Hostname: page-component-7479d7b7d-qs9v7 Total loading time: 0 Render date: 2024-07-12T22:02:26.235Z Has data issue: false hasContentIssue false

Responses of African bovids to Pliocene climatic change

Published online by Cambridge University Press:  05 October 2018

René Bobe
Affiliation:
Evolution of Terrestrial Ecosystems Program, Department of Paleobiology, NHB MRC 121, Smithsonian Institution, Washington, D.C. 20560. E-mail: Bobe.Rene@nmnh.si.edu
Gerald G. Eck
Affiliation:
Department of Anthropology, Box 353100, University of Washington, Seattle, Washington 98195. E-mail: ggeck@u.washington.edu

Abstract

The record of fossil mammals from the Shungura Formation, lower Omo Valley of southern Ethiopia, represents one of the largest and most carefully controlled samples for deciphering the responses of land faunas to global-scale climatic change. We use the abundant and continuous fossil record of the family Bovidae to analyze the effects of a late Pliocene climatic shift toward increased aridity in Africa beginning at 2.8 Ma and intensifying at about 2.4 Ma. A database consisting of 4233 specimen-based records collected under well-defined procedures is used to define patterns through time in bovid abundances while also controlling for taphonomic and other potential biases. Univariate and multidimensional (correspondence analysis) methods are used to study changes in bovid abundances through time. Our results indicate that bovids experienced an increase in species richness and a rapid episode of change in taxonomic abundances at 2.8 ± 0.1 Ma (between Members B and C), and that this shift was followed by gradual and prolonged changes in abundance between 2.8 and 2.0 Ma (between Member C and upper Member G). An analysis of skeletal-element abundances through the Shungura sequence shows that only moderate changes in taphonomic conditions occurred between 3.0 and 2.1 Ma, when the lower Omo Valley was dominated by a large, meandering river, but that significant changes in the mode of preservation accompanied the onset of lacustrine depositional environments at 2.1 Ma (between lower and upper Member G). A juxtaposition of taxonomic with taphonomic patterns shows that the shift in taxonomic abundances at 2.8 Ma occurred in the absence of significant changes in taphonomic conditions. The main changes in bovid relative abundances and diversity appear to have been driven by broad environmental and climatic changes in Africa. As environmental indicators, bovids show a transition in the Omo at about 2.8 Ma from closed and wet environments in Member B to closed but dry environments in Member C. This drying trend intensified in Members D, E, and F, between about 2.5 and 2.3 Ma. In lower Member G, between 2.3 and 2.1 Ma, there was an increase in bovid abundance and diversity, which may be a result of greater environmental heterogeneity. The pattern of environmental change depicted by Shungura bovids is consistent with independently derived evidence of Omo paleoenvironments (from paleosols, paleoflora, and micromammals), and with regional and global evidence of climatic changes, especially acute between 2.8 and 2.3 Ma, that caused the initiation of glacial cycles in the north and drier climate in the tropics of Africa. Even though the Omo bovids showed distinct responses to large-scale climatic and environmental change, the Omo bovid community also had important attributes of long-term stability: two species, Aepyceros shungurae and Tragelaphus nakuae, dominated the community for nearly one million years. This study highlights the importance of carefully controlled collection procedures of fossil vertebrates and provides an important demonstration of the potential complexity in mode and rate of responses of land faunas to climatic change.

