Journal of Archaeological Science: Reports 25 (2019) 85–93

Contents lists available at ScienceDirect

Journal of Archaeological Science: Reports

journal homepage: www.elsevier.com/locate/jasrep

Analysis of a coprolite from Conejo Shelter, : Potential ritualistic T viperous snake consumption ⁎ Elanor M. Sondermana, , Crystal A. Dozierb, Morgan F. Smithc a Texas A&M University, USA b Department of Anthropology, Wichita State University, Wichita, KS, USA c Department of Anthropology, Texas A&M University, College Station, TX, USA

ABSTRACT

This paper presents an analysis of the floral and faunal remains of a single human coprolite recovered from Conejo Shelter, Texas (41VV162). The unique contentsof this specimen warrant full description. Floral macrobotanical analysis revealed a high density of Agave lechuguilla and Dasylirion spp. fibers. Calcium oxalate crystals confirm the ingestion of Opuntia. Palynological analysis found evidence for a variety of plants with known economic and medicinal uses, with pollen from the Liliaceae (new: Asparagales) family predominating. Zooarchaeological analysis found the remains of a small rodent, evidently eaten whole, with no indication of preparation or cooking. Notably, the bones, scales and a fang of a snake in the Viperidae family were also recovered from the coprolite, which is the first direct archaeological evidence of venomous snake consumption known to the researchers. With the exception of the Viperidae remains, the coprolite evidence is consistent with previous research at Conejo Shelter and the Lower Pecos region. Recently assayed radiocarbon samples from this coprolite and a second, unprocessed coprolite from the same archaeological provenience produced a date range of 1460–1528 cal BP. Future analyses of coprolites from this lens and the surrounding contexts will further our current understanding of this unique gastrological event and better situate it in the context of diet patterns and paleoenvironmental adaptions in the Lower Pecos.

1. Introduction Grande, the Lower Pecos extends throughout the state of Coahuila. On the Texas side, the Lower Pecos includes all of Val Verde County and Here we present the findings from the analysis of a single coprolite some of Crocket County, to the north. This eco-cultural region has been recovered from Late Archaic period deposits (stratigraphic Lens 9) in occupied by humans for much of the last 12–14,000 years. The cultural Conejo Shelter (41VV162), a dry rockshelter in the Lower Pecos, TX. landscape of the region is defined by both rockshelter and surface ha- The majority of the identified faunal, macrobotanical, and micro- bitations, extensive burned rock middens, a robust woven plant-fiber botanical remains are consistent with other dietary analyses in the re- industry, and elaborate rock art. gion (Bryant, 1974a; Williams-Dean, 1978; Sobolik, 1988a). However, During the Pleistocene/Holocene transition, this landscape sup- in this coprolite we also identified the fang of a viperous snake, likely ported mesic vegetation, including the extension of grasslands from the Crotalus atrox (western diamondback rattlesnake) or Agkistrodon con- Southern Plains (Bryant and Holloway, 1985). Scattered sites in the tortrix (copperhead). We argue that the consumption of a poisonous region, particularly Bonfire Shelter (41VV218) and Cueva Quebrada snake head is anomalous both in the archaeological and ethnographic (41VV162a) provide evidence for some large bodied fauna – bison, records, and considering the dangerous nature of such an action, is horse and camel (Dibble and Lorrain, 1968; Lundelius Jr and Ernest, potentially interpreted as a ritualistic event. Snakes hold important 2003; Turpin, 2004). Progressive and largely uninterrupted aridifica- symbolic status within the region, as in many areas of the world, and tion of the Lower Pecos throughout the Early and Middle Archaic per- the consumption of a venomous snake in the manner indicated in this iods (~9000 to 4500 years BP) limited the available plant and animal coprolite (eaten whole, with no preparation), reflects ritualistic con- resources to xeric-adapted species (Story and Bryant, 1966; Williams- sumption behavior among the foraging peoples of the lower Pecos. Dean, 1978; Reinhard et al., 2007; Riley, 2012). Small rodents, lago- morphs, fish and reptiles, and several species of desert succulents were 1.1. Lifeways in the Lower Pecos Canyonlands the primary food resources for inhabitants of the Lower Pecos (Bryant, 1974b; Williams-Dean, 1978; Stock, 1983; Reinhard et al., 2007; Riley, The Lower Pecos Canyonlands encompass approximately 2008; Jurgens, 2008). Indigenous populations might have encountered 8000 mile2 of southwest Texas and Northern . South of the Rio deer occasionally, but the faunal record shows that this was a relatively

⁎ Corresponding author at: 516 8th St. NE, Washington, DC 20002, USA. E-mail address: [email protected] (E.M. Sonderman). https://doi.org/10.1016/j.jasrep.2019.03.032 Received 8 November 2018; Received in revised form 20 March 2019; Accepted 25 March 2019 Available online 13 April 2019 2352-409X/ © 2019 Published by Elsevier Ltd. E.M. Sonderman, et al. Journal of Archaeological Science: Reports 25 (2019) 85–93 rare occurrence (Alexander, 1974). sample of 43 human coprolites from Conejo Shelter was analyzed While the indigenous inhabitants of the Lower Pecos were primarily during the time of excavation (Bryant, 1974a). The pollen and macro- migratory foragers, their ranges were likely limited by the availability fossil results of that analysis suggest reliance on desert succulents as a of water (Taylor, 1964; Taylor et al., 2003). Many of the canyons in the food source, particularly the flowers of Yucca, Agave, and Dasylirion. Lower Pecos are only watered in times of heavy rainfall, but the three Pollen data from the majority of the coprolites analyzed support a main rivers in this area, the Rio Grande, Pecos, and Devils, would have spring or summer seasonal occupation of Conejo (Bryant, 1974a). Fol- been perennial sources of fresh water. Natural springs and bedrock lowing excavation and these initial analyses, the artifact and coprolite basins called tinajas would also have been readily available sources of materials from this site have been stored at the Texas Archaeological fresh water. Foraging ranges and group sizes were likely controlled by Research Lab at the University of Texas, Austin and the Anthropology the availability of desert succulents, particularly members of Dasylirion, Research Collections at Texas A&M University. Yucca, and Agave, and the wood required to cook them (Dering, 1999). Inhabitants of the area relied on these plant resources extensively, not only for food, but also as raw material for baskets, mats, and sandals 2.2. Coprolite processing (Shafer and Bryant, 1975; Alexander, 1974; Andrews et al., 1980; Stock, 1983; Dering, 1999; Riley, 2012). of the Lower Pecos has Analysis of this coprolite was completed as part of a larger project to long-relied on both traditional examination of fiber, bone, and lithic re-analyze the excavated components of Conejo Shelter (Sonderman, artifacts as well as coprolite analysis in order to reconstruct the lifeways 2018). This coprolite was processed following Bryant (1974b) and of the early inhabitants of this area. Williams-Dean (1978). A photograph of the coprolite taken prior to analysis is shown in Fig. 2. The specimen was first bisected along its 1.2. Coprolite analysis longest axis. One half was selected, photographed, weighed, and as- signed a color from the Munsell color system – 7.5 YR 6/4. Next, the

