Hostname: page-component-8448b6f56d-mp689 Total loading time: 0 Render date: 2024-04-24T03:26:19.787Z Has data issue: false hasContentIssue false

Enamel hypoplasia and dental wear of North American late Pleistocene horses and bison: an assessment of nutritionally based extinction models

Published online by Cambridge University Press:  03 June 2019

Christina I. Barrón-Ortiz
Affiliation:
Department of Biological Sciences, University of Calgary, Calgary, Alberta T2N 1N4, Canada. E-mail: christina.barron-ortiz@gov.ab.ca
Christopher N. Jass
Affiliation:
Quaternary Palaeontology Program, Royal Alberta Museum, Edmonton, Alberta T5J 0G2, Canada
Raúl Barrón-Corvera
Affiliation:
Programa de Ingeniería Civil, Universidad Autónoma de Zacatecas, Zacatecas 98000, Mexico
Jennifer Austen
Affiliation:
Department of Archaeology, University of Reading, Reading RG6 6AH, United Kingdom
Jessica M. Theodor
Affiliation:
Department of Biological Sciences, University of Calgary, Calgary, Alberta T2N 1N4, Canada. E-mail: christina.barron-ortiz@gov.ab.ca

Abstract

Approximately 50,000–11,000 years ago many species around the world became extinct or were extirpated at a continental scale. The causes of the late Pleistocene extinctions have been extensively debated and continue to be poorly understood. Several extinction models have been proposed, including two nutritionally based extinction models: the coevolutionary disequilibrium and mosaic-nutrient models. These models draw upon the individualistic response of plant species to climate change to present a plausible scenario in which nutritional stress is considered one of the primary causes for the late Pleistocene extinctions.

In this study, we tested predictions of the coevolutionary disequilibrium and mosaic-nutrient extinction models through the study of dental wear and enamel hypoplasia of Equus and Bison from various North American localities. The analysis of the dental wear (microwear and mesowear) of the samples yielded results that are consistent with predictions established for the coevolutionary disequilibrium model, but not for the mosaic-nutrient model. These ungulate species show statistically different dental wear patterns (suggesting dietary resource partitioning) during preglacial and full-glacial time intervals, but not during the postglacial in accordance with predictions of the coevolutionary disequilibrium model. In addition to changes in diet, these ungulates, specifically the equid species, show increased levels of enamel hypoplasia during the postglacial, indicating higher levels of systemic stress, a result that is consistent with the models tested and with other climate-based extinction models. The extent to which the increase in systemic stress was detrimental to equid populations remains to be further investigated, but suggests that environmental changes during the late Pleistocene significantly impacted North American equids.

Type
Articles
Copyright
Copyright © The Paleontological Society. All rights reserved 2019 

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

Footnotes

*

Present address: Quaternary Palaeontology Program, Royal Alberta Museum, Edmonton, Alberta T5J 0G2, Canada.

Data available from the Dryad Digital Repository: https://doi.org/10.5061/dryad.3kf3gg2