Type
Research Article
Copyright
Copyright © The Paleontological Society, 2001 

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

Literature Cited

Alemseged, Z. 1998. L'Hominidé Omo-323: sa position phylétique et son environnement dans le cadre de l'évolution des communautés de mammifères du Plio-Pléistocène dans la basse vallée de l'Omo (Éthiopie). Ph.D. dissertation. Muséum national d'Histoire naturelle, Paris.Google Scholar
Arambourg, C. 1947. Contribution à l'étude géologique et paléontologique du Bassin du Lac Rodolphe et de la Basse Vallée de l'Omo, deuxième partie: Paléontologie. Mission Scientifique de l'Omo 1932–1933, 1(3). Paris.Google Scholar
Beden, M. 1976. Proboscideans from the Omo group formations. Pp. 193208 in Coppens et al. 1976.Google Scholar
Beden, M. 1987. Les faunes Plio-Pléistocène de la basse vallée de l'Omo (Éthiopie), Tome 2. Les eléphantidés (Mammalia-Proboscidea). Editions du CNRS, Paris.Google Scholar
Begon, M., Harper, J. L., and Townsend, C. R. 1990. Ecology: individuals, populations, and communities. Blackwell Scientific, Boston.Google Scholar
Behrensmeyer, A. K. 1982. Time resolution in fluvial vertebrate assemblages. Paleobiology 8:211227.Google Scholar
Behrensmeyer, A. K. 1988. Vertebrate preservation in fluvial channels. Palaeogeography, Palaeoclimatology, Palaeoecology 63:183199.CrossRefGoogle Scholar
Behrensmeyer, A. K., Todd, N. E., Potts, R., and McBrinn, G. E. 1997. Late Pliocene faunal turnover in the Turkana Basin, Kenya. Science 278:15891594.Google Scholar
Benzécri, J. P. 1992. Correspondence analysis handbook. Dekker, New York.Google Scholar
Boaz, N. T. 1977. Paleoecology of Plio-Pleistocene Hominidae in the Lower Omo Basin, Ethiopia. Ph.D. dissertation. University of California, Berkeley.Google Scholar
Bobe, R. 1996. Pliocene environmental changes derived from faunal analysis of the Shungura Formation, Ethiopia. Journal of Vertebrate Paleontology 16(3):23. A.Google Scholar
Bobe, R. 1997. Hominid environments in the Pliocene: an analysis of fossil mammals from the lower Omo valley, Ethiopia. Ph.D. dissertation, University of Washington, Seattle.Google Scholar
Bobe, R. and Behrensmeyer, A. K. 1999. Environmental changes in early hominid evolution derived from analysis of fossil mammals from the lower Omo Valley, Ethiopia. Journal of Human Evolution 36(4):A. 3.Google Scholar
Bonnefille, R. 1983. Evidence for a cooler and drier climate in the Ethiopian uplands towards 2. 5 Myr ago. Nature 303:487491.Google Scholar
Bonnefille, R. 1994. Palynology and paleoenvironment of East African hominid sites. Pp. 415427 in Corruccini and Ciochon 1994.Google Scholar
Bonnefille, R. 1995. A reassessment of the Plio-Pleistocene pollen record of East Africa. Pp. 299310 in Vrba et al. 1995.Google Scholar
Bonnefille, R. and Dechamps, R. 1983. Data on fossil flora. Pp. 191207 in de Heinzelin 1983.Google Scholar
Bonnefille, R. and Letouzey, R. 1976. Fruits fossiles d'Antrocaryon dans la vallée de l'Omo (Ethiopie). Adansonia 16:6582.Google Scholar
Brown, F. H. 1982. Tulu Bor Tuff at Koobi Fora correlated with the Sidi Hakoma Tuff at Hadar. Nature 300:631633.Google Scholar
Brown, F. H. 1994. Development of Pliocene and Pleistocene chronology of the Turkana basin, East Africa, and its relation to other sites. Pp. 285312 in Corruccini and Ciochon 1994.Google Scholar
Brown, F. H. 1995. The potential of the Turkana Basin for paleoclimatic reconstruction in East Africa. Pp. 319330 in Vrba et al. 1995.Google Scholar
Brown, F. H. and Feibel, C. S. 1986. Revision of lithostratigraphic nomenclature in the Koobi Fora region, Kenya. Journal of the Geological Society, London 143:297310.CrossRefGoogle Scholar
Brown, F. H. and Heinzelin, J. de. 1983. The lower Omo basin. Pp. 724 in de Heinzelin 1983.Google Scholar
Brown, F. H., McDougall, I., Davies, I., and Maier, R. 1985. An integrated Plio-Pleistocene chronology for the Turkana basin. Pp. 8290 in Delson, E., ed. Ancestors: the hard evidence. Alan R. Liss, New York.Google Scholar