While the term “coprolite” was originally attributed to mineralized specimen was rehydrated in a bath of trisodium phosphate (Na3PO4) for animal fecal matter, it has been more commonly applied to desiccated a period of two weeks. During the rehydration process, the liquid or frozen human or animal feces, often from archaeological contexts portion of the coprolite turned black, frequently interpreted as an in- (Heizer and Napton, 1969; Williams-Dean, 1978). The utility of co- dicator of human origin (Reinhard and Bryant, 1992; Fry, 1977). We prolite analysis for the reconstruction of human diets has long been washed the rehydrated specimen through nested sieves (1.4 mm > recognized in this region (Fry and Hall, 1975; Fry, 1977; Alexander, 590 μm > 125 μm) with EtOH (ethyl alcohol or ethanol) into a final 1970; Bryant, 1974b; Williams-Dean, 1978; Stock, 1983; Sobolik, collection plate. We collected the contents of each fraction separately in 1991a; Riley, 2008; Tito et al., 2011). specimen jars and briefly stored with EtOH to prevent bacterial growth during analysis. After size sorting, the < 125 μm fraction was processed 2. Materials and methods for plant microfossils. Botanical and faunal components of the larger fractions were separated for identification. 2.1. Excavations at Conejo Shelter

Conejo Shelter (41VV162) was excavated from 1967 to 1968 as part 2.3. Microfossil analysis of the archaeological mitigation efforts prior to the inundation of Amistad Reservoir (Graham and Davis, 1958; Collins, 1969; Alexander, Pollen analysis from the coprolite followed standard procedures 1974). The shelter is situated in a side canyon just north of the con- adapted from Williams-Dean (1978) and follows all standard proce- fluence of the Rio Grande and Pecos Rivers (Fig. 1). It is 170 ft wide, dures for the Palynological Research Laboratory at Texas A&M Uni- approximately 50 ft deep and the overhang is 40 ft high at its highest versity. The microfossil sample (< 125 μm fraction) was centrifuged at point. The base of the shelter is high up the sloping canyon wall, about 3500 rpm for 2 min to concentrate microremains. The supernatant li- 1200 ft above sea level. This elevation not only places the site out of quid was removed and the sample was placed in a 5% potassium hy- reach of flood waters but also largely isolates it from potential entryof droxide bath for 10 min in a heating block at 80 °C. The sample was predators (Alexander, 1974). Robert Alexander, the principal in- then washed with water, ethyl alcohol, and glacial acetic acid before vestigator of the project, recognized the importance of coprolite re- undergoing Erdtman's acetolysis, a solution of acetic anhydride and search and promoted coprolite studies when contemporaneous ex- sulphuric acid (in a 9:1 ratio, respectively) in a heating block at 80 °C cavations routinely ignored them (Bryant, 1987; Sobolik, 1991a). To for 10 min. The sample was then washed in glacial acetic acid, and then prove their worth, Alexander systematically collected coprolites during water until clear. Because phytoliths or other non-pollen palynomorphs the excavation of Conejo Shelter. In doing so, Alexander's work at may provide additional information into dietary patterns, the sample Conejo was the first shelter excavated as part of the Amistad Reservoir then underwent heavy density separation with zinc bromide to isolate project to include the systematic collection of coprolites as a primary the pollen from other microparticles. The heavy fraction was kept for excavation and research goal. Alexander completed the analysis of the future study. The sample was then washed and stained in water and macrobotanical and faunal component of the site to develop broad diet washed again in ethyl alcohol. Stained pollen grains were mounted on a reconstructions (Alexander, 1974). Just over 1000 presumed-human slide with glycerin. coprolites were recovered during excavation. The majority of these The pollen slide was analyzed under light microscopy. Several specimens were concentrated in multiple stratigraphic lenses within the transects of the slide were inspected and over 200 pollen grains were unit designated “N95 W100”. This unit is located approximately 20 ft counted (Barkley, 1934), including unknown and degraded specimens. from the dripline of the shelter. Excavators identified cultural Lens 9 Identification of consumed pollen sources, rather than background only within this unit (Alexander, 1969). environmental pollen, is reliant on pollen concentration (Reinhard The stratigraphic levels directly underlying Lens 9 (especially Lens et al., 1991). While pollen counts over 200 can increase the diversity of 12) were noted as being particularly dense in coprolites (Alexander, taxa identified, increasing pollen counts do not significantly alterthe 1969). Because there is such a high concentration of coprolites within pollen concentration profiles (Jones and Bryant, 1998); therefore, a 200 this and the adjacent areas of the surrounding units, we propose in- count was sufficient to capture the pollen concentration values tode- terpreting this as a potential latrine area. Latrines have been observed termine direct consumption. Grains were identified to taxonomical fa- at other rockshelters in the region, notably at Hinds (shown in mily or genus; identifications were compared to reference material held Fig. 1) and (Shafer and Bryant, 1975; Sobolik, 1991b). A at the Palynological Research Laboratory at Texas A&M University.

86 E.M. Sonderman, et al. Journal of Archaeological Science: Reports 25 (2019) 85–93

Fig. 1. Location of Conejo Shelter and Hinds Cave.

required to confirm identifications. Phytoliths were compared tore- ferences generated during analyses of coprolites from Hinds Cave (lo- cation shown on the map in Fig. 1) and modern samples of Opuntia (Jones and Bryant, 1992; Riley, 2012). This specimen contained a great deal of mammal hair, which had become entangled with the macrobotanical remains. Separation of these material types was a challenge, so the hair was identified in tandem with the macrobotanical remains. Hair was compared to the FBI's Microscopy of Hair Part II: A Practical Guide and Manual for Animal Hairs (Deedrick and Koch, 2004).