References

Literature Cited

Alroy, J. 1999. Putting North America's end-Pleistocene megafaunal extinction in context: large-scale analyses of spatial patterns, extinction rates, and size distributions. Pp. 105143 in MacPhee, R. D. E. and Sues, H., eds. Extinctions in near time: causes, contexts, and consequences. Advances in vertebrate paleobiology. Springer Science + Business Media, New York.Google Scholar
Anderson, P. M., Edwards, M. E., and Brubaker, L. B.. 2003. Results and paleoclimate implications of 35 years of paleoecological research in Alaska. Pp. 427440 in Gillespie, A. E., Porter, S. C., and Atwater, B. F., eds. The Quaternary period in the United States. Developments in Quaternary science. Elsevier, New York.Google Scholar
Azzaroli, A. 1998. The genus Equus in North America—the Pleistocene species. Paleontographia Italica 85:160.Google Scholar
Barnosky, A. D. 1986. “Big Game” extinction caused by late Pleistocene climatic change: Irish elk (Megaloceros giganteus) in Ireland. Quaternary Research 25:128135.Google Scholar
Barrón-Ortiz, C. I., Rodrigues, A. T., Theodor, J. M., Kooyman, B. P., Yang, D. Y., and Speller, C. F.. 2017. Cheek tooth morphology and ancient mitochondrial DNA of late Pleistocene horses from the western interior of North America: implications for the taxonomy of North American late Pleistocene Equus. PLoS ONE 12(8):e0183045.Google Scholar
Barrón-Ortiz, C. I., Jass, C. N., and Bravo-Cuevas, V. M.. 2018. Equus: Where is the genus? Phylogenetic assessment of Haringtonhippus francisci (Perissodactyla, Equidae) and other horses traditionally assigned to Equus. 78th Annual Meeting of the Society of Vertebrate Paleontology, Program and Abstracts, pp. 86–87.Google Scholar
Barrón-Ortiz, C. R., Theodor, J. M., and Arroyo-Cabrales, J.. 2014. Dietary resource partitioning in the Late Pleistocene horses from Cedral, north-central Mexico: evidence from the study of dental wear. Revista Mexicana de Ciencias Geológicas 31:260269.Google Scholar
Bell, R. H. V. 1971. A grazing ecosystem in the Serengeti. Scientific American 224:8693.Google Scholar
Boddaert, P. 1785. Elenchus animalium, volumen 1: Sistens quadrupedia huc usque nota, erorumque varietates. C. R. Hake, Rotterdam.Google Scholar
Braunn, P. R., Ribeiro, A. M., and Ferigolo, J.. 2014. Microstructural defects and enamel hypoplasia in teeth of Toxodon Owen, 1837 from the Pleistocene of Southern Brazil. Lethaia 47:418431.Google Scholar
Bravo-Cuevas, V. M., Jiménez-Hidalgo, E., and Priego-Vargas, J.. 2011. Taxonomía y hábito alimentario de Equus conversidens (Perissodactyla, Equidae) del Pleistocene tardío de Hidalgo, centro de México. Revista Mexicana de Ciencias Geológicas 28:6582.Google Scholar
Burns, J. A. 1996. Vertebrate paleontology and the alleged ice-free corridor: the meat of the matter. Quaternary International 32:107112.Google Scholar
Byerly, R. M. 2007. Palaeopathology in late Pleistocene and early Holocene Central Plains bison: dental enamel hypoplasia, fluoride toxicosis and the archaeological record. Journal of Archaeological Science 34:18471858.Google Scholar
Byerly, R. M. 2009. Late Pleistocene to Holocene variability in bison health: implications for human bison-based subsistence on the Northwestern and Central Great Plains. PhD dissertation. Southern Methodist University, Dallas, Tex.Google Scholar
Canadian Archaeological Radiocarbon Database. CARD 2.0 home page. http://www.canadianarchaeology.ca, accessed 17 March 2015.Google Scholar
Cinq-Mars, J. 1979. Bluefish Cave l: a late Pleistocene eastern Beringian cave deposit in the northern Yukon. Canadian Journal of Archaeology 3:132.Google Scholar
Cinq-Mars, J. 1990. La place des grottes du Poisson–Bleu dans la prehistoire Beringienne. Revista de Arqueología Americana 1: 932.Google Scholar
Cerling, T. E., Harris, J. M., and Passey, B. H.. 2003. Diets of East African Bovidae based on stable isotope analysis. Journal of Mammalogy 84:456470.Google Scholar
Clauss, M., Nunn, C., Fritz, J., and Hummel, J.. 2009. Evidence for a tradeoff between retention time and chewing efficiency in large mammalian herbivores. Comparative Biochemistry and Physiology A 154:376382.Google Scholar
Coltrain, J. B., Harris, J. M., Cerling, T. E., Ehleringer, J. R., Dearing, M.-D., Ward, J., and Allen, J.. 2004. Rancho La Brea stable isotope biogeochemistry and its implications for the palaeoecology of late Pleistocene, coastal southern California. Palaeogeography, Palaeoclimatology, Palaeoecology 205:199219.Google Scholar
Connin, S. L., Betancourt, J., and Quade, J.. 1998. Late Pleistocene C4 plant dominance and summer rainfall in the southwestern United States from isotopic study of herbivore teeth. Quaternary Research 50:179193.Google Scholar
Cooper, A., Turney, C., Hughen, K. A., Brook, B. W., McDonald, H. G., and Bradshaw, C. J. A.. 2015. Abrupt warming events drove Late Pleistocene Holartic megafaunal turnover. Science 349:602606.Google Scholar
Diamond, J. M. 1989. Quaternary megafaunal extinctions: variations on a theme by Paganini. Journal of Archaeological Science 16:167175.Google Scholar
Dobney, K., and Ervynck, A. 2000. Interpreting developmental stress in archaeological pigs: the chronology of linear enamel hypoplasia. Journal of Archaeological Science 27:597607.Google Scholar
Dobney, K., Ervynck, A., Albarella, U., and Rowley-Conwy, P.. 2004. The chronology and frequency of a stress marker (linear enamel hypoplasia) in recent and archaeological populations of Sus scrofa in northwest Europe, and the effects of early domestication. Journal of Zoology 264:197208.Google Scholar
Du Toit, J. T. 1990. Feeding-height stratification among African browsing ruminants. African Journal of Ecology 28:5561.Google Scholar
Emery-Wetherell, M. M., McHorse, B. K., and Davis, E. B.. 2017. Spatially explicit analysis sheds new light on the Pleistocene megafaunal extinction in North America. Paleobiology 43:642655.Google Scholar
