Hostname: page-component-7bb8b95d7b-lvwk9 Total loading time: 0 Render date: 2024-09-26T10:03:01.307Z Has data issue: false hasContentIssue false

Mid-Pleistocene Change in Large Mammal Faunas of East Africa

Published online by Cambridge University Press:  20 January 2017

Richard Potts
Affiliation:
Department of Anthropology, National Museum of Natural History, Smithsonian Institution, Washington, DC 20560 National Museums of Kenya, P.O. Box 40658, Nairobi, Kenya
Alan Deino
Affiliation:
Berkeley Geochronology Center, 2455 Ridge Road, Berkeley, California 94709

Abstract

Single-crystal 40Ar/39Ar age estimates of 392,000 ± 4000 to 330,000 ± 6000 yr from Lainyamok, a middle Pleistocene fossil locality in the southern Kenya rift, document the oldest evidence from sub-Saharan Africa of a diverse, large mammal fauna consisting entirely of extant species. The inferred age of this fauna implies an upper limit for extinction of species that characterize well-calibrated, mid-Pleistocene fossil assemblages in East Africa. For its age and species richness, the Lainyamok fauna is surprising for its lack of extinct forms (e.g., the bovine Pelorovis) well documented in later faunal assemblages of East and South Africa. Definitive presence of the South African blesbok (Damaliscus dorcas) is also unexpected, especially as this alcelaphine bovid is the dominant large mammal in the Lainyamok fauna. These age estimates and the faunal composition at Lainyamok indicate that geographic ranges and taxonomic associations of extant largebodied mammals were susceptible to wide fluctuations in sub-Saharan Africa over the past 330,000 yr. This inference is consistent with the hypothesis of nonanalogue, or ephemeral, biotas believed to characterize late Quaternary ecosystems of northern continents.

Type
Research Article
Copyright
University of Washington

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.)