2.5. Faunal analysis

As with the macrobotanical material, the faunal remains were col- lected from the largest fraction sieve (1.4 mm). This material was rinsed Fig. 2. Coprolite prior to processing. in EtOH and stored in a series of Petri dishes during analysis. Initial sorting of the material by skeletal element was completed to ease 2.4. Macrobotanical analysis identifications. Bones were examined under magnification with aste- reoscope. Elements were compared to two reference collections: the The macrobotanical remains were collected from the largest fraction Texas A&M Zooarchaeology Laboratory Collection and the Texas A&M plate (1.4 mm). After the faunal component of this fraction was re- Biodiversity and Research Collection. moved, groups of specimens were examined under a dissecting scope for initial sorting. Specimens that could be readily identified at this 2.6. Radiocarbon analysis magnification were separated and grouped by taxa. Several ofthe succulent species ubiquitous to the Lower Pecos look quite similar even During analysis, a small fragment of a wild onion bulb (Allium spp.) under magnification. For these taxa (Yucca spp., Dasylirion spp., Agave was collected and sent to the WM Keck Carbon Cycle Accelerator Mass spp.) examination of phytoliths under high power magnification was Spectrometry Laboratory at the University of California Irvine for

87 E.M. Sonderman, et al. Journal of Archaeological Science: Reports 25 (2019) 85–93 radiocarbon analysis. Because the coprolite material had been pro- cessed and washed with ethanol, we sampled an unprocessed coprolite from the same unit and lens (N95 W100, Lens 9) as a comparative sample.

2.7. Human author of coprolite

Because of the archaeological context of the shelter and the evi- dence of relatively continuous human occupation, all coprolites re- covered from Conejo Shelter are presumed human. As a result of these contextual conditions, additional steps to verify the human origin of this coprolite were deemed unnecessary at the time of analysis. Fecal biomarker analyses and DNA analyses are well established for coprolite studies, but the excavation and storage conditions of these specimens prohibited these analyses as feasible options. Because the specimens were not excavated and collected under sterile conditions and were not always stored in sterile conditions, it would have been impossible to rule out DNA contributions resultant of contamination. While fecal biomarker analysis does not necessitate sterile conditions, this metho- dology would have required sampling of the coprolite prior to rehy- dration (Sistiaga et al., 2014). Micromorphology has been used to de- termine coprolite author, particularly in cases of controversial timing of human occupations (Goldberg et al., 2009). Like fecal biomarker ana- lysis, this methodology can only be employed with an intact coprolite and was, thus, not possible to carry out post-analysis. The morphology Fig. 3. Pollen concentrations (grains/gram). of the intact coprolite, as well as the contents, suggest that the author of the coprolite must have been a large-bodied omnivore. Within the Lower Pecos, only humans both consume the diversity of foods found in the coprolite and are large enough to produce excretions of the size of the studied coprolite. The archaeological context of the coprolite also suggests human authorship. More than 1,000 coprolites were recovered from the main excavation block, which was a 20 by 25 ft rectangle with cultural de- posits extending to about 80 in. below the surface. As noted above, based on the density of coprolites recovered from the shelter, and precedent for the presence of latrines in similar occupation areas, we propose that a portion of the excavated area was used as a latrine, at least occasionally, by the inhabitants of the shelter. The coprolites re- covered from Lens 9 are closely associated with Lens 12, which, in unit N95 W100, directly underlies Lens 9. The excavation notes from the 1969 field season indicate a high density of coprolites in Lens12 (Alexander, 1969). The horizontal distribution of this cultural lens centers on the N95 W100 unit. It is possible that Lens 9 may represent the upper portion of the latrine deposits identified in Lens 12.

3. Results

3.1. Microfossils

The pollen concentrations are related in Fig. 3. The pollen concentration for the coprolite exceeded 156,000 grains per gram of coprolite. Preservation of pollen was fairly good; only three grains were degraded beyond recognition and only ten grains uni- dentified to family or genus. The pollen assemblage was dominated by pollen in the Asparagales family, sensu lato, whereas many genera were formerly in the Liliaceae family (Byng et al., 2016), with a pollen concentration of over 100,300 grains per coprolite gram. Identification of the Asparagales pollen to genus was not attempted due to the overlapping size and similar mor- phology under light microscopy. Genera such as Allium, Agave, Dasy- lirion, and Yucca all share similar monocolpate form with overlapping pollen sizes. Both Yucca and Allium have been observed in coprolites from the region before (Bryant, 1974a) and are known food taxa. The Fig. 4. Comparison of sample and Yucca reference pollen. A: sample. B: re- high concentration of such zoophilous taxa indicates direct consump- ference. tion of the pollen or flower. Of the two previously identified taxa, Yucca flowers have been ethnographically recorded as a food source(Bell and

88 E.M. Sonderman, et al. Journal of Archaeological Science: Reports 25 (2019) 85–93

Fig. 5. Comparison of sample and Pistacia reference pollen. A. Sample. B. Reference.

Fig. 6. Agave lechuguilla and Dasylirion spp. phytoliths.

Castetter, 1941; Castetter and Opler, 1936). Examples from the sample which have similar pollen grain morphology. Pistacia has not been as compared to reference pollen can be seen in Fig. 4. identified in other coprolite studies of the region (Bryant, 1974b). The second most populous pollen type was identified from the The third most abundant taxon was only represented by nine pollen Anacardiaceae family as a Pistacia-type. The identification was made in grains of the genus Gilia. At just over 4% of the pollen recovered, the comparison to an Old World species, Pistacia chinensis (see Fig. 5) concentration may indicate direct consumption of entomophilous Gilia though Pistacia (mexicana and texana) does grow in the Lower Pecos (Reinhard et al., 1991). There are numerous species of Gilia in Texas today (Hatch et al., 1990; Simpsson, 1988). There are no ethno- (Tull and Miller, 1991), a small wildflower that blooms in spring and graphically recorded uses for Pistacia in regions adjacent to the Lower summer. There are many known ethnographic uses for the plant across Pecos, so the relatively high concentration (30,000 grains per gram) the continental , from the Navajo to the Shoshone, mostly may be understood in terms of water contamination or unintentional as a medicinal herb (Moerman, 2003). A single Mirabilis pollen was also consumption. The shrub blooms from March to August and is wind- identified, which has several species within the state(Hatch et al., pollinated. The degraded nature of the sample made exact identifica- 1990) and is also known to have been used for medicinal purposes tion difficult; it is possible that this archaeological class contains Pis- (Moerman, 2003). A number of taxa were identified on the slide during tacia and heavily degraded pollen grains from other taxa, such as Ulmus, an examination that followed the initial pollen count; these include