Faith, J. T., and Surovell, T. A.. 2009. Synchronous extinction of North America's Pleistocene mammals. Proceedings of the National Academy of Sciences USA 106:2064120645.Google Scholar
Famoso, N. A., Feranec, R. S., and Davis, E. B.. 2013. Occlusal enamel complexity and its implications for lophodonty, hypsodonty, body mass, and diet in extinct and extant ungulates. Palaeogeography, Palaeoclimatology, Palaeoecology 387:211216.Google Scholar
Faunmap Working Group. 1994. FAUNMAP: a database documenting Late Quaternary distributions of mammal species in the United States. Illinois State Museum Scientific Papers 25:1690.Google Scholar
Federation Dentaire Internationale. 1982. An epidemiological index of developmental defects of dental enamel (DDE Index). Technical report 16, International Dental Journal 32:159167.Google Scholar
Feranec, R. S. 2004. Geographic variation in the diet of hypsodont herbivores from the Rancholabrean of Florida. Palaeogeography, Palaeoclimatology, Palaeoecology 207:359369.Google Scholar
Feranec, R. S., Hadly, E. A., and Paytan, A.. 2009. Stable isotopes reveal seasonal competition for resources between late Pleistocene bison (Bison) and horse (Equus) from Rancho La Brea, southern California. Palaeogeography, Palaeoclimatology, Palaeoecology 271:153160.Google Scholar
Ficcarelli, G., Coltorti, M., Moreno-Espinosa, M., Pieruccini, P. L., Rook, L., and Torre, D.. 2003. A model for the Holocene extinction of the mammal megafauna in Ecuador. Journal of South American Earth Sciences 15:835845.Google Scholar
Firestone, R. B., West, A., Kennett, J. P., Becker, L., Bunch, T. E., Revay, Z. S., Schulz, P. H., et al. 2007. Evidence for an extraterrestrial impact 12,900 years ago that contributed to the megafaunal extinctions and the Younger Dryas cooling. Proceedings of the National Academy of Sciences USA 104:1601616021.Google Scholar
Fisher, D. C. 2009. Paleobiology and extinction of proboscideans in the Great Lakes region of North America. Pp. 5575 in Haynes, G., ed. American megafaunal extinctions at the end of the Pleistocene. Springer, Dordrecht, Netherlands.Google Scholar
Forster, M. A. 2004. Self-organised instability and megafaunal extinctions in Australia. Oikos 103:235239.Google Scholar
Fortelius, M., and Solounias, N.. 2000. Functional characterization of ungulate molars using the abrasion-attrition wear gradient: a new method for reconstructing paleodiets. American Museum Novitates 3301:136.Google Scholar
Fox-Dobbs, K., Leonard, J. A., and Koch, P. L.. 2008. Pleistocene megafauna from eastern Beringia: paleoecological and paleoenvironmental interpretations of stable carbon and nitrogen isotope and radiocarbon records. Palaeogeography, Palaeoclimatology, Palaeoecology 261:3046Google Scholar
Franz-Odendaal, T. A. 2004. Enamel hypoplasia provides insights into early systemic stress in wild and captive giraffes (Giraffa camelopardalis). Journal of Zoology 263:197206.Google Scholar
Franz-Odendaal, T. A., and Kaiser, T. M.. 2003. Differential mesowear in the maxillary and mandibular cheek dentition of some ruminants (Artiodactyla). Annales Zoologici Fennici 40:395410.Google Scholar
Franz-Odendaal, T. A., Chinsamy, A., and Lee-Thorp, J.. 2004. High prevalence of enamel hypoplasia in an early Pliocene giraffid (Sivatherium hendeyi) from South Africa. Journal of Vertebrate Paleontology 24:235244.Google Scholar
Fraser, D., Mallon, J.C., Furr, R., and Theodor, J. M.. 2009. Improving the repeatability of low magnification microwear methods using high dynamic range imaging. Palaios 24:818825.Google Scholar
Fritz, J., Hummel, J., Kienzle, E., Arnold, C., Nunn, C. L. and Clauss, M.. 2009. Comparative chewing efficiency in mammalian herbivores. Oikos 118:16231632.Google Scholar
Gadbury, C., Todd, L., Jahren, A. H., and Amundson, R.. 2000. Spatial and temporal variations in the isotopic composition of bison tooth enamel from the Early Holocene Hudson–Meng Bone bed, Nebraska. Palaeogeography, Palaeoclimatology, Palaeoecology 157:7993.Google Scholar
Gentry, A., Clutton-Brock, J., and Groves, C. P.. 1996. Proposed conservation of usage of 15 mammal specific names based on wild species which are antedated by or contemporary with those based on domestic animals. Bulletin of Zoological Nomenclature 53:2833.Google Scholar
Gentry, A., Clutton-Brock, J., and Groves, C. P.. 2004. The naming of wild animal species and their domestic derivatives. Journal of Archaeological Science 31:645651.Google Scholar
Gomes Rodrigues, H., Merceron, G., and Viriot, L.. 2009. Dental microwear patterns of extant and extinct Muridae (Rodentia, Mammalia): ecological implications. Naturwissenschaften 96:537542.Google Scholar
Goodman, A. H., and Rose, J. C.. 1990. Assessment of systemic physiological perturbations from dental enamel hypoplasias and associated histological structures. Yearbook of Physical Anthropology 33:59110.Google Scholar
Goodman, A. H., Armelagos, G. J., and Rose, J. C.. 1980. Enamel hypoplasias as indicators of stress in three prehistoric populations from Illinois. Human Biology 3:515528.Google Scholar
Graham, R. W., and Lundelius, E. L. Jr. 1984. Coevolutionary disequilibrium and Pleistocene extinctions. Pp. 223249 in Martin, P. S. and Klein, R. G., eds. Quaternary extinctions: a prehistoric revolution. University of Arizona Press, Tucson.Google Scholar
Graham, R. W., Lundelius, E. L. Jr., Graham, M. A., Schroeder, E. K., Toomey, R. S. III, Anderson, E., Barnosky, A. D., et al. 1996. Spatial response of mammals to late Quaternary environmental fluctuations. Science. 272:16011606.Google Scholar
Grayson, D. K. 1991. Late Pleistocene mammalian extinctions in North America: taxonomy, chronology, and explanations. Journal of World Prehistory 5:193231.Google Scholar
Grayson, D. K. 2007. Deciphering North American Pleistocene extinctions. Journal of Anthropological Research 63:185214.Google Scholar