Burckle, L. H. 1995. Current issues in Pliocene paleoclimatology. Pp. 37 in Vrba et al. 1995.Google Scholar
Buzas, M. A. 1990. Another look at confidence limits for species proportions. Journal of Paleontology 64:842843.Google Scholar
Carr, C. J. 1976. Plant ecological variation and pattern in the lower Omo basin. Pp. 432467 in Coppens et al. 1976.Google Scholar
Cerling, T. E. 1992. Development of grasslands and savannas in East Africa during the Neogene. Palaeogeography, Palaeoclimatology, Palaeoecology 97:241247.Google Scholar
Cerling, T. E. and Harris, J. M. 1999. Carbon isotope fractionation between diet and bioapatite in ungulate mammals and implications for ecological and paleoecological studies. Oecologia 120:347363.Google Scholar
Cooke, H. B S. 1976. Suidae from Plio-Pleistocene strata of the Rudolf basin. Pp. 251263 in Coppens et al. 1976.Google Scholar
Cooke, H. B S. 1978. Faunal evidence for the biotic setting of early African hominids. Pp. 267281 in Jolly, C. J., ed. Early hominids of Africa. Duckworth, London.Google Scholar
Cooke, H. B S. 1985. Pliocene-Pleistocene Suidae in relation to African hominid deposits. Pp. 101115 in Beden, M. et al., eds. L'environnement des hominidés au Plio-Pléistocène. Masson, Paris.Google Scholar
Coppens, Y. 1975. Evolution des hominidés et de leur environnements au cours du Plio-Pléistocène dans la basse vallée de l'Omo en Éthiopie. Comptes Rendus de l'Académie des Sciences D 281:16931696.Google Scholar
Coppens, Y. 1978. Evolution of the hominids and of their environment during the Plio-Pleistocene in the lower Omo valley, Ethiopia. Pp. 499506 in Bishop, W.W., ed. Geological background to fossil man. Scottish Academic Press, Edinburgh.Google Scholar
Coppens, Y. 1994. East side story: the origin of humankind. Scientific American 270(5):8895.Google Scholar
Coppens, Y. and Howell, F. C. 1976. Mammalian faunas of the Omo Group: Distributional and biostratigraphic aspects. Pp. 177192 in Coppens et al. 1976.Google Scholar
Coppens, Y. and Howell, F. C. 1985. Les faunes Plio-Pléistocène de la basse vallée de l'Omo (Éthiopie), Tome 1. Périssodactyles-Artiodactyles (Bovidae). Editions du CNRS, Paris.Google Scholar
Coppens, Y., Howell, F. C., Isaac, G. L., and Leakey, R. E. F., eds. 1976. Earliest man and environments in the Lake Rudolf basin. University of Chicago Press, Chicago.Google Scholar
Corruccini, R. S. and Ciochon, R. L. 1994. Integrative paths to the past: paleoanthropological advances in honor of F. Clark Howell. Prentice-Hall, Englewood Cliffs, N.J.Google Scholar
Coryndon, S. C. 1976. Fossil Hippopotamidae from Plio-Pleistocene successions of the Rudolf basin. Pp. 238250 in Coppens et al. 1976.Google Scholar
Dechamps, R. and Maes, F. 1985. Essai de reconstitution des climats et des végétations de la basse vallée de l'Omo au Plio-Pléistocène à l'aide de bois fossiles. Pp. 175221 in Beden, M. et al., eds. L'environnement des hominidés au Plio-Pléistocène. Masson, Paris.Google Scholar
Dechant Boaz, D. 1994. Taphonomy and the fluvial environment: examples from Pliocene deposits of the Shungura Formation, Omo Basin, Ethiopia. Pp. 377414 in Corruccini and Ciochon 1994.Google Scholar
de Heinzelin, J., ed. 1983. The Omo group: archives of the International Omo Research Expedition. Musée Royale de l'Afrique Centrale Annale, série in 8°. Sciences Géologiques No. 85. Tervuren, Belgium.Google Scholar
de Heinzelin, J. and Haesaerts, P. 1983. The Shungura Formation. Pp. 25127 in de Heinzelin 1983.Google Scholar
de Heinzelin, J., Haesaerts, P., and Howell, F. C. 1976. Plio-Pleistocene formations of the lower Omo basin, with particular reference to the Shungura Formation. Pp. 2449 in Coppens et al. 1976.Google Scholar
deMenocal, P. B. 1995. Plio-Pleistocene African climate. Science 270:5359.Google Scholar