89 E.M. Sonderman, et al. Journal of Archaeological Science: Reports 25 (2019) 85–93

Cirsium, Ephedra, and at least one Selaginella spore (Bryant, 2017, oral Table 1 communication). All taxa have been historically identified within Specimen counts by anatomical unit for Muridae modern plant communities in the region (Story and Bryant, 1966). and Viperidae. Taxa Counts

3.2. Macrobotanicals Muridae Distal Phalanx 2 Upon visual examination of the macrobotanical sample, several Metatarsals 9 difficulties with quantification were encountered. The density and Phalanges 7 fragility of the various plant fibers made complete separation virtually Maxilla 1 Femur 2 impossible. Additionally, the mammal hair contained within the co- Vertebrae 3 prolite had become interwoven with the fibers, particularly the struc- Scapula 2 tural fibers of Opuntia pads. The entire sample was sorted and presence/ Tibia 2 absence data were collected on species represented and which plant Incisor 1 Fibula 1 parts were represented. The macrobotanical assemblage almost ex- clusively comprised of desert succulents (Opuntia spp., Agave lechuguilla, Viperidae Ribs 11 and Dasylirion spp.). These taxa were represented by epidermal frag- Vertebrae 11 ments, vascular bundles, structural fibers (in the case of Opuntia), and Scales 48 numerous phytoliths and calcium oxalate crystals. Some of the epi- Incisor 1 dermal fragments were fairly large (ca. > 1 cm2). Notably, no spines were present on the epidermal tissue of the Opuntia spp., indicating that their removal prior to consumption. This discovery also supports this coprolite represents ingestion of one whole or nearly whole spe- human origin of the coprolite. cimen for both Muridae and Viperidae. None of the bones or scales Stylus and raphide phytoliths of Agave lechuguilla and Dasylirion are identified exhibited clear signs of cooking. Charring was not observed pictured in Fig. 6; reference photos from these taxa can be found in on the bones, scales, or fur. Additionally, none of the bones showed Riley (2012). pitting, which, if present, is indicative of prolonged boiling (Williams- Other economically useful wild plant species were also recovered, Dean, 1978). including epidermal fragments from wild onion (Allium). The sample Identification of the hair preserved in the coprolite supports the from this taxon sent for radiocarbon assay returned a radiocarbon date identification of Muridae. The hair is largely intact, undamaged, and of 1560 ± 15 years BP (UCIAMS 17089). The second radiocarbon shows no evidence of fire or heat damage, indicating little preparation sample (from the unprocessed coprolite) returned a radiocarbon date of or cooking. It is also possible that the Muridae specimen was consumed 1575 ± 15 years BP (UCIAMS 176090) (Sonderman, 2017). by the snake first and ingested by the human individual secondarily. Of the Viperidae remains, 48 scales were identified. The morphology of the scales and the size of the fang are both consistent with scales and fangs 3.3. Faunal remains from Crotalus atrox. The scales themselves are strongly keeled, sug- gesting C. atrox versus the more weakly-keeled scales of A. contortrix The primary taxa identified in this coprolite are members of (Tennant, 2006). The abundance of scales recovered from this coprolite Rodentia and Viperidae. Comparisons between a rodent incisor re- suggests that the snake was also consumed whole, with little to no covered from the coprolite with those of various taxa in Muridae sup- preparation. Further supporting the C. atrox identification is the length port identification to this Family. Skeletal elements from the Viper of the fang recovered (11 mm). This size lies within the normal range specimen were examined and compared to the Texas A&M Biodiversity for rattlesnakes (5–20 mm), but far out of range for copperheads and Research Collection (fang shown in Fig. 7). This examination, in (maximum 7 mm) (Ernst, 1982; Cundall, 2008). concert with SEM imaging of the fang, particularly the hollow venom channel confirms identification to the family Viperidae. Geographic ranges of Viperidae suggest species identification to Crotalus atrox or 4. Discussion Agkistrodon contortrix, though C. atrox is more common (Tennant, 2006). Counts by anatomical unit for Muridae and Viperidae are pre- The contents of this coprolite specimen were generally consistent sented in Table 1. The Muridae and Viperidae remains include portions with other presumed-human coprolites from this region. Based on of both the appendicular and axial skeleton. Element counts indicate studies of nearby rockshelters, particularly Hinds Cave (indicated in Fig. 1) which had a large, well-defined latrine area, human inhabitants of this area primarily exploited various desert succulents and small mammals, lizards and fishes (Bryant, 1974a; Shafer and Bryant, 1975; Williams-Dean, 1978; Stock, 1983; Sobolik, 1991a, 1988b, 1991b). Coprolites from this region tend to be very fibrous (Sobolik, 1991a; Williams-Dean, 1978; Stock, 1983). Many of the available succulents, particularly sotol (Dasylirion spp.) and lechuguilla (Agave lechuguilla) are baked in earth ovens until the nutritional components of them (primarily fructan) are broken down and more easily digested (Dering, 1999; Riley, 2012). The edible portions are stripped from the leaf bases with the teeth, frequently leading to the ingestion of fibers and large pieces of epidermal fragments from these plants. Ethnohistoric records suggest that prickly pear pads (Opuntia spp.) would have been baked in a similar fashion to sotol and lechuguilla to maximize nutrients (Riley, 2012). The recovery of bone and fur elements from coprolites is also quite Fig. 7. Image of the Viperidae fang recovered during coprolite analysis. Note common in this region. Williams-Dean recorded bone in 97% of the venom channel. coprolites she analyzed from Hinds Cave (Williams-Dean, 1978). In

