived among the predominantly Luisefio-Juaneno amalgamation of Indians at Mission San Juan Capistrano from May 1814 to January 1826. While there, he assiduously recorded all details of life in the pre-mission period with the help of three Luisefio-Juanefio men—two of whom were local chiefs, and the other a shaman. The resulting account, properly known as the Relacién Hist6rica, was probably first compiled around 1822, and remains one of the earliest and most detailed descriptions of aboriginal life in native southern California.’ In several brief passages Boscana mentions the therapeutic use of large red ants by the local Indians when they were still ‘in their heathen state.” The ants were applied externally in the treatment of unspecified “pains”: ... the most frequent and commonest practice, especially when in pain, was to whip the place where the pain was with nettles, and to put them right on the place of the pain, and likewise ants, and these latter especially on sores, and in this manner they cured themselves. [Harrington 1934: 49} Boscana’s most extensive description relates to the “ant ordeal’’ that formed the conclusion of the boys’ initiation ceremony into the Chingichngish religion of the Gabrielino, Luisefio, and Juaneno Indians. All boys were subjected to this ordeal, which was performed during early adolescence in order to “harden” the youths, to provide luck and skill in hunting, and to ensure a long life. Robinson's 1846 translation of the Relacién Hist6rica describes it in the following terms: The Indians were obliged to undergo still greater martyrdom to be called men, and to be admitted among the already initiated, for, after the cer- emony of the potense [ritual initiatory branding with Artemisia vulgaris L.], they were whipped with nettles and covered with ants that they might become robust. This infliction was always performed in summer, during the months of July and August when the nettle was in its most fiery state. They gathered small bunches which they fastened together and the poor deluded Indian was chastised by inflicting blows with them upon his naked limbs until he was unable to walk. He was then carried to the nest of the nearest and most furious species of ants, and laid down among them, while some of his friends, with sticks, kept annoying the insects to make them still more violent. What torments did they not un- dergo! What pain! What hellish inflictions! Yet their faith gave them power to endure all without a murmur, and they remained as if dead. Having undergone these dreadful ordeals, they were considered as invulnerable, and believed that the arrows of their enemies could no longer harm them. [Robinson 1846; reproduced and annotated in Harrington 1933: 47] A slightly different account of this event is found in J.P. Harrington's translation of the ““Cessac manuscript” of the Relacién, which reads as follows: Winter 2001 JOURNAL OF ETHNOBIOLOGY 139 After this sacrifice [the potense ceremony], having been well lashed with nettles, they placed the patient on a nest of fierce ants, and another one was stirring them up to make them still fiercer, and since the patient had no more clothes on than what he brought from the belly of his mother, we can imagine in what condition he must have been, after having been thoroughly lashed with nettles, as a result of those fierce ants, which even cause fever. And so great was their patience, that they seemed like dead, without a groan or movement. These were the ones called cured. There were some who suffered through this torture several times over, and many went through it alone or with some companion, for they believed that when thus cured, they were from that time on more agile, and that the arrows of their enemies could not harm them.” [Harrington 1934: 19] In his annotation to the first of these two passages, Harrington elaborates on Boscana’s basic account, including observations derived from his own ethnograph- ic research among the Luisefio and Juaneno Indians: The ants used in the ant stinging of the boys’ ceremony were [called] ‘aanat, pl. ‘antum, Pogonomyrmex californicus Buckley, California Harvest- ing Ant. This is a good-sized red ant, the medicinal ant of these people. It is plentiful throughout the region, making large nests in the ground, and is not much of a climber, being unable to climb out of a bottle. When irritated, it stings with its abdomen, injecting formic acid, and bites with its mandibles at the same time. The ant dies after a time, his carcass still clinging to the skin of the person stung if the attachment is successful. The sting is claimed by the Indians to be as painful as a European bee sting, and hurts noticeably for fifteen minutes or more.* Doubtless when the Indians lay about the camps naked they were stung much more fre- quently than at present. : When these ants were used as medicine, to relieve rheumatism, 1n- ternal pains, and the like, one method was to pick a number of the ants, one after another, and place them on the afflicted part, where they stung and were allowed to remain until they dropped off or got accidentally brushed off; Eustaquio [Lugo] once cured himself by putting a dozen or more of them on his bosom thus and leaving them on for hours. Another and evidently more modern method is to put a goodly number of the ants in a piece of cheesecloth and press it against the afflicted part, where- upon the ants sting through the cheesecloth. This cloth method is said to have been used in the boys’ ceremony, but the earlier method was un- doubtedly to seat and lay the named boy on a nest of these ants, or better to dig out the nest and seat and lay him in the teeming hole. There was not a part of the boy that was not stung and the ordeal was continued until the boy fainted or weakened, and all this without a murmur on the Part of the boy. The ants were also administered as medicine given to Sick people internally, being swallowed alive, but | have not found an informant who recalled that they were swallowed in the boys’ ceremony igh [Harrington 1933:164, note 128] 140 GROARK Vol. 21, No. 2 Later in the note, Harrington indicates that the Luiseno-Juaneno referred to this ritual as ‘antush (< ’aanat “red ant’”)—literally, “an anting’’ (Harrington 1933: 164, note 128)! Although this account was published in 1933, Harrington’s notes indicate that he had been collecting data on ritual and initiatory ant ingestion intermittently since at least 1910 among the Kitanemuk, Interior Chumash, and various Yokutsan groups. Unfortunately, the descriptions contained in his manuscript fieldnotes contain only indigenous names for the ants—no common name or Latin binomial was provided. The above account is therefore of great importance, as it provides us with the first proper taxonomic identification of the species involved. Harrington’ Identification: Inference or Scientific Determination?—Despite the excite- ment of finding Harrington's note confirming my earlier speculative identification, a nagging question remained: How did Harrington arrive at this identification? Was it merely an inference derived from a general familiarity with the southern California environment, or was it based on properly determined voucher speci- mens? We know that Harrington was an obsessively meticulous fieldworker. In ad- dition to collecting careful data on indigenous nomenclature and usage, he was also a conscientious collector of botanical and zoological specimens (most of which, unfortunately, have not survived in an identifiable state). In an interesting twist to this story, Dr. Ted Shultz—a myrmecologist at the Smithsonian Institu- tion—discovered a set of Harrington’s vouchers in the Smithsonian’s ant collection after reading a draft version of this paper. Stored just 15 feet from his office door, Dr. Schultz found a specimen set consisting of six pins holding four workers, one male, and one female. The spec- imens are collectively identified as “ Pogonomyrmex californicus (Buck) sp. det Roh.”, and each bears an identical label reading: ‘’J.P. Harrington, Collector.” According to Schultz, the identification label