Grayson, D. K. 2016. Giant sloths and sabertooth cats: extinct mammals and the archaeology of the Ice Age Great Basin. University of Utah Press, Salt Lake City.Google Scholar
Guatelli-Steinberg, D. 2000. Linear enamel hypoplasia in gibbons (Hylobates larcarpenteri). American Journal of Physical Anthropology 112:395410.Google Scholar
Guatelli-Steinberg, D. 2003. Macroscopic and microscopic analysis of linear enamel hypoplasia in Plio-Pleistocene South-African hominins with respect to aspects of enamel development and morphology. American Journal of Physical Anthropology 120:309322.Google Scholar
Guatelli-Steinberg, D., and Benderlioglu, Z.. 2006. Brief communication: linear enamel hypoplasia and the shift from irregular to regular provisioning in Cayo Santiago rhesus monkeys (Macaca mulatta). American Journal of Physical Anthropology 131:416419.Google Scholar
Guatelli-Steinberg, D., Ferrell, R. J., and Spence, J.. 2012. Linear enamel hypoplasia as an indicator of physiological stress in great apes: reviewing the evidence in light of enamel growth variation. American Journal of Physical Anthropology 148:191204.Google Scholar
Guatelli-Steinberg, D. J. 1998. Prevalence and etiology of linear enamel hypoplasia in non-human primates. PhD dissertation. University of Oregon, Eugene.Google Scholar
Guthrie, R. D. 1984. Mosaics, allelochemics and nutrients: an ecological theory of late Pleistocene megafaunal extinctions. Pp. 259298 in Martin, P. S. and Klein, R. G., eds. Quaternary extinctions: a prehistoric revolution. University of Arizona Press, Tucson.Google Scholar
Guthrie, R. D. 2006. New carbon dates link climatic change with human colonization and Pleistocene extinctions. Nature 44:207209.Google Scholar
Gwynne, M. D., and Bell, R. H. V.. 1968. Selection of vegetation components by grazing ungulates in the Serengeti National Park. Nature 220:390393.Google Scholar
Hall, S. A. 2005. Ice age vegetation and flora of New Mexico. In Lucas, S. G., Morgan, G. S., and Zeigler, K. E., eds. New Mexico's ice ages. New Mexico Museum of Natural History and Science Bulletin 28:171–183.Google Scholar
Hammer, Ø., and Harper, D. A. T.. 2006. Paleontological data analysis. Blackwell, Malden, Mass.Google Scholar
Hammer, Ø., Harper, D. A. T., and Ryan, P. D.. 2001. PAST: paleontological statistics software package for education and data analysis. Palaeontologia Electronica 4:4A.Google Scholar
Harris, A. H. 1987. Reconstruction of mid-Wisconsin environments in southern New Mexico. National Geographic Research 3:142151.Google Scholar
Harris, A. H. 1989. The New Mexican late Wisconsin—east versus west. National Geographic Research 5:205217.Google Scholar
Harris, A. H. 2015. Pleistocene vertebrates of southwestern U.S.A. and northwestern Mexico. https://www.utep.edu/leb/pleistnm/default.html, accessed 14 June 2015.Google Scholar
Haynes, C. V. 1995. Geochronology of paleoenvironmental change, Clovis type site, Blackwater draw, New Mexico. Geoarchaeology 10:317388.Google Scholar
Heintzman, P. D., Zazula, G. D., MacPhee, R. D. E., Scott, E., Cahill, J. A., McHorse, B. K., Kapp, J. D., et al. 2017. A new genus of horse from Pleistocene North America. eLife 6:e29944.Google Scholar
Hillson, S. 1996. Dental anthropology. Cambridge University Press, Cambridge.Google Scholar
Hillson, S. 2005. Teeth, 2nd ed. Cambridge University Press, Cambridge.Google Scholar
Hillson, S. 2014. Tooth development in human evolution and bioarchaeology. Cambridge University Press, Cambridge.Google Scholar
Hillson, S., and Bond, S.. 1997. Relationship of enamel hypoplasia to the pattern of tooth crown growth. American Journal of Physical Anthropology 104:89103.Google Scholar
Hofreiter, M., and Stewart, J.. 2009. Ecological change, range fluctuations and population dynamics during the Pleistocene. Current Biology 19:R584R594.Google Scholar
Holliday, T. V., and Meltzer, D. J.. 1996. Geoarchaeology of the midland (Paleoindian) site, Texas. American Antiquity 61:755771.Google Scholar
Holliday, V., Surovell, T., Meltzer, D., Grayson, D., and Boslough, M.. 2014. The Younger Dryas impact hypothesis: a cosmic catastrophe. Journal of Quaternary Science 29:515530.Google Scholar
Hoppe, K. A., and Koch, P. L.. 2006. The biochemistry of the Aucilla River fauna. Pp. 379401 in Webb, S. D., ed. First Floridians and last mastodons: the Page-Ladson site in the Aucilla River. Springer, Dordrecht, Netherlands.Google Scholar
Hoppe, K. A., Stover, S. M., Pascoe, J. R., and Amundson, R.. 2004. Tooth enamel biomineralization in extant horses: implications for isotopic microsampling. Palaeogeography, Palaeoclimatology, Palaeoecology 206:355365.Google Scholar
Hutchinson, G. E. 1957. Concluding remarks. Cold Springs Harbor Symposia on Quantitative Biology 22:415427Google Scholar
International Commission on Zoological Nomenclature. 2003. Opinion 2027 (Case 3010): usage of 17 specific names based on wild species which are predated by or contemporary with those based on domestic animals (Lepidoptera, Osteichthyes, Mammalia): conserved. Bulletin of Zoological Nomenclature 60:8184.Google Scholar
Janis, C. 1976. The evolutionary strategy of the Equidae and the origins of rumen and cecal digestion. Evolution 30:757774.Google Scholar
Janis, C. 1988. An estimation of tooth volume and hypsodonty indices in ungulate mammals and the correlation of these factors with dietary preference. Pp. 367–387 in Russell, D. E., Santoro, J.-P., and Sigogneau-Russell, D., eds. Teeth revisited: Proceedings of the VIIth International Symposium on Dental Morphology, Paris, 1986. Memoirs de Musee National and Histoire Naturelle, series C, Paris, France.Google Scholar
Jarman, P. J., and Sinclair, A. R. E.. 1979. Feeding strategy and the pattern of resource-partitioning in ungulates. Pp. 130163 in Sinclair, A. R. E. and Norton-Griffiths, M., eds. Serengeti, dynamics of an ecosystem. University of Chicago Press, Chicago.Google Scholar