deMenocal, P. B. and Bloemendal, J. 1995. Plio-Pleistocene climatic variability in subtropical Africa and the paleoenvironment of hominid evolution: a combined data-model approach. Pp. 262288 in Vrba et al. 1995.Google Scholar
Dorst, J. and Dandelot, P. 1969. A field guide to the larger mammals of Africa. Houghton Mifflin, Boston.Google Scholar
Eck, G. G. 1976. Cercopithecoidea from the Omo Group deposits. Pp. 332344 in Coppens et al. 1976.Google Scholar
Eck, G. G., Jablonski, N. G., and Leakey, M. 1987. Les faunes Plio-Pléistocène de la basse vallée de l'Omo (Éthiopie), Tome 3. Cercopithecidae de la formation de Shungura. Editions du CNRS, Paris.Google Scholar
Eisenmann, V. 1976. Equidae from the Shungura Formation. Pp. 225233 in Coppens et al. 1976.Google Scholar
Eisenmann, V. 1985. Les équidés des gisements de la vallée de l'Omo en Éthiopie. Pp. 1365 in Coppens and Howell 1985.Google Scholar
Estes, R. D. 1991. The behavior guide to African mammals. University of California Press, Berkeley.Google Scholar
Feibel, C. S., Brown, F. H., and McDougall, I. 1989. Stratigraphic context of fossil hominids from the Omo Group deposits: northern Turkana basin, Kenya and Ethiopia. American Journal of Physical Anthropology 78:595622.Google Scholar
Feibel, C. S., Harris, J. M., and Brown, F. H. 1991. Palaeoenvironmental context for the Late Neogene of the Turkana Basin. Pp. 321346 in Harris 1991.Google Scholar
Fisher, L. D. and van Belle, G. 1993. Biostatistics: a methodology for the health sciences. Wiley, New York.Google Scholar
Gentry, A. W. 1976. Bovidae of the Omo Group deposits. Pp. 275292 in Coppens et al. 1976.Google Scholar
Gentry, A. W. 1985. The Bovidae of the Omo Group deposits, Ethiopia. Pp. 119191 in Coppens and Howell 1985.Google Scholar
Geraads, D. 1995. Simatherium shungurense n. sp., un nouveau Bovini (Artiodactyla, Mammalia) du Pliocène terminal de l'Omo (Éthiopie). Annales de Paleontologie 81:8796.Google Scholar
Grattard, J. L., Howell, F. C., and Coppens, Y. 1976. Remains of Camelus from the Shungura Formation, Ethiopia. Pp. 268274 in Coppens et al. 1976.Google Scholar
Gray, B. T. 1980. Environmental reconstruction of the Hadar Formation (Hadar, Ethiopia). Ph.D. dissertation. Case Western Reserve University, Cleveland.Google Scholar
Grayson, D. K. 1984. Quantitative zooarchaeology. Academic Press, Orlando.Google Scholar
Greenacre, M. J. 1993. Correspondence analysis in practice. Academic Press, London.Google Scholar
Greenacre, M. J. and Vrba, E. S. 1984. Graphical display and interpretation of antelope census data in African wildlife areas, using correspondence analysis. Ecology 65:984997.Google Scholar
Guérin, C. 1976. Rhinocerotidae and Chalicotheriidae (Mammalia, Perissodactyla) from the Shungura Formation, lower Omo Basin. Pp. 214221 in Coppens et al. 1976.Google Scholar
Guérin, C. 1985. Les rhinocéros et les chalicothères (Mammalia, Perissodactyla) des gisements de la vallée de l'Omo en Éthiopie. Pp. 6789 in Coppens and Howell 1985.Google Scholar
Haesaerts, P., Stoops, G., and van Vliet-Lanoe, B. 1983. Data on sediments and fossil soils. Pp. 149185 in de Heinzelin 1983.Google Scholar
Harris, J. M. 1991. Koobi Fora Research Project, Vol. 3. The fossil ungulates: geology, fossil artiodactyls, and palaeoenvironments. Clarendon, 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 (Los Angeles) No. 399.Google Scholar
Hayek, L. A. and Buzas, M. A. 1997. Surveying natural populations. Columbia University Press, New York.Google Scholar
Hillman, J. C. 1986. Aspects of the biology of the bongo antelope, Tragelaphus euryceros Ogilby 1837, in southwest Sudan. Biological Conservation 38:255272.Google Scholar
Hooijer, D. A. and Churcher, C. S. 1985. Perissodactyla of the Omo Group deposits. Pp. 97117 89 in Coppens and Howell 1985.Google Scholar
Howell, F. C. 1968. Omo research expedition. Nature 219:567572.Google Scholar
Howell, F. C. 1978. Overview of the Pliocene and earlier Pleistocene of the lower Omo basin, southern Ethiopia. Pp. 85130 in Jolly, C. J., ed. Early hominids of Africa. Duckworth, London.Google Scholar