indicates that the species determination was made by Sievert Allen Rohwer, a hymenopterist who worked at the Smithsonian's National Museum of Natural History from 1909 to 1951. From 1925 to 1937, most ant identifications were referred to Rohwer, suggesting that Harrington deposited the specimens during this period. Although the specimen labels indicate that the ants were collected in Cottonia, Arizona (and not southern California), their discovery—when considered along with their probable date of deposit—strongly suggests that Harrington’s 1933 identification was indeed based on properly documented and determined voucher specimens (or at the very least, that his published identification derived from voucher specimens collected after his Luisefio-Juanefio fieldwork, but before his 1933 Boscana annotation). CONCLUSIONS The combination of three lines of evidence—the physical and ecological de- scription of the species, Harrington’s precise 1933 entomological identification, as well as the discovery of his Pogonomyrmex voucher specimens—allows us to make a strong argument that Pogonomyrmex californicus was, in fact, the ant species used Winter 2001 JOURNAL OF ETHNOBIOLOGY 141 for visionary and medicinal purposes in native California. That such an identifi- cation can be confirmed more than a century after the species’ last known use is eloquent testimony to the importance of voucher specimens in anthropological research, as well as to the importance of the collections that preserve such ma- terials. Despite the fact that our knowledge of red ant ingestion comes principally from a patchwork of early ethnohistoric accounts, these narratives—when consid- ered in their entirety—provide us with a remarkably complete and well-attested ethnographic example of the use of an insect as an hallucinogenic agent. Although there have been scattered references to non-botanical hallucinogens, most prior claims have suffered from a lack of documentation—either inadequate ethno- graphic descriptions or a confusion surrounding the identity of the species in question.> With the publication of this report, the taxonomic identity of the red ant used in native California has been confirmed, and the descriptive record of its use is supplemented with several additional ethnographic accounts. This new taxonomic certainty places future toxicological investigations on a much firmer footing, adding a key piece to our reconstruction of “hallucinogenic” harvester ant use in native south-central California. NOTES 'The Chingichngish religion is classified as one of two major religious subsystems that developed out of the Datura-based toloache cult of southern California (Kroeber 1925; Black- burn 1974). The Chingichngish religion appears to have originated among the Gabrielino during the proto-historic period, then spread to neighboring groups, possibly through indigenous evangelization (Bean and Vane 1978). Its doctrine centered around mythic ac- set of beliefs, which appear to have become integrated with older local traditions. Unlike the toloache cult—an egalitarian religion based on vision seeking and the acquisition: of ‘dream helpers” through the ceremonial ingestion of Datura wrightii—the Chingichngish religion was characterized by esoteric doctrine, highly formalized rituals and initiations, and the construction of ceremonial enclosures into which only the initiates were admitted (hence the frequent reference to the Chingichngish “Cult’). For more detailed information, see Johnson (1962) on the Gabrielino, and Sparkman (1908), DuBois (1908), and White (1963) on the Luisefo. *There were at least three versions of Boscana’s original account, only one of which is known to have survived. Based on the surviving copy, the original title appears to have been “Relacién histérica de la creencia, usos, costumbres, y extravagancias de los indios de esta Mission de San Juan Capistrano llamado la nacién Acagchemem.” The first full published version of Boscana’s account was Robinson's (1846) English translation, retitled “Chinigchinich and