Jass, C. N., Burns, J. A., and Milot, P. J.. 2011. Description of fossil muskoxen and relative abundance of Pleistocene megafauna in central Alberta. Canadian Journal of Earth Sciences 48:793800.Google Scholar
Kaiser, T. M. 2011. Feeding ecology and niche partitioning of the Laetoli ungulate faunas. Pp. 329354 in Harrison, T., ed. Paleontology and geology of Laetoli: human evolution in context, Vol. I. Geology, geochronology, paleoecology and paleoenvironment. Springer, Berlin.Google Scholar
Kaiser, T. M., and Solounias, N.. 2003. Extending the tooth mesowear method to extinct and extant equids. Geodiversitas 25:321345.Google Scholar
Kaiser, T. M., Müller, D. W. H., Fortelius, M., Schulz, E., Codron, D., and Claus, M.. 2013. Hypsodonty and tooth facet development in relation to diet and habitat in herbivorous ungulates: implications for understanding tooth wear. Mammal Review 43:3446.Google Scholar
Kierdorf, H., and Kierdorf, U.. 1997. Disturbances of the secretory stage of amelogenesis in fluorosed deer teeth: a scanning electron-microscopic study. Cell Tissue Research 289:125135.Google Scholar
Kierdorf, H., Kierdorf, U., Richards, A., and Sedlacek, F.. 2000. Disturbed enamel formation in wild boars (Sus scrofa L.) from fluoride polluted areas in Central Europe. Anatomical Record 259:1224.Google Scholar
Kierdorf, H., Kierdorf, U., Richards, A., and Josephsen, K.. 2004. Fluoride-induced alterations of enamel structure: an experimental study in the miniature pig. Anatomy and Embryology 207:463474.Google Scholar
Kierdorf, H., Zeiler, J., and Kierdorf, U.. 2006. Problems and pitfalls in the diagnosis of linear enamel hypoplasia in cheek teeth of cattle. Journal of Archaeological Science 33:16901695.Google Scholar
Kierdorf, H., Witzel, C., Upex, B., Dobney, K., and Kierdorf, U.. 2012. Enamel hypoplasia in molars of sheep and goats, and its relationship to the pattern of tooth crown growth. Journal of Anatomy 220:484495.Google Scholar
Kierdorf, U., Kierdorf, H., and Fejerskov, O.. 1993. Fluoride induced developmental changes in enamel and dentine of European roe deer (Capreolus capreolus L.) as a result of environmental pollution. Archives of Oral Biology 38:10711081.Google Scholar
Kiltie, R. A. 1984. Seasonality, gestation time, and large mammal extinctions. Pp. 299314 in Martin, P. S. and Klein, R. G., eds. Quaternary extinctions: a prehistoric revolution. University of Arizona Press, Tucson.Google Scholar
King, J. E., and Saunders, J. J.. 1984. Environmental insularity and the extinction of the American mastodont. Pp. 315339 in Martin, P. S., and Klein, R. G., eds. Quaternary extinctions: a prehistoric revolution. University of Arizona Press, Tucson.Google Scholar
King, T., Humphrey, L. T., and Hillson, S.. 2005. Linear enamel hypoplasia as indicators of systemic physiological stress: evidence from two known age-at-death and sex populations from postmedieval London. American Journal of Physical Anthropology 128:546559.Google Scholar
Koch, P. L., and Barnosky, A. D.. 2006. Late Quaternary extinctions: state of the debate. Annual Review of Ecology, Evolution, and Systematics 37:215250.Google Scholar
Koch, P. L., Hoppe, K. A., and Webb, S. D.. 1998. The isotopic ecology of late Pleistocene mammals in America. Part 1. Florida. Chemical Geology 152:119138.Google Scholar
Kooyman, B., Newman, M. E., Cluney, C., Lobb, M., Tolman, S., McNeil, P., and Hills, L. V.. 2001. Identification of horse exploitation by Clovis hunters based on protein analysis. American Antiquity 66:686691.Google Scholar
Kooyman, B., Hills, L. V., McNeil, P., and Tolman, S.. 2006. Late Pleistocene horse hunting at the Wally's Beach site (DhPg-8), Canada. American Antiquity 71:101121.Google Scholar
Lamb, H. F., and Edwards, M. E.. 1988. In Huntley, B. and Webb, T. III, eds. Vegetation history. Handbook of Vegetation Science 7:519–555. Kluwer Academic, Boston.Google Scholar
Larsen, C. S. 1997. Bioarchaeology. Cambridge University Press, Cambridge.Google Scholar
Linnaeus, C. 1758. Systema naturae per regna tria naturae, secundum classis, ordines, genera, species cum characteribus, differentiis, synonymis, locis. Tenth ed. Vol. 1. Laurentii Salvii, Stockholm.Google Scholar
Lucas, P.W., Omar, R., Al-Fadhalah, K., Almusallam, A. S., Henry, A. G., Michael, S., Arockia Thai, L., et al. 2013. Mechanisms and causes of wear in tooth enamel: implications for hominin diets. Journal of the Royal Society Interface 10:20120923.Google Scholar
Lukacs, J. R. 2001. Enamel hypoplasia in the deciduous teeth of early Miocene catarrhines: evidence of perinatal physiological stress. Journal of Human Evolution 40:319329.Google Scholar
Lukacs, J. R. 2009. Markers of physiological stress in juvenile bonobos (Pan paniscus): are enamel hypoplasia, skeletal development and tooth size interrelated? American Journal of Physical Anthropology 39:339352.Google Scholar
Lundelius, E. L. Jr. 1972. Vertebrate remains from the Gray Sand. Pp. 148163 in Hester, J. J., ed. Blackwater Locality No. 1: a stratified early man site in eastern New Mexico. Fort Burgwin Research Center, Southern Methodist University, Rancho de Taos, N.Mex.Google Scholar
Lundelius, E. L. Jr. 1984. A late Pleistocene mammalian fauna from Cueva Quebrada, Val Verde County, Texas. Carnegie Museum of Natural History Special Publication 8.Google Scholar
Lyons, S. K., Smitha, F. A., and Brown, J. H.. 2004. Of mice, mastodons, and men: human mediated extinctions on four continents. Evolutionary Ecology Research 6:339358.Google Scholar
MacPhee, R. D. E., and Marx, P. A.. 1997. The 40,000 year plague: humans, hyperdiseases, and first-contact extinctions. Pp. 169217 in Goodman, S. M. and Patterson, B. R., eds. Natural change and human impact in Madagascar. Smithsonian Institution Press, Washington, D.C.Google Scholar