90 E.M. Sonderman, et al. Journal of Archaeological Science: Reports 25 (2019) 85–93 slight contrast, Sobolik observed bone elements in only 53% of the archaeological record has primarily been discussed through the con- coprolites she analyzed from a latrine area in Baker Cave that was more struction of monumental features (Lightfoot et al., 2011; Kidder, 2011) contemporaneous with the Lens 9 deposits at Conejo Shelter. Both Dean or rock art (Boyd, 1998; Boyd, 2003; Shafer, 2013; Shafer, 1977; and Stock observed the frequency with which animal taxa were re- Turpin, 1982). Ritualistic consumption events are more commonly re- presented by relatively large (1–3 cm) fragments of bone within co- cognized at feasting occasions, which are recognized archaeologically prolites from Hinds Cave (Williams-Dean, 1978; Stock, 1983). The size through the presence of luxury foods (van der Veen, 2003) which is and completeness of skeletal elements from various mammal, reptile, gratuitous, whether in terms of exotic origin, quantity, or volume and fish species suggest common, if not preferred, ingestion ofwhole, (Hayden, 1998; Hayden and Adams, 2004; Hayden and Villeneuve, or nearly whole animals (Williams-Dean, 1978; Stock, 1983; Sobolik, 2011). Such ritualistic consumption behaviors are rarely recognized in 1991b). Sobolik observed fur in all of the coprolite samples that con- hunter-gatherer archaeology; the nature of archaeology among small- tained bones, but also recorded the presence of fur and occasionally scale societies rarely provides insight into individual actions, events, or feathers even when bones were absent from coprolite specimens people (Bettinger et al., 2015), thereby limiting the capabilities of ar- (Sobolik, 1991b). While remains from both lizards and snakes (bone chaeologists to recognize ritualistic patterns at such a granular scale. and scales) were observed in coprolites from Hinds Cave, these did not We believe that coprolite analysis provides an avenue to understand include cranial portions or teeth from either family (Williams-Dean, diet and non-dietary consumption patterns that are critically under- 1978; Stock, 1983). studied among small-scale, foraging societies. Pollen analysis from coprolites can provide dietary/consumption, seasonality, and environmental data (Reinhard and Bryant, 1992). The 4.2. Typical versus atypical snake consumption analysis of this specific coprolite from Conejo Shelter can inform onthe former two. Consumption of snakes, even venomous snakes is well documented The pollen concentration from this coprolite, totaling over 156,000 in the ethnographic records of several cultures around the world, par- grains per gram, indicates a direct consumption of flowers or pollen ticularly in the American Southwest and Northern Mexico, and Great from the Asparagales family a few days prior to deposition (Sobolik, Basin (Mackenzie, 1996; Pennington, 1969; Smith, 1974; Krieger, 1988a). The most likely botanical candidate for the pollen is in the 2002). Ethnographies of Ute and Tepehuan groups specifically mention Yucca genus. Ethnographic reports from adjacent regions indicate that common consumption of rattlesnakes. In the case of the Tepehuan of Yucca flowers themselves were eaten by early inhabitants (Bell and Northeastern Mexico, processing of rattlesnakes for food involved re- Castetter, 1941; Castetter and Opler, 1936). Yucca does not flower moval of both the head, rattle, and skin prior to cooking in an olla seasonally, but rather blooms following periods of sufficient rain. This (Pennington, 1969). For certain bands of Ute, rattlesnakes were skinned interpretation of Yucca is consistent with other coprolite studies of the and then roasted on coals (Smith, 1974). Ethnographies also indicate Lower Pecos (Bryant, 1974b). The low concentration of the other eco- that, while snake was not a preferred food, it might be consumed in nomic plants, Gilia and Mirabilis, make it hard to determine if those taxa times of dietary stress. Cabeza de Vaca provides an account of inland were consumed directly or indirectly (in the water source, or non- Karankawa encountered on his journey through Texas in the 1530s flower/pollen parts of the plant). Nonetheless, their presence doesat (Krieger, 2002 p.194). least indicate that the individual from whom the coprolite originated …and their hunger so great that they eat spiders and ant eggs and had access to those medicinal resources. worms and lizards and salamanders and snakes and vipers that kill The substantial number of Pistacia-type pollen are yet too few to the men they bite, and they eat earth and wood and whatever they indicate definite direct consumption and the plant has no known eth- can get and the dung of deer and other things which I refrain from nographic uses. Rather, as a wind-pollinated genus, it is likely that the telling, and I believe truthfully that if in that land were rocks, they pollen was ingested indirectly as part of the water cycle. Because would eat them. They keep the fine bones of the fish that they eat Pistacia blooms only seasonally (March–August), it is likely that this and of the snakes and other things, in order to grind it all later and coprolite represents a spring or summer incident. All other seasonally eat the powder from it. blooming plant taxa represented in the coprolite also fall into this range. Given that accounts of snake consumption as food rather than ritual The macrobotanical sample we analyzed independently confirms involve removal of the skin, usually accompanied by removal of the our palynological analysis. We attempted-species level identification for head and/or rattle, it is significant and atypical that the snake con- Opuntia. However, calcium oxalate (druse) crystals are common to all sumed by the coprolite's author has the scales intact. Opuntia species and are largely indistinguishable from each other The results of an analysis of the contents of nineteen coprolites from (Jones and Bryant, 1992). While exact counts and weights of macro- Conejo Shelter, identified teeth and cranial portions of reptiles botanical specimens were not recorded, visual comparison of groups of (Sonderman, 2017, personal communication). Several analyses of taxa following macrobotanical sorting and identification suggests that nearby shelters including Hinds Cave and Baker Cave have been con- Opuntia pads may have been the primary component of the meal or ducted, which also recognize the importance of lizards and snakes as a meals resulting in this coprolite. Since seeds were not recovered, it is food resource in the Lower Pecos (Williams-Dean, 1978; Stock, 1983; likely that Opuntia consumption was limited to the pads. Given the ease Sobolik, 1988b). However, none of the snake bones or teeth were at- of harvest, their relative abundance (when fruiting), and accounts from tributed to members of Crotalus and no fangs of any member of Vi- the ethnographic record, we can assume that, had prickly pear tunas peridae were recovered in any of the previously published and un- (fruits) been available, they would have been consumed (Riley, 2012). published coprolite studies from Conejo, Hinds, and Baker The absence of seeds from this coprolite further supports a spring or (Williams-Dean, 1978; Stock, 1983; Sobolik, 1988b). Indeed, none of early summer incident, since these plants typically produce fruit in the the authors have encountered any published archaeological evidence of late summer/fall. Other taxa identified from the macrobotanical re- the consumption of the head of a viperous snake. mains further supports currently understood dietary reliance on desert succulents, particularly the heart or central portion of Dasylirion (likely 4.3. Symbolic importance of snakes texanum). Snakes are revered in many world cultures and are frequent char- 4.1. Ritual in the archaeological record acters in indigenous oral histories and mythologies (Attala, 2016). Whether benign or not, snakes are considered to hold power to act upon Recognizing ritualistic events among foraging peoples in the certain elements of the earth. Because of their power and role in various