published as an appendix to the first-edition of his book Life in California. (Robinson chose the title “Chinigchinich” because of the prominence of this mythical figure in Bos- cana’s account, and it has since become the de facto name for this document.) His translation appears to have relied upon two slightly different original manuscripts, both of which have been lost (however, stylistic peculiarities suggest that the Cessac manuscript described below was one of the source versions). In 1933, J.P. Harrington republished Robinson's Translation, supplementing it with 132 pages of ethnographic annotations (as a result, this edition is often referred to as ““Harrington’s Chinigchinich.”) Sometime during this period, 142 GROARK Vol. 21, No. 2 Harrington also succeeded in locating a ‘‘new lost original Boscana’’ manuscript in the Bibliotéque Nationale in Paris. This version—now known as the ‘‘Cessac Manuscript’’—is written in Boscana’s own hand, providing us with the only surviving original manuscript. This version, which differs in some details from Robinson's translation, was published in English by Harrington (1934) and in the original Spanish by Reichlen and Reichlen (1971). For the sake of clarity, I will refer to all versions of the text as Boscana’s Relacién Historica, but I cite them under the surname of the translator or editor in order to distinguish between the numerous variant editions. } This practice appears to have been based on the principle of counter-irritation, and was widespread among southern and south-central Californian groups. Interestingly, the venom of the ant Pseudomyrmex has been shown to be an efficacious treatment for chronic rheu- matoid arthritis (Schultz and Arnold 1978), and there is evidence that a component in oney-bee venom alleviates arthritic pain and associated symptoms (Dr. Roy Snelling, Los Angeles Museum of Natural History: personal communication 1995). * Pogonomyrmex stings are exceedingly painful ont eee and have been described as spproxinating ‘ripping muscles or tendons” or “turning a screw in the flesh around the sting site’—and all of this accompanied iy a nervous, chilling sensation that sweeps upward from the site of the sting (Schmidt 1986). * The only well-documented hallucinogen of nen botanical Re go comes from the Sonoran Desert Toad (Bufo alvarius Girard), which ties of 5-MeO-DMT in its venom glands (Weil and Davis 1994). A number of seh initere toads and frogs (mostly Dendrobates, Phyllobates, and Phyllomedusa) also secrete toxins which are used by the Amahuaca and Matsés Indians of the Peruvian Amazon in hunting magic, although visions are usually not reported (Carneiro 1970, Amato 1992). Interestingly, these intoxicating cu- taneous alkaloids are not endogenously produced—rather, they are sequestered from di- etary sources which include alkaloid-rich myrmicine ant species (Daly 1994). The only reference to an insect-based hallucinogen is an anecdotal report by Saint-Hilaire (1824) referring to a larval moth (Myelobia smerintha Huebner) used by the Malali Indians of Brazil to produce an opium-like, dream-filled sleep. While Britton (1984) has proposed that the gut or salivary glands of this larval moth be classified as a new hallucinogen, Ott (1993: 414) argues that, if confirmed, the moth is more accurately regarded as an ‘‘oneirogenic” or “dream-inducing”’ agent, and classifies all of these cases as ‘putative’ hallucinogens. LITERATURE CITED AGINSKY, BURT W. 1943. Culture gyre Distributions: XXIV Central Sierr. versity OF California AntTopolseial Records 3-468. AMATO, I. 1992. From “Hunter Magic,” rmacopoeia? Science 258: 1306. Smithsonian Institution, Washington, BLACKBURN, THOMAS C. 1974. Cere- ane Integration I 110 in ‘Antap: Califo Pharmacop ANDERTON, ALICE JEANNE. 