Martin, P. S. 1967. Prehistoric overkill. Pp. 75120 in Martin, P. S. and Wright, H. E. J., eds. Pleistocene extinctions: the search for a cause. Yale University Press, New Haven, Conn.Google Scholar
Martin, P. S. 1984. Prehistoric overkill: The global model. Pp. 384403 in Martin, P. S., and Klein, R. G., eds. Quaternary Extinctions: a Prehistoric Revolution. University of Arizona Press, Tucson.Google Scholar
MathWorks. 2018. MATLAB: the language of technical computing, Version R2018a. Natick, Mass.Google Scholar
McDonald, J. N. 1981. North American Bison, their classification and evolution. University of California Press, Berkeley.Google Scholar
McNaughton, S. J., and Georgiadis, N. J.. 1986. Ecology of African grazing and browsing mammals. Annual Review of Ecology and Systematics 17:3965.Google Scholar
Mead, A. J. 1999. Enamel hypoplasia in Miocene rhinoceroses (Teleoceras) from Nebraska: evidence of severe physiological stress. Journal of Vertebrate Paleontology 19:391397.Google Scholar
Meltzer, D. J. 2015. Pleistocene overkill and North American Mammalian extinctions. Annual Review of Anthropology 44:3353.Google Scholar
Meltzer, D. J., Holliday, V. T., Cannon, M. D., and Miller, D. S.. 2014. Chronological evidence fails to support claims for an isochronous widespread layer of cosmic impact indicators dated to 12,800 years ago. Proceedings of the National Academy of Sciences USA 111:E2162E2171.Google Scholar
Merceron, G., Blondel, C., Brunet, M., Sen, S., Solounias, N., Viriot, L., and Heintz, E.. 2004. The Late Miocene paleoenvironment of Afghanistan as inferred from dental microwear in artiodactyls. Palaeogeography, Palaeoclimatology, Palaeoecology 207:143163.Google Scholar
Merceron, G., Blondel, C., de Bonis, L., Koufos, G. D., and Viriot, L.. 2005. A new method of dental microwear analysis: application to extant primates and Ouranopithecus macedoniensis (Late Miocene of Greece). Palaios 20:551561.Google Scholar
Merceron, G., Schulz, E., Kordos, L., and Kaiser, T. M.. 2007. Paleoenvironment of Dryopithecus brancoi at Rudabánya, Hungary: evidence from dental meso- and micro-wear analyses of large vegetarian mammals. Journal of Human Evolution 53:331349.Google Scholar
Merceron, G., Escarguel, G., Angibault, J. M., and Verheyden-Tixier, H.. 2010. Can dental microwear textures record inter-individual dietary variations? PLoS ONE 5:e9542.Google Scholar
Mihlbachler, M. C., Beatty, B. L., Caldera-Siu, A., Chan, D., and Lee, R.. 2012. Error rates and observer bias in dental microwear analysis using light microscopy. Palaeontologia Electronica 15:12A.Google Scholar
Mihlbachler, M. C., Campbell, D., Ayoub, M., Chen, C., and Ghani, I.. 2016. Comparative dental microwear of ruminant and perissodactyl molars: implications for paleodietary analysis of rare and extinct ungulate clades. Paleobiology 42:98116.Google Scholar
Miles, A. E. W, and Grigson, C.. 1990. Colyer's variations and diseases of the teeth of animals, rev. ed. Cambridge University Press, Cambridge.Google Scholar
Moggi-Cecchi, J., and Crovella, S.. 1991. Occurrence of enamel hypoplasia in the dentitions of simian primates. Folia Primatologica 57:106110.Google Scholar
Morlan, R. E. 1989. Paleoecological implications of Late Pleistocene and Holocene microtine rodents from the Bluefish Caves, northern Yukon Territory. Canadian Journal of Earth Sciences 26:149156.Google Scholar
Mosimann, J. E., and Martin, P. S.. 1975. Simulating overkill by paleoindians. American Scientist 63:304313.Google Scholar
Murray, M. G., and Brown, D.. 1993. Niche separation of grazing ungulates in the Serengeti: an experimental test. Journal of Animal Ecology 62:380389.Google Scholar
Nelson, S. V., Badgley, C., and Zakem, E.. 2005. Microwear in modern squirrels in relation to diet. Palaeontologia Electronica 8:14A.Google Scholar
Niven, L. B. 2002. Enamel hypoplasia in bison: paleoecological implications for modeling hunter-gatherer procurement and processing on the Northwestern Plains. Archaeozoologica 11:101112.Google Scholar
Niven, L. B., Egeland, C. P., and Todd, L. C.. 2004. An inter-site comparison of enamel hypoplasia in bison: implications for paleoecology and modelling Late Plains archaic subsistence. Journal of Archaeological Science 31:17831794.Google Scholar
Pérez-Crespo, V. A., Arroyo-Cabrales, J., Alva-Valdivia, L. M., Morales-Puente, P., and Cienfuegos-Alvarado, E.. 2012. Datos isotópicos (δ13C, δ18O) de la fauna pleistocenica de la Laguna de las Cruces, San Luis Potosí, México. Revista Mexicana de Ciencias Geológicas 29:299307.Google Scholar
Polyak, V. J., Asmerom, Y., Burns, S. J., and Lachniet, M. S.. 2012. Climatic backdrop to the terminal Pleistocene extinction of North American mammals. Geology 40:10231026.Google Scholar
Prado, J. L., Alberdi, M. T., Azanza, B., Sanchez, B., and Frassinetti, D.. 2005. The Pleistocene Gomphotheriidae (Proboscidea) from South America. Quaternary International 126/128:2130.Google Scholar
Priego-Vargas, J., Bravo-Cuevas, V. M., and Jiménez-Hidalgo, E.. 2017. Revisión taxonómica de los équidos del Pleistoceno de México con base en la morfología dental. Revista Brasileira de Paleontologia 20:239268.Google Scholar
Ripple, W. J., and Van Valkenburgh, B.. 2010. Linking top-down forces to the Pleistocene megafaunal extinctions. BioScience 60:516526.Google Scholar
Ritchie, J. C., Cinq-Mars, J., and Cwynar, L. C.. 1982. L'environnement tardiglaciaire du Yukon septentrional, Canada. Géographie physique et Quaternaire 36:241250.Google Scholar
Rivals, F., and Athanassiou, A.. 2008. Dietary adaptations in an ungulate community from the late Pliocene of Greece. Palaeogeography, Palaeoclimatology, Palaeoecology 265:134139.Google Scholar
Rivals, F., Solounias, N., and Mihlbachler, M. C.. 2007. Evidence for geographic variation in the diets of late Pleistocene and early Holocene Bison in North America, and differences from the diets of recent Bison. Quaternary Research 68:338346.Google Scholar