91 E.M. Sonderman, et al. Journal of Archaeological Science: Reports 25 (2019) 85–93 mythologies, many cultures around the world include snakes as a fea- Declarations of interest ture of ceremonies and rituals. Pueblo tribes of the southwestern United States, particularly the Hopi, hold Snake ceremonies each year to en- None. sure water availability and a successful harvest. In the culminating event of this multi-day ceremony, snake priests and their associates Acknowledgments collect many snakes (including rattlesnakes and other vipers) from each of the four cardinal directions. These snakes are carried by dancers as The authors would like to thank Dr. Vaughn M. Bryant, of the each snake is then passed to the snake priest who holds the snake in his Department of Anthropology at Texas A&M University for his valuable mouth, imparting the snake with messages for snake spirits to bring counsel during the analysis and writing stages of this project. We also water (Lawrence, 1976; Lawrence et al., 1980). The Aztecs revered thank Jack Johnson, Park Archeologist, Amistad National Recreation snakes as water-related spirits as well (Mackenzie, 1996; Looper, 2009). Area and Marybeth Tomka, Collections Manager, Texas Archeological Rattlesnakes, in particular, were considered to be water-confining Research Lab, and the Anthropology Research Collections at Texas A& bringers of drought. These “drought demons” would be slayed (literally M University for providing access to the Conejo Shelter assemblage. The or figuratively) to “release the water and bring the season of drought to Texas A&M Biodiversity and Research Teaching Collection was instru- an end” (Mackenzie, 1996). In the Primeros memoriales, images de- mental in helping identify the viper and other remains in this coprolite. picting ceremonies to the Aztec rain god Tlaloc, show human figures We are also grateful to Dr. David L. Carlson, of the Department of with, what appear to be rattlesnakes, in their mouths (Looper, 2009). Anthropology at Texas A&M University for editorial comments through Presumably by consuming the snakes during this ceremony, the prac- multiple iterations of this manuscript. Finally, thanks go to Robert K. titioners are ritualistically killing the animal in hopes of releasing the Alexander, without whom this work would not have been possible, and rain. The Dresden Codex also depicts the relationship of serpents and the comments of two anonymous reviewers. rain through panels depicting a human or god headed serpent with Funding for radiocarbon analyses included in this research was water pouring from openings in its body (Mackenzie, 1996). granted by the Council for Texas Archeologists Student Research Grant Serpentine figures are also present in many of the rock art panelsin (no Grant number specified when funds were awarded) and the rockshelters in the Lower Pecos region (Boyd, 2003). In a discussion of Summerlee Foundation Texas History Grant (Sponsor Award Number representations of snakes in the rock art of the Lower Pecos, Boyd M1701481). highlights the roles of serpents in the Shamanic ideologies of several prominent cultures of Mesoamerica and the American Southwest (Boyd, References 1996, 2003). The four cultures noted in this discussion, Aztec, Huichol, Yaqui, and Hopi, each feature the snake as being a gateway or barrier Alexander, Robert Kirk, 1969. Excavation Notes. between earth and supernatural realms. Exploration of thematic and Alexander, Robert Kirk, 1970. Archaeological excavations at Parida Cave, Val Verde County, Texas. In: Papers of the Texas Archeological Salvage Project: No. 19. Texas stylistic elements of the rock art of the Lower Pecos has led Boyd to Archeological Salvage Project, Austin. draw a connection between the ancient cultural group(s) present in that Alexander, Robert Kirk, 1974. The Archaeology of Conejo Shelter: A Study of Cultural region and those of north and northwest Mexico (Boyd, 2003; Boyd and Stability at an Archaic Rockshelter Site in Southwestern Texas. Andrews, Rhonda Lynette, Adovasio, James M., Dirkmaat, Dennis, 1980. Perishable Cox, 2016). Recent aDNA analyses support biological affinity between Industries from Hinds Cave, Val Verde County, Texas. University of Pittsburgh, Dept. the ancient inhabitants of the Lower Pecos and present-day indigenous of Anthropology. populations in Mexico (Raff et al., 2018). Attala, Luci, 2016. Digesting “cryptid” snakes: a phenomenological approach to the Studies of altered states from the cognitive neurosciences have mythic and cosmogenic properties of serpent hallucinations. In: Hurn, Samanth (Ed.), Anthropology and Cryptozoology: Exploring Encounters with Mysterious Creatures. identified certain visual experiences common to trance states, regard- Routledge, New York. less of the cultural background of the individual experiencing them Barkley, Fred A., 1934. The statistical theory of pollen analysis. Ecology 15 (3), 283–289. (Boyd, 2003). One such “entopic phenomena” is the visual experience Bell, Willis Harvey, Castetter, Edward Franklin, 1941. The UItilization of Yucca, Sotol, and Beargrass by the aborigines in the American southwest. In: Ethnobiological of zigzag or undulating lines, often interpreted as snakes or serpents by Studies in the American Southwest. The University of New Mexico Bulletin 5(5). pp. the person in the altered state. Boyd suggests that the serpentine figures 1–75. common to so many rock art panels represent these common visions Bettinger, Robert L., Garvey, Raven, Tushingham, Shannon, 2015. Hunter-Gatherers: Archaeological and Evolutionary Theory. Springer. achieved during altered states (Boyd, 2003). In the Lower Pecos, the Boyd, Carolyn E., 1996. Shamanic journeys into the otherworld of the archaic Chichimec. mescal bean (Sophora secundiflora), peyote (Lophophora williamsii), and Lat. Am. Antiq. 7 (2), 152–164. Datura (Datura spp.) are all relatively easily accessed for the purpose of Boyd, Carolyn E., 1998. In: Boyd, Carolyn Elizabeth (Ed.), The Work of Art: Rock Art and Adaptation in the Lower Pecos, Texas Archaic. Texas A&M University. reaching altered states of consciousness (Boyd, 2003). Symbologies of Boyd, Carolyn E., 2003. Rock art of the lower Pecos. In: Texas A&M Anthropology Series: shamanism associated with the use of these plants are all present in the No. 8 Texas A&M University Press, College Station. Pecos River-style rock art, which first appeared approximately Boyd, Carolyn E., Cox, Kim, 2016. The White Shaman Mural: An Enduring Creation Narrative in the Rock Art of the Lower Pecos. University of Texas Press. 4000 years ago (Boyd, 2003; Turpin, 2004; Boyd et al., 2013). While Boyd, Carolyn E., Castaneda, Amanda M., Koenig, Charles W., 2013. A reassessment of neither peyote nor Datura was recovered from Conejo Shelter, pods and red linear pictographs in the Lower Pecos Canyonlands of Texas. Am. Antiq. 78 (3), seeds of Sophora secundiflora were recovered during excavation 456–482. (Alexander, 1974). Bryant, Vaughn M., 1974a. Prehistoric diet in Southwest Texas: the coprolite evidence. Am. Antiq. 39 (3), 407–420. Bryant, Vaughn M., 1974b. The role of coprolite analysis in archeology. Bulletin of the 5. Conclusion Texas Archeological Society 45. Bryant, Vaughn M., 1987. Pollen grains: the tiniest clues in archaeology. Environment Southwest 519. The recovery of Viper skeleton remains from a human coprolite is Bryant, Vaughn M., Holloway, Richard G., 1985. A late-Quaternary paleoenvironmental remarkable, not only because of the rarity of such an occurrence but record of Texas: an overview of the pollen evidence. In: Pollen Record of Late also because this find provides a glimpse of the ritualistic behaviors of Quaternary North American Sediments. American Association of Stratigraphic Palynologists, Dallas, pp. 39–70. small-scale societies. We propose that the ingestion of an entire veno- Byng, James W., Chase, Mark W., Christenhusz, Maarten J.M., Fay, Michael F., Judd, mous snake is not typical behavior for the occupants of the Lower Pecos Walter S., Mabberley, David J., Sennikov, Alexander N., Soltis, Douglas E., Soltis, or Conejo Shelter. It is also clear from ethnographic analogy and rock Pamela S., Stevens, Peter F., 2016. An update of the Angiosperm Phylogeny Group classification for the orders and families of flowering plants: APG IV. Bot. J.Linn.Soc. art from this region that snakes hold ritual significance to the in- 181 (1), 1–20. digenous populations of the Lower Pecos. We propose that a likely Castetter, Edward Franklin, Opler, Morris Edward, 1936. The ethnobiology of the explanation for the ingestion of an entire snake is that the individual Chiricahua and Mescalero apache: A. In: The Use of Plants for Foods, Beverages and Narcotics. vol. 4 University of New Mexico Press. did so for a distinctly ceremonial or ritualistic purpose.