1988. The Language of the Kitanemuks of Califor- nia. Ph.D. Dissertation, University of California, Los Angeles. BEAN, LOWELL JOHN and SYLVIA BRAKKE VANE. 1978. Cults and their Transformations. Pp. 662-672 in Hand- book of North American Indians? Vol. mayer horn (Robert E Heizer, ed.). cal and Economic Basen (Lowell J. Bean and Thomas F. King, eds.). Bal- lena Press: Ramona California. BRITTON, E. B. 1984. A Pointer to a New Specs of Insect Origin. Journal ates 12:331-333. oh ROBERT L. 1970. Hunting and ting Magic Among the Amahuaca. peti 9:331-341. i { Winter 2001 DALY, JOHN W. 1994. Dietary Source for Skin Alkaloids of Poison Frogs (Den- drobatidae). Journal of Chemical Ecol- ogy 20:943-955. DRIVER, HAROLD E. 1937. Culture Ele- ment Distributions: VI Southern Sierra Nevada. University of California An- thropological Records 1:53~154. DU BOIS, CONSTANCE (GODDARD). 1908. The Religion of the Luisefio and Dieguenio Indians of Southern Califor- nia. University of California Publica- tions in American Archaeology and Ethnology 8:69-186. GROARK, KEVIN P. 1996. Ritual and Ther- apeutic Use of ‘“Hallucinogenic” Har- vester Ants (Pogonomyrmex) in Native South-Central California. Journal of Ethnobiology 16(1):1-29. TON, JOHN PEABODY. 1933. Chinigchinich (chi-fti’ch-ftich): A Re- vised and Annotated Version of Alfred Robinson‘s Translation of the Belief, Us- ages, Customs and Extravagencies [sic.] of the Indians of this Mission of San = School and Junior College. (Reprinted as facsimile edition in 1976 sons Capistrano Indians of Southern alifornia (With Two Plates). Smithson- > Miscellaneous Collections 94(4):1- Pe. 1986a. John P. Harrington Papers, ay 2 (Northern and Central Califor- lia). Washington: Smithsonian Institu- hk National Anthropological Ar- [Microfilm editi Publications, Millwood, NY]. , TRAVIS (editor). 1979. Breath of JOURNAL OF ETHNOBIOLOGY 143 the Sun: Life in Early California as Told by a Chumash Indian, Fernando Libra- do, to John P. Harrington. Malki Muse- um Press, Morongo Indian Reservation, anning, CA. JOHNSTON, BERNICE EASTMAN. 1962. California’s Gabrielino Indians. South- west Museum, Los Angeles. KROEBER, ALFRED. 1925. Handbook of the Indians of California. Bureau of American Ethnology Bulletin 78. Wash- ington D.C. OTT, JONATHAN. 1993. Pharmacotheon: Entheogenic Drugs, Their Plant Sources and History. Natural Products Co., Ken- newick, WA. REICHLEN, HENRY and PAUL REI- CHLEN. 1971. Le Manuscrit Boscana de la Bibliothéque Nationale de Paris: Re- lation sur les Indiens Acagchemem de la Mission de San Juan Capistrano, Cal- ifornie. Journal de la Société des Amér- icanistes LX: 233-273. REID, HUGO. 1968 [1852]. The Indians of Los Angeles County: Hugo Reid’s Let- ters of 1852 (edited and annotated by Robert E Heizer). Southwest Museum Papers 21. Los Angeles: Southwest Mu- seum. ROBINSON, ALFRED. 1846. Life in Cali- fornia. Wiley & Putnam, New York [Boscana’s text appears as an appendix, hg lege: : SAINT-HILAIRE, A.E.C.P. de. 1824. Histo- ire du Plantes les plus Remarquables tanica 10:24-61] SCHMIDT, JUSTIN O. and MURRAY S. BLUM. 1978a. A Harvester Ant Venom: Chemistry and Pharmacology. Science 200:1064—1066. _ 1978b. Pharmacological and Toxi- cological Properties of Harvester Ant, Pogonomyrmex badius, Venom. Toxi- con 16: 645-651. _ 1978c. The Biochemical Constitu- ents of the Venom of the Harvester Ant, Pogonomyrmex_badius. Comparative Biochemistry and Physiology 61C: 239- 247. SCHULTZ, D.R. and Pl. ARNOLD. 1978. Ant Venom (Pseudomyrmex sp.) as an ac- tivator of C1 and an inactivator of C3b Inactivator: Its Use in Rheumatoid Ar- thritis. Pp. 172-186 in Clinical Aspects 144 GROARK of the af ommgae sears (W. Opferk- K. Rother and D. R. Schultz, eds.). PS | Pub biicheak, Massachusetts. SPARKMAN, PHILIP S. 1908. The Cultu of the Luisefio Indians. University of California Publications in porn Ar- chaeology and Ethnology 8: VOEGELIN, ERMINIE W. W838. Tabetula- . University of Califor- nia Anthropological Records 2:1-84. WALKER, PHILLIP L. and TRAVIS HUD- 993. Chumash Healing: Chang- ing Health and Medical Practices in an American Indian Society. Malki Muse- um Press, Morongo Indian Reservation, Banning, CA. WEIL, ANDREW T. and WADE DAVIS. 