Howell, F. C. and Coppens, Y. 1976. An overview of Hominidae from the Omo succession, Ethiopia. Pp. 522532 in Coppens et al. 1976.Google Scholar
Howell, F. C. and Petter, G. 1976. Carnivora from Omo Group Formations, southern Ethiopia. Pp. 314331 in Coppens et al. 1976.Google Scholar
Howell, F. C., Haesaerts, P., and Heinzelin, J. de. 1987. Depositional environments, archeological occurrences and hominids from Members E and F of the Shungura Formation (Omo basin, Ethiopia). Journal of Human Evolution 16:665700.Google Scholar
Jaeger, J. J. and Wesselman, H. B. 1976. Fossil remains of micromammals from the Omo Group deposits. Pp. 351360 in Coppens et al. 1976.Google Scholar
Kennett, J. P. 1995. A review of polar climatic evolution during the Neogene, based on the marine sediment record. Pp. 4964 in Vrba et al. 1995.Google Scholar
Kimbel, W. H., Walter, R. C., Johanson, D. C., Reed, K. E., Aronson, J. L., Assefa, Z., Marean, C., Eck, G. G., Bobe, R., Hovers, E., Rak, Y., Vondra, C., Yemane, T., York, D., Chen, Y., Evensen, N. M., and Smith, P. E. 1996. Late Pliocene Homo and Oldowan tools from the Hadar Formation (Kada Hadar Member), Ethiopia. Journal of Human Evolution 31:549561.Google Scholar
Kingdon, J. 1982. East African mammals: an atlas of evolution in Africa, Vol. III. Parts C and D, Bovids. Academic Press, London.Google Scholar
Kingdon, J. 1989. Island Africa: the evolution of Africa's rare animals and plants. Princeton University Press, Princeton, N.J.Google Scholar
Kukla, G. 1987. Loess stratigraphy in central China. Quaternary Science Reviews 16:191219.Google Scholar
Lind, E. M. and Morrison, M. E. 1974. East African vegetation. Longman, London.Google Scholar
Ludwig, A. J. and Reynolds, J. F. 1988. Statistical ecology: a primer on methods and computing. Wiley, New York.Google Scholar
Magurran, A. E. 1988. Ecological diversity and its measurement. Princeton University Press, Princeton, N.J.Google Scholar
May, R. M. 1975. Patterns of species abundance and diversity. Pp. 81120 in Cody, M. L. and Diamond, J. M, eds. Ecology and evolution of communities. Belknap Press of Harvard University Press, Cambridge.Google Scholar
McDoughall, I. 1985. K-Ar and 40Ar/39Ar dating of the hominid-bearing Pliocene-Pleistocene sequence at Koobi Fora, Lake Turkana, northern Kenya. Geological Society of America Bulletin 96:159175.Google Scholar
McDougall, I., Davies, T., Maier, R., and Rudowski, R. 1985. Age of the Okote Tuff Complex at Koobi Fora, Kenya. Nature 316:792794.Google Scholar
McGougall, I., Brown, F. H., Cerling, T. E., and Hillhouse, J. W. 1992. A reappraisal of the geomagnetic polarity time scale to 4 Ma using data from the Turkana Basin, East Africa. Geophysical Research Letters 19:23492352.Google Scholar
Mloszewski, M. J. 1983. The behavior and ecology of the African buffalo. Cambridge University Press, Cambridge.Google Scholar
Murray, M. G. 1981. Structure of association in impala, Aepyceros melampus . Behavioral Ecology and Sociobiology 9:2333.Google Scholar
Reed, K. E. 1997. Early hominid evolution and ecological change through the African Plio-Pleistocene. Journal of Human Evolution 32:289322.Google Scholar
Shackleton, N. J. 1995. New data on the evolution of Pliocene climatic variability. Pp. 243248 in Vrba et al. 1995.Google Scholar
Shackleton, N. J., Backman, J., Zimmerman, H., Kent, D. V., Hall, M. A., Roberts, D. J., Schnitker, D., Baldauf, J. G., Desprairies, A., Homrighausen, R., Huddlestun, P., Keene, J. B., Kaltenback, A. J., Krumsiek, K. A., Morton, A. C., Murray, J. W., and Westberg-Smith, J. 1984. Oxygen isotope calibration of the onset of ice rafting and history of glaciation in the North Atlantic region. Nature 307:620623.Google Scholar
Shipman, P. and Harris, J. 1988. Habitat preference and paleoecology of Australopithecus boisei in Eastern Africa. Pp. 343381 in Grine, F. E., ed. Evolutionary history of the “robust” australopithecines. Aldine, New York.Google Scholar