92 E.M. Sonderman, et al. Journal of Archaeological Science: Reports 25 (2019) 85–93

Collins, Michael B., 1969. Test excavations at Amistad International Reservoir, fall, 1967. Raff, Jennifer A., Reynolds, Austin W., Turpin, Solveig, Rohland, Nadin, Reich, David, In: Papers of the Texas Archeological Salvage Project: No. 16. Texas Archeological Bolnick, Deborah A., 2018. Paleogenomic investigations of the ancient inhabitants of Salvage Project, Austin. the Lower Pecos region of Texas and Northern Mexico. In: American Journal of Cundall, David, 2008. Viper fangs: functional limitations of extreme teeth. Physiol. Physical Anthropology 165. Wiley, 111 River St, Hoboken 07030-5774, NJ USA, pp. Biochem. Zool. 82 (1), 63–79. 217–218. Deedrick, Douglas W., Koch, Sandra L., 2004. Microscopy of Hair Part II: A Practical Reinhard, Karl J., Bryant, Vaughn M., 1992. Coprolite analysis: a biological perspective Guide and Manual for Animal Hairs. FBI Forensic Science Communications. on archaeology. Archaeol. Method Theory 4, 245–288. Dering, Phil, 1999. Earth-oven plant processing in archaic period economies: an example Reinhard, Karl J., Hamilton, Donny L., Hevly, Richard H., 1991. Use of pollen con- from a semi-arid Savannah in South-Central . Am. Antiq. 64 (4), centration in paleopharmacology: coprolite evidence of medicinal plants. J. 659–674. Ethnobiol. 11 (1), 117–132. Dibble, David S., Lorrain, Dessamae, 1968. Bonfire shelter: a stratified bison kill site, Val Reinhard, Karl J., Ambler, J.R., Szuter, C.R., 2007. Hunter-gatherer use of small animal Verde County, Texas. In: A Publication of the Texas Memorial Museum. food resources: coprolite evidence. Int. J. Osteoarchaeol. 17, 416–428. Miscellaneous Papers, No. 1. University of Texas, Texas Memorial Museum, Austin. Riley, Timothy E., 2008. Diet and seasonality in the Lower Pecos: evaluating coprolite Ernst, Carl H., 1982. A study of the fangs of snakes belonging to the Agkistrodon-complex. data sets with cluster analysis. J. Archaeol. Sci. 35 (10), 2726–2741. J. Herpetol. 72–80. Riley, Timothy E., 2012. Assessing diet and seasonality in the Lower Pecos canyonlands: Fry, Gary F., 1977. Analysis of prehistoric coprolites from Utah. In: Anthropological an evaluation of coprolite specimens as records of individual dietary decisions. J. Papers; No. 97. University of Utah Press, Salt Lake City. Archaeol. Sci. 39 (1), 145–162. Fry, Gary, Hall, H.J., 1975. Human coprolites from Antelope house: preliminary analysis. Shafer, Harry J., 1977. Art and territoriality in the Lower Pecos archaic. Plains Anthropol. Kiva 41 (1), 87–96. 22 (75), 13–21. Goldberg, Paul, Berna, Francesco, Macphail, RIchard I., 2009. Comment on “DNA from Shafer, Harry J., 2013. In: Shafer, Harry J. (Ed.), Painters in Prehistory: Archaeology and pre-Clovis human coprolites in Oregon, North America”. Science 325, 148. Art of the Lower Pecos Canyonlands. Trinity University Press, San Antonio (published Graham, John A., Davis, W.A., 1958. Appraisal of the Archeological Resources of Diablo in association with the Witte Museum, [2013]). Reservoir Val Verde County, Texas. Shafer, Harry, Bryant, Vaughn M., 1975. In: Shafer, Harry J. (Ed.), A Preliminary Report Hatch, Stephen L., Gandhi Kancheepuram, N., Brown, Larry E., 1990. Checklist of the of Hinds Cave, Val Verde County, Texas. Anthropology Laboratory, Texas A&M Vascular Plants of Texas (MP-1655). Texas Agricultural Experiment Station, College University: no. 8. Anthropology Laboratory, Texas A&M University, College Station Station. [and others]. Technical report. Hayden, Brian, 1998. Practical and prestige technologies: the evolution of material sys- Simpsson, Benny J., 1988. A Field Guide to Texas Trees. Texas Monthly Press, Austin, tems. J. Archaeol. Method Theory 5 (1), 1–55. Texas. Hayden, Bryan, Adams, Ron, 2004. Ritual structures in transegalitarian communities. In: Sistiaga, A., Berna, F., Laursen, R., Goldberg, P., 2014. Steroidal biomarker analysis of a Complex Hunter-Gatherers: Evolution and Organization of Prehistoric Communities 14,000 years old putative human coprolite from Paisley Cave, Oregon. J. Archaeol. on the Plateau of Northwestern North America. University of Utah Press, Salt Lake Sci. 41, 813–817. City, pp. 84–102. Smith, Anne M., 1974. Ethnography of the Northern Utes. In: Papers in Anthropology. Hayden, Brian, Villeneuve, Suzanne, 2011. A century of feasting studies. Annu. Rev. Museum of New Mexico Press, Santa Fe. Anthropol. 