1994. Bufo aluarius: A Potent Hallucino- gen of Animal Origin. Journal of Eth- nopharmacology ng 1-8. WHITE, RAYMOND C. 1963. Luisefio So- cial Organization. Univesity of Califor- nia Publications in American Archae- ology and Ethnology 48(2): 1-194. Uni- versity of California Press, Berkeley, ZIGMOND, MAURICE. 1977. The Super- natural World of the Kawaiisu. Pp. 59- 95 in Flowers of the Wind: Papers on _ nee and Symbolism in Califor- nia and the Southwest (T. C. gga adie, Ballena Press, Socorro, Vol. 21, No. 2 a Winter 2001 ERRATUM. Editor’s Note: This appendix to the artich was inadvertently omitted from vol. 21(1) [Summer JOURNAL OF ETHNOBIOLOGY 145 e by Glenn Shepard, Jr. et al. 2001] and is here reproduced. APPENDIX.—PLANT SPECIES AUTHORS AND VOUCHERS Plant Species Alibertia pilosa Apuleia leiocarpa Astrocaryum murumuru Attalea butyracea Attalea phalerata Attalea tesmannii Bactris concinna Calliandra amazonica cropia membranacea “sles polystachya cropia scia lla Cedrela gg Cedrelinga cataeniformis Ceiba pentandra Chrysochlamys cf. ulei Clavija cf. longifolia Cordia nodosa Davilla nitida Gallesia integrifolia Geonoma brongniartii onoma deversa a maxima Guadua angustifolia Guadua glomerata Guadua weberbaueri Guazuma crinita Gynerium saccharoides in a brasiliensis nartea deltoidea Raephas macrocarpa 3 almia breviscapa “i cf, herzogii ratea exorrhiza Author Krause (Vog.) Macbr. Mart. (Mutis ex L. f.) Wess. Boer Mart. ex Spreng. Burret L.f. (Spreng.) Harms (Poit.) Kunth (Poit.) Kunth Kunth Munro Pilger Mart. Bonpl. (Aubl.) Beauv. Pl. & Linden Muell. Arg. (R. & P) Wedd. Lf. Mart. R. &P. King R. & P. HBK P&E. Voucher* GHS 918 GHS 424 GHS 3345 MIT 351 MIT 8 GHS 1218 GHS 964 GHS 1223 GHS 890 GHS 1346 GHS 1113 GHS 897 GHS 695 GHS 1325 GHS 333 Matsigenka Name matsityanana toaro tiroti shevo tsigaro kovanti setiko / inkona tonko yaaro santari / santaviri paria shirigari / pasaro kachopitoki piamentsishi / pakitsashi matyagiroki tsororoapini shitiro memerishi | metakishi tsikeroshi / choginashi tyonkinto / chigeroshi manipi yaivero onto konori | konori kamona niapashi koshi / toturoki paroto sega kompiro porenki shimashiri Le vakirintsi | kontirt kompapari | konkapart Specimen collector codes: GHS—G.H. Shepard Jr; DWY—D.W. Yu; MIT—M. Italiano. 146 ERRATUM Vol. 21, No. 2 ERRATUM. Editor’s Note: This table to the article by Glenn Shepard, Jr. et al. was inadvertently omitted from vol. 21(1) [Summer 2001] and is here reproduced. TABLE 13.—Correspondence between abiotically (Tables 1-4) and _biotically-defined habitats (Tables 5-12). Widespread soil types S1-S3 (Table 3) have been omitted. Stud sites: T—Tayakome; Y.—Yomybato; M.—Mayapo/. Huallana; C.—Camana. (+) habitat occurs in vicinity; (*) does not occur in vicinity, but is known to occur at a distance. Topography / Hydrology Floodplain—Table 1 Uplands—Table 2 Ref'Ti oT? TS" Tes" 16 17 Te TS TIO Te ra 722 115 Tl4 Tis Tie Ti Ts Palms Table 5 ~ ~ > x ~ KK XK Se ea ~ ~*~ K j= No ~*~ >< a es ~ x eee BR KK KK mK KOK OX ~*~ ~< — ie) ~ > ~*~ xX S x ~*~ *K ~*~ > ~*~ >< ~~ >< ~ eR A a 9 Pe PR eS ON Pt ae EU oP aS > Ferns & Herbs Table 7 25 N ron *< ~ XS ~ XS ~K ~ x ~*~ x ~*~ ~ ss x » x x x > Trees & Shrubs Table 8 37 Xx 38 ~*~ ~*~ an 147 JOURNAL OF ETHNOBIOLOGY Winter 2001 TABLE 13 (extended) Soils Table 4 Disturbance Study sites Table 3 mee 15 14 DS D6 D7 DB S455 S6°S7 SB OU SIO UCU Me +++ + ++4++ 4444+ +++ + ++4++ 4444 t++tteett+ © ++ F444 t++t++4+4+ * ++ + +4 ~< x MK KK xX Pe Pe AG OS eon eS RS Pom SS eh db Se ete tt + ++ +4+4+4+4+4+4+4++4+ ++t+tt¢e¢¢4¢4¢4+4+ ++ ++4+4+4+4+4+4++ ~< xX xX < ~ 8 x be oe >< * >< * >< Montafia Table 10 58 59 60 61 62 63 Vegetation Aspect Table 11 ~*~ ~*~ X ~ >< ~< ~ >< ete x~ KK ~< >< ~ > ~*~ X< ~*~ > ~< >< KK OK OK OX ~ >< ee a a ni ia eaten 149 JOURNAL OF ETHNOBIOLOGY Winter 2001 TABLE 13 (extended, continued) Soils Disturbance Study sites Table 4 Table 3 mer >. 14 D5.D6 D7 DS. S4: SS SO. S7 S859 SIO Tt: YY M C t+++++4++ 44 ++++ ++ +4 t++tt++ +++ 444+ +++ ++++4+4 < p< APs eS See x ~< x< x< < 2 on oe telat t++++4+4++ t++tte++4+ ++++4+4+4+ t+++++4++ xxx x x< >< >< “* * & ee Hh # * + + * &- x XK +++% * +++* * ++++* +++ ees Po ee ea x xX xX xX +++4+4 x re xx x x < <>< mM