References

Alroy, J. (1992). Conjunction among taxonomic distributions and the Miocene mammalian biochronology of the Great Plains. Paleobiology 18, 326343.Google Scholar
Azzaroli, A. (1983). Quaternary mammals and the “End-Villafranchian” dispersal event—A turning point in the history of Eurasia. Palaeogeography Palaeoclimatology Palaeoecology 44, 117139.Google Scholar
Baker, B. H. (1958). “Geology of the Magadi Area.” Geological Survey of Kenya, Nairobi.Google Scholar
Barry, J. C. Flynn, L. J., and Pilbeam, D. R. (1990). Faunal diversity and turnover in a Miocene terrestrial sequence. In “Causes of Evolution: A Paleontological Perspective” (Ross, R. M. and Allmon, W. D., Eds.), pp. 381421. University of Chicago Press, Chicago.Google Scholar
Beden, M. (1983). Family Elephantidae. In “Koobi Fora Research Project, Vol. 2” (Harris, J. M., Ed.), pp. 40129. Clarendon Press, Oxford.Google Scholar
Churcher, C. S. (1993). Equus grevyi. Mammalian Species No. 453, 19.Google Scholar
Cole, K. (1985). Past rates of change, species richness, and a model of vegetational inertia in the Grand Canyon, Arizona. American Natu-ralist 125, 289303.CrossRefGoogle Scholar
Davis, M. B. (1976). Pleistocene biogeography of temperate deciduous forests. Geoscience and Man 13, 1326.Google Scholar
Deino, A., and Potts, R. (1990). Single-crystal 40Ar/39Ar dating of the Olorgesailie Formation, southern Kenya rift. Journal of Geophysical Research 95, 84538470.Google Scholar
Deino, A., and Potts, R. (1992). Age-probability spectra for examination of single-crystal 40Ar/39Ar dating results: Examples from Olorgesailie, southern Kenya rift valley. Quaternary International 7/8, 8189.Google Scholar
Delcourt, H. R. Delcourt, P. A., and Webb, T. (1983). Dynamic plant ecology: The spectrum of vegetational change in space and time. Quaternary Science Reviews 1, 153175.Google Scholar
Eisenmann, V. (1979). Le genre Hipparion (Mammalia, Perissodactyla) et son intérêt biostratigraphique en Afrique. Bulletin de la Société Géologique de France 21, 277281.Google Scholar
Eisenmann, V. (1983). Family Equidae. In “Koobi Fora Research Project, Vol. 2,” (Harris, J. M., Ed.), pp. 156214. Clarendon Press, Oxford.Google Scholar
Eugster, H. P. (1981). Lake Magadi, Kenya, and its precursors. In “Hypersaline Brines and Evaporitic Environments” (Nissenbaum, A., Ed.), pp. 195232. Elsevier, Amsterdam.Google Scholar
Fairhead, J. D. Mitchell, J. G„ and Williams, L. (1972). New K/Ar determinations on rift volcanics of S. Kenya and their bearings on age of rift faulting. Nature 238, 6669.Google Scholar
Graham, R. W. (1976). Late Wisconsin mammal faunas and environmental gradients of the eastern United States. Paleobiology 2, 343350.Google Scholar
Graham, R. W., and Lundelius, E. L. Jr. (1984). Coevolutionary disequilibrium and Pleistocene extinctions. In “Quaternary Extinc-tions” (Martin, P. S. and Klein, R. G., Eds.), pp. 223249. Univ. of Arizona Press, Tucson.Google Scholar
Graham, R. W., and Mead, J. I. (1987). Environmental fluctuations and evolution of mammalian faunas during the last deglaciation in North America. In “North America and Adjacent Oceans during the Last Deglaciation” (Ruddiman, W. F. and Wright, H. E. Jr., Eds.), pp. 371402. Geological Soc. of America, Boulder.Google Scholar
Harris, J. M. (Ed.) (1983). “Koobi Fora Research Project, Vol. 2.” Clarendon Press, Oxford.Google Scholar
Harris, J. M. (Ed.) (1991). “Koobi Fora Research Project, Vol. 3” Clarendon Press, Oxford.Google Scholar
Harris, J. M. Brown, F. H., and Leakey, M. G. (1988). Stratigraphy and Paleontology of Pliocene and Pleistocene Localities West of Lake Turkana, Kenya “Contributions in Science,” Vol. 339, pp. 1128, Natural History Museum of Los Angeles County, Los Angeles.Google Scholar
Hay, R. L. (1966). Zeolites and zeolitic reactions in sedimentary rocks. Geological Society of America Special Paper 85, 1130.Google Scholar
Hay, R. L. (1976). “Geology of the Olduvai Gorge.” Univ. of California Press, Berkeley.Google Scholar
Hay, R. L. (1978). Geologic occurrence of zeolites. In “Natural Zeolites: Occurrences, Properties, Use” (Sand, L. B. and Mumpton, F. A., Eds.), pp. 135143. Pergamon, New York.Google Scholar
Hay, R. L. (1990). Olduvai Gorge: A case history in the interpretation of hominid paleoenvironments in East Africa. In “Establishment of a Geologic Framework for Paleoanthropology” (Laporte, L. F., Ed.), pp. 2337. Special Paper 242, Geological Society of America, Boulder, CO.Google Scholar
Hendey, Q. B. (1974). The late Cenozoic Carnivora of the southwestern Cape Province. Annals of the South African Museum 63,169.Google Scholar
Huntley, B., and Webb, T. (1989). Migration: Species response to climatic variations caused by changes in the Earth’s orbit. Journal of Biogeography 15, 519.Google Scholar
Kalb, J. E. Jolly, C. J. Mebrate, A. Tebedge, S. Smart, C. Oswald, E. B. Cramer, D. Whitehead, P. Wood, C. B. Conroy, G. C. Adefris, T. Sperling, L., and Kana, B. (1982). Fossil mammals and artefacts from the Middle Awash Valley. Nature 298, 2529.CrossRefGoogle ScholarPubMed
Klein, R. G. (1975), Middle Stone Age man-animal relationships in southern Africa: Evidence from Die Kelders and Klasies River Mouth. Science 190, 265267.Google Scholar
Klein, R. G. (1980). Environmental and ecological implications of large mammals from Upper Pleistocene and Holocene sites in southern Africa.” Annals of the South African Museum 81, 223283.Google Scholar
Klein, R. G. (1984). Mammalian extinctions and Stone Age people in Africa. In “Quaternary Extinctions: A Prehistoric Revolution” (Martin, P. S., and Klein, R. G., Eds.), pp. 553573. Univ. of Arizona Press, Tucson.Google Scholar
Klein, R. G., and Cruz-Uribe, K. (1991). The bovids from Elandsfontein South Africa and their implications for the age, palaeoenvironment, and origins of the site. African Archaeological Review 9, 2179.Google Scholar
Koch, C. P. (1986). “The Vertebrate Taphonomy and Paleoecology of the Olorgesailie Formation (Middle Pleistocene, Kenya).” Ph.D. dis-sertation, Univ. of Toronto.Google Scholar
Kowallis, B. J. Heaton, J. S., and Bringhurst, K. (1986). Fission-track dating of volcanically derived sedimentary rocks. Geology 14, 1922.Google Scholar
Leakey, M. D. (in press). “Olduvai Gorge, Volume 5.” Cambridge Univ. Press, New York.Google Scholar
Lundelius, E. L. Jr. (1976). Vertebrate paleontology of the Pleistocene: An overview. In “Geoscience and Man” (West, R. C., Ed.), Vol. 13, pp. 4959. Louisiana State Univ., Baton Rouge.Google Scholar
Lundelius, E. L. Jr. Graham, R. W. Anderson, E. Guilday, J. Holman, J. A. Steadman, D. W., and Webb, S. D. (1983). Terrestrial vertebrate faunas. In “Late Quaternary Environments of the United States: Vol. 1” (Porter, S. C., Ed.), pp. 311353. Univ. of Minnesota Press, Minneapolis.Google Scholar
Maglio, V. J. (1975). Pleistocene Faunal Evolution in Africa and Eurasia. In “After the Australopithecines” (Butzer, K. W. and Isaac, G. LI., Eds.), pp. 419476. Mouton, The Hague.Google Scholar
Maglio, V. J. (1978). Patterns of faunal evolution. In “Evolution of African Mammals” (Maglio, V. J. and Cooke, H. B. S., Eds.), PP. 603619. Harvard Univ. Press, Cambridge, MA.Google Scholar
Maglio, V. J., and Cooke, H. B. S. (Eds.) (1978). “Evolution of African Mammals.” Harvard Univ. Press, Cambridge, MA.Google Scholar
Marean, C. W., and Gifford-Gonzalez, D. (1991). Late Quaternary extinct ungulates of East Africa and paleoenvironmental implications. Nature 350, 418420.Google Scholar
Potts, R. Shipman, P., and Ingall, E. (1988). Taphonomy, paleoecology and hominids of Lamyamok, Kenya. Journal of Human Evolution 17, 597614.CrossRefGoogle Scholar
Shipman, P. Potts, R., and Pickford, M. (1983). Lainyamok: A new middle Pleistocene hominid site. Nature 306, 365368.Google Scholar
Steiger, R. H., and Jager, E. (1977). Subcommision on geochronology: Convention on the use of decay constants in geoand cosmochronology. Earth and Planetary Science Letters 36, 359362.Google Scholar
West, R. G. (1964). Inter-relations of ecology and Quaternary palaeobotany. Journal of Ecology 52 (Suppl.), 4757.Google Scholar