Rivals, F., Schulz, E., and Kaiser, T. M.. 2008. Climate-related dietary diversity of the ungulate faunas from the middle Pleistocene succession (OIS 14-12) at the Caune de l'Arago (France). Paleobiology 34:117127.Google Scholar
Rivals, F., Mihlbachler, M. C., Solounias, N., Mol, D., Semprebon, G. M., de Vos, J., and Kalthoff, D. C.. 2010. Palaeoecology of the Mammoth Steppe fauna from the late Pleistocene of the North Sea and Alaska: separating species preferences from geographic influence in paleoecological dental wear analysis. Palaeogeography, Palaeoclimatology, Palaeoecology 286:4254.Google Scholar
Roohi, G., Raza, S. M., Khan, A. M., Ahmad, R. M., and Akhtar, M.. 2015. Enamel hypoplasia in Siwalik rhinocerotids and its correlation with Neogene climate. Pakistan Journal of Zoology 47:14331443.Google Scholar
Sánchez, B., Prado, J. L., and Alberdi, M. T.. 2006. Ancient feeding, ecology and extinction of Pleistocene horses from the Pampean region, Argentina. Ameghiniana 43:427436.Google Scholar
Sanson, G. D., Kerr, S. A., and Gross, K. A.. 2007. Do silica phytoliths really wear mammalian teeth? Journal of Archaeological Science 34:526531.Google Scholar
Schulz, E., Piotrowski, V., Clauss, M., Mau, M., Merceron, G., and Kaiser, T. M.. 2013. Dietary abrasiveness is associated with variability of microwear and dental surface texture in rabbits. PLoS ONE 8:e56167.Google Scholar
Schwartz, G. T., Reid, D. J., Dean, M. C., and Zihlman, A. L.. 2006. A faithful record of stressful life events preserved in the dental development of a juvenile gorilla. International Journal of Primatology 27:12011219.Google Scholar
Scott, E. 1996. The small horse from Valley Wells, San Bernardino County, California. In Reynolds, R. E. and Reynolds, J., eds. Punctuated chaos in the northeastern Mojave desert. SBCM Association Quarterly 43(1,2):85–89. San Bernardino County Museum Association, Redland, Calif.Google Scholar
Scott, E. 2010. Extinctions, scenarios, and assumptions: changes in latest Pleistocene large herbivore abundance and distribution in western North America. Quaternary International 217:225239.Google Scholar
Scott, J. R. 2012. Dental microwear texture analysis of extant African Bovidae. Mammalia 76:157174.Google Scholar
Scott, R. S., Ungar, P. S., Bergstrom, T. S., Brown, C. A., Grine, F. E., Teaford, M. F., and Walker, A.. 2005. Dental microwear texture analysis shows within-species diet variability in fossil hominins. Nature 436:693695.Google Scholar
Scott, R. S., Ungar, P. S., Bergstrom, S., Brown, C. A., Childs, B. E., Teaford, M. F., and Walker, A.. 2006. Dental microwear texture analysis: technical considerations. Journal of Human Evolution 51:339349.Google Scholar
Semprebon, G. M., Godfrey, L. R., Solounias, N., Sutherland, M. R., and Jungers, W. L.. 2004. Can low-magnification stereomicroscopy reveal diet? Journal of Human Evolution 47:115144.Google Scholar
Shearer, T. R., Kolstad, D. L., and Suttie, J. W.. 1978. Bovine dental fluorosis: histologic and physical characteristics. American Journal of Veterinary Research. 39:597602.Google Scholar
Shupe, J. L., and Olson, A. E.. 1983. Clinical and pathological aspects of fluoride toxicosis in animals. Pp. 319338 in Shupe, J. L., Peterson, H. B., and Leone, N. C., eds. Fluorides: effects on vegetation, animals and humans. Paragon Press, Salt Lake City, Utah.Google Scholar
Skinner, M., and Goodman, A. H.. 1992. Anthropological uses of developmental defects of enamel. Pp. 153174 in Saunders, S. R. and Katzenberg, M. A., eds. Skeletal biology of past peoples. Wiley-Liss, New York.Google Scholar
Skinner, M. F., and Hopwood, D.. 2004. A hypothesis for the causes and periodicity of repetitive linear enamel hypoplasia (rLEH) in large, wild African (Pan troglodytes and Gorilla gorilla) and Asian (Pongo pygmaeus) apes. American Journal of Physical Anthropology 123:216235.Google Scholar
Skinner, M. F., and Hung, J. T. W.. 1986. Localized enamel hypoplasia of the primary canine. Journal of Dentistry for Children 53:197200.Google Scholar
Smith, F. A., Tomé, C. P., Elliot Smith, E. A., Lyons, S. K., Newsome, S. D., and Stafford, T. W.. 2016. Unraveling the consequences of the terminal Pleistocene megafauna extinction on mammal community assembly. Ecography 39:223239.Google Scholar
Smith, T. M., and Boesch, C.. 2015. Developmental defects in the teeth of three wild chimpanzees from the Taï forest. American Journal of Physical Anthropology 157:556570.Google Scholar
Solounias, N., and Semprebon, G.. 2002, Advances in the reconstruction of ungulate ecomorphology with application to early fossil equids. American Museum Novitates 3366:149.Google Scholar
Solounias, N., Teaford, M. F., and Walker, A.. 1988. Interpreting the diet of extinct ruminants: the case of a non-browsing giraffid. Paleobiology 14:287300.Google Scholar
Spencer, L. M. 1995. Morphological correlates of dietary resource partitioning in the African Bovidae. Journal of Mammalogy 76:448471.Google Scholar
Stewart, J. R. 2009. The evolutionary consequence of the individualistic response to climate change. Journal of Evolutionary Biology 22:23632375.Google Scholar
Stewart, K. M., Bowyer, R. T., Kie, J. G., Cimon, N. J., and Johnson, B. K.. 2002. Temporospatial distributions of elk, mule deer and cattle: resource partition and competitive displacement. Journal of Mammalogy 83:229244.Google Scholar
Strong, W. L., and Hills, L. V.. 2005. Late-glacial and Holocene palaeovegetation zonal reconstruction for central and north-central North America. Journal of Biogeography 32:10431062.Google Scholar
Stuart, J. A. 2015. Late Quaternary megafaunal extinctions on the continents: a short review. Geological Journal 50:338363.Google Scholar
Suckling, G., Elliott, D. C., and Thurley, D. C.. 1986. The macroscopic appearance and associated histological changes in the enamel organ of hypoplastic lesions of sheep incisor teeth resulting from induced parasitism. Archives of Oral Biology 31:427439.Google Scholar