40 (1), 433–449. Sobolik, Kristin D., 1988a. The importance of pollen concentration values from coprolites: Heizer, Robert F., Napton, Lewis K., 1969. Biological and cultural evidence from pre- an analysis of Southwest Texas samples. Palynology 12, 201–214. historic human coprolites. Science 165 (3893), 563 (LP-568). Sobolik, Kristin D., 1988b. The Prehistoric Diet and Subsistence of the Lower Pecos Jones, John G., Bryant, Vaughn M., 1992. Phytolith taxonomy in selected species of Texas Region, as Reflected in Coprolites from Baker Cave, Val Verde County, Texas. TexasA cacti. In: Phytolithihic Systematics, pp. 215–238. &M University. Jones, Gretchen D., Bryant, Vaughn M., 1998. Are all counts created equal? In: Bryant Sobolik, Kristin D., 1991a. Paleonutrition in the Lower Pecos Region of the Chihuahuan Jr.Vaughn M., Wrenn, J. (Eds.), New Developments in Palynomorph Sampling, Desert. Texas A&M University. Extinction and Analysis. American Association of Stratigraphic Palynologists Sobolik, Kristin D., 1991b. Prehistoric diet from the Lower Pecos Region of Texas. Plains Foundation, Dallas, TX, pp. 115–120 Contributi. Anthropol. 36 (135), 139–152. Jurgens, Christopher J., 2008. The fish fauna from Arenosa shelter (41VV99), Lower Sonderman, Elanor, 2017. CTA student grant report. In: Council of Texas Archeologists Pecos region, Texas. Quat. Int. 185 (1), 26–33. Newsletter, (March). Kidder, Tristam R., 2011. Transforming hunter-gatherer history at poverty point. In: Sonderman, Elanor, 2018. Re-analysis of Conejo Shelter: A Legacy Collection from the Sassaman, Kenneth E., Jr, Donald H. Holly (Eds.), Hunter-Gatherer Archaeology as Amistad Reservoir Area, Texas. Historical Process. Amerind Studies in Archaeology University of Arizona Press, Stock, Janet Ann, 1983. The Diet of Hinds Cave (41VV456), Val Verde County, Texas: The Tucson, pp. 95–119. Coprolite Evidence. Texas A&M University. Krieger, Alex Dony, 2002. In: Krieger, Margery H. (Ed.), We Came Naked and Barefoot: Story, Dee Ann, Bryant, Vaughn M., 1966. A Preliminary Study of the Paleoecology of the The Journey of Cabeza de Vaca Across North America. Texas Arch. University of Amistad Reservoir Area. (Austin). Texas Press, Austin, Texas. Taylor, Walter W., 1964. Tethered nomadism and water territoriality: an hypothesis. Lawrence, David Herbert, 1976. The Hopi snake dance. Salmagundi 33/34, 133–148. Actas y Memorias 2, 197–203. Lawrence, D.H., Laird, David, Laird, Linda, Sanders, Mark, 1980. The Hopi Snake Dance. Taylor, Walter W., Demerath, Nicholas J., Kennedy, Mary C., Watson, Patty Jo, Quilter, Al Lowman Printing Arts Collection and Research Archive. Peccary Press, Flagstaff. Jeffrey, Browman, David L., Adovasio, James M., 2003. Sandals from Coahuila caves: Lightfoot, Kent G., Luby, Edward M., Pesnichak, Lisa, 2011. Evolutionary typologies and with an introduction to the Coahuila project Coahuila, Mexico: 1937–1941, 1947. In: hunter-gatherer research: rethinking the mounded landscapes of Central California. Studies in Pre-Columbian Art and Archaeology. JSTOR. In: Sassaman, Kenneth E., Holly Jr.Donald H. (Eds.), Hunter-Gatherer Archaeology as Tennant, Alan, 2006. Lone Star Field Guide to Texas Snakes. Taylor Trade Publishing. Historical Process. Amerind Studies in Archaeology University of Arizona Press, Tito, Raul Y., Belknap 3rd, Samuel L., Sobolik, Kristin D., Ingraham, Robert C., Cleeland, Tucson, pp. 55–78. Lauren M., Lewis Jr, Cecil M., 2011. Brief communication: DNA from early Holocene Looper, Matthew G., 2009. To Be like Gods: Dance in Ancient Maya Civilization. American dog. Am. J. Phys. Anthropol. 145 (4), 653–657. University of Texas Press. Tull, Delena, Miller, George Oxford, 1991. Field guide to wildflowers, trees & shrubs of Lundelius Jr., , Ernest, L., 2003. A history of paleontological investigations of Quaternary Texas. In: Texas Fieldguide Series Gulf Publishing Company, Houston, Texas. cave deposits on the Edwards Plateau, central Texas. In: Ice Age Cave Faunas of North Turpin, Solveig A., 1982. The Archeology and Rock Art of Seminole Canyon: A Study in America. Indiana University Press, Bloomington, Indiana. Indiana University Press, the Lower Pecos River Region of Southwest Texas. University of Texas. Bloomington, pp. 201–214. Turpin, Solveig A., 2004. The Lower Pecos River Region of Texas and Northern Mexico. Mackenzie, Donald Alexander, 1996. Myths of Pre-Columbian America. Courier In: Perttula, Timothy K. (Ed.), The Prehistory of Texas. 086. Texas A&M University Corporation. Press, College Station, pp. 266–280. Moerman, Dan, 2003. Native American Ethnobotany Database. Botanic Research Institue van der Veen, Marijke, 2003. When is food a luxury? World Archaeol. 34 (3), 405–427. of Texas. Williams-Dean, Glenna Joyce, 1978. Ethnobotany and Cultural Ecology of Prehistoric Pennington, Campbell W., 1969. The Tepehuan of Chihuahua: Their Material Culture. Man in Southwest Texas. pp. 1–287. University of Utah Press, Salt Lake City.

93