Suckling, G., Thurley, D. C., and Nelson, D. G. A.. 1988. The macroscopic and scanning electron-microscopic appearance and microhardness of the enamel, and the related histological changes in the enamel organ of erupting sheep incisors resulting from a prolonged low daily dose of fluoride. Archives of Oral Biology 33:361373.Google Scholar
Suckling, G. W. 1989. Developmental defects of enamel—historical and present-day perspectives of their pathogenesis. Advances in Dental Research 3:8794.Google Scholar
Surovell, T., Holliday, V. T., Gingerich, J. A. M., Kreton, C., Haynes, C. V., Hilman, I., Wagner, D. P., Johnson, E., and Claeys, P.. 2009. An independent evaluation of the Younger Dryas extraterrestrial impact hypothesis. Proceedings of the National Academy of Sciences USA 106:1815518158.Google Scholar
Tebedge, S. 1988. Paleontology and paleoecology of the Pleistocene mammalian fauna of Dark Canyon Cave, Eddy County, New Mexico. PhD dissertation. University of Texas at Austin.Google Scholar
Timperley, C., and Lundelius, E. L. Jr. 2008. Dental enamel hypoplasia in late Pleistocene Equus from Texas and New Mexico. Pp. 45–50 in Farley and Chaote, eds. Unlocking the unknown: papers honoring Dr. Richard J. Zakrzewski. Fort Hays State University Special Issue No. 2, Hays, Kans.Google Scholar
Tütken, T., Kaiser, T. M., Vennemann, T., and Merceron, G.. 2013. Opportunistic feeding strategy for the earliest Old World hypsodont equids: evidence from stable isotope and dental wear proxies. PLoS ONE 8:e74463.Google Scholar
Ungar, P. S., Teaford, M. F., Glander, K. E., and Pastor, R. F.. 1995. Dust accumulation in the canopy: a potential cause of dental microwear in primates. American Journal of Physical Anthropology 97:9399.Google Scholar
Ungar, P. S., Brown, C. A., Bergstrom, T. S., and Walker, A.. 2003. Quantification of dental microwear by tandem scanning confocal microscopy and scale-sensitive fractal analyses. Scanning 25:185193.Google Scholar
Ungar, P. S., Merceron, G., and Scott, R. S.. 2007. Dental microwear texture analysis of Varswater bovids and early Pliocene paleoenvironments of Langebaanweg, Western Cape Province, South Africa. Journal of Mammalian Evolution 14:163181.Google Scholar
Ungar, P. S., Scott, R. S., Grine, F. E., and Teaford, M. F.. 2010. Molar microwear textures and the diets of Australopithecus anamensis and Australopithecus afarensis. Philosophical Transactions of the Royal Society of London B 365:33453354.Google Scholar
Upex, B., Balasse, M., Tresset, A., Arbuckle, B., and Dobney, K.. 2014. Protocol for recording enamel hypoplasia in modern and archaeological caprine populations. International Journal of Osteoarchaeology 24:7989.Google Scholar
Walker, A., Hoeck, H. H., and Perez, L.. 1978. Microwear of mammalian teeth as an indicator of diet. Science 201:908910.Google Scholar
Waters, M. R., Stafford, T. W. Jr., Kooyman, B., and Hills, L. V.. 2015. Late Pleistocene horse and camel hunting at the southern margin of the ice-free corridor: reassessing the age of Wally's Beach, Canada. Proceedings of the National Academy of Sciences USA 112:42634267.Google Scholar
Weinstock, J., Willerslev, E., Sher, A., Tong, W., Ho, S. Y. W., Rubenstein, D., Storer, J., et al. 2005. Evolution, systematics, and phylogeography of Pleistocene horses in the New World: a molecular perspective. PLoS Biology 3:e241.Google Scholar
Willerslev, E., Davison, J., Moora, M., Zobel, M., Coissac, E., Edwards, M. E., Lorenzen, E. D., et al. 2014. Fifty thousand years of Arctic vegetation and megafaunal diet. Nature. 506:4751.Google Scholar
Wilson, D. E., and Reeder, D. A. M., eds. 2005. Mammal species of the world. A taxonomic and geographic reference, 3rd ed. Smithsonian Institution Press, Washington, D.C.Google Scholar
Winans, M. C. 1985. Revision of North American fossil species of the genus Equus (Mammalia: Perissodactyla: Equidae). PhD dissertation. University of Texas at Austin.Google Scholar
Winans, M. C. 1989. A quantitative study of North American fossil species of the genus Equus. Pp. 262297 in Prothero, D. R. and Schoch, R. M., eds. The evolution of Perissodactyls. Clarendon Press, Oxford.Google Scholar
Witzel, C., Kierdorf, U., Dobney, K., Ervynck, A., Vanpoucke, S., and Kierdorf, H.. 2006. Reconstructing impairment of secretory ameloblasts function in porcine teeth by analysis of morphological alterations in dental enamel. Journal of Anatomy 209:93110.Google Scholar
Witzel, C., Kierdorf, U., Schultz, M., and Kierdorf, H.. 2008. Insights from the inside: histological analysis of abnormal enamel microstructure associated with hypoplastic enamel defects in human teeth. American Journal of Physical Anthropology. 136:400414.Google Scholar
Wyckoff, D. G., and Dalquest, W. W.. 1997. From whence they came: the paleontology of Southern Plains bison. Memoir 29. Plains Anthropologist 42:532.Google Scholar
Young, R. R., Burns, J. A., Smith, D. G., Arnold, L. D., and Rains, R. B.. 1994. A single, late Wisconsin, Laurentide glaciation, Edmonton area and southwestern Alberta. Geology 22:683686.Google Scholar
Young, R. R., Burns, J. A., Rains, R. B., and Schowalter, D. B.. 1999. Late Pleistocene glacial geomorphology and environment of the Hand Hills region and southern Alberta, related to Middle Wisconsin fossil prairie dog sites. Canadian Journal of Earth Sciences 36:15671581.Google Scholar
Zazula, G. D., Schweger, C. E., Beaudoin, A. B., and McCourt, G. H.. 2006. Macrofossil and pollen evidence for full-glacial steppe within an ecological mosaic along the Bluefish River, eastern Beringia. Quaternary International 142–143:219.Google Scholar
Zhou, L., and Corruccini, R. S.. 1998. Enamel hypoplasias related to famine stress in living Chinese. American Journal of Human Biology 10:723733.Google Scholar