Sinclair, A. R E. 1977. The African buffalo. University of Chicago Press, Chicago.Google Scholar
Spencer, L. M. 1997. Dietary adaptations of Plio-Pleistocene Bovidae: implications for hominid habitat use. Journal of Human Evolution 32:201228.Google Scholar
Spinage, C. A. 1982. A territorial antelope: the Uganda waterbuck. Academic Press, New York.Google Scholar
Spinage, C. A. 1986. The natural history of antelopes. Facts on File, New York.Google Scholar
Sponheimer, M., Reed, K. E., and Lee-Thorp, J. A. 1999. Combining isotopic and ecomorphological data to refine bovid paleodietary reconstruction: a case study from the Makapansgat Limeworks hominid locality. Journal of Human Evolution 36:705718.Google Scholar
Suwa, G., White, T. D., and Howell, F. C. 1996. Mandibular postcanine dentition from the Shungura Formation, Ethiopia: crown morphology, taxonomic allocations, and Plio-Pleistocene hominid evolution. American Journal of Physical Anthropology 101:247282.Google Scholar
Vrba, E. S. 1980a. The significance of bovid remains as indicators of environment and predation patterns. Pp. 247271 in Behrensmeyer, A. K. and Hill, A. P., eds. Fossils in the making. University of Chicago Press, Chicago.Google Scholar
Vrba, E. S. 1980b. Evolution, species, and fossils: how does life evolve? South African Journal of Science 76:6184.Google Scholar
Vrba, E. S. 1984. Evolutionary pattern and process in the sister-group Alcelaphini-Aepycerotini (Mammalia: Bovidae). Pp. 6279 in Eldredge, N. and Stanley, S. M., eds. Living fossils. Springer, New York.Google Scholar
Vrba, E. S. 1985a. Ecological and adaptive changes associated with early hominid evolution. Pp. 6371 in Delson, E. ed. Ancestors: the hard evidence. Liss, New York.Google Scholar
Vrba, E. S. 1985b. Paleoecology of early Hominidae, with special reference to Sterkfontein, Swartkrans and Kromdraai. Pp. 345369 in Beden, M. et al., eds. L'environnement des hominidés au Plio-Pléistocène. Masson, Paris.Google Scholar
Vrba, E. S. 1987. Ecology in relation to speciation rates: some case histories of Miocene-Recent mammals clades. Evolutionary Ecology 1:283300.Google Scholar
Vrba, E. S. 1988. Late Pliocene climatic events and hominid evolution. Pp. 405426 in Grine, F. E., ed. Evolutionary history of the “robust” australopithecines. Aldine, New York.Google Scholar
Vrba, E. S. 1995. The fossil record of African antelopes (Mammalia, Bovidae) in relation to human evolution and paleoclimate. Pp. 385424 in Vrba et al. 1995.Google Scholar
Vrba, E. S. 1996. Climate, heterochrony, and human evolution. Journal of Anthropological Research 52:128.CrossRefGoogle Scholar
Vrba, E. S., Denton, G. H., Partridge, T. C., Burckle, L. H., and eds. 1995. Paleoclimate and evolution, with emphasis on human origins. Yale University Press, New Haven, Conn.Google Scholar
Walter, R. C. and Aronson, J. L. 1993. Age and source of the Sidi Hakoma Tuff, Hadar Formation, Ethiopia. Journal of Human Evolution 25:229240.Google Scholar
Wesselman, H. B. 1984. The Omo micromammals: systematics and paleoecology of early man sites from Ethiopia. Contributions to Vertebrate Evolution 7:1219.Google Scholar
Wesselman, H. B. 1995. Of mice and almost-men: regional paleoecology and human evolution in the Turkana basin. Pp. 356368 in Vrba et al. 1995.Google Scholar
White, F. 1983. The vegetation of Africa: a descriptive memoir to accompany Unesco/AETFAT/UNSO vegetation map of Africa. Unesco, Paris.Google Scholar
Williamson, P. G. 1985. Evidence of an early Plio-Pleistocene rainforest expansion in East Africa. Nature 315:487489.Google Scholar
Wing, S. L., Sues, H. D., Potts, R., DiMichele, W. A., and Behrensmeyer, A. K. 1992. Evolutionary paleoecology. Pp. 113 in Behrensmeyer, A. K., Damuth, J. D., DiMichele, W. A., Potts, R., Sues, H. D., and Wing, S. L., eds. Terrestrial ecosystems through time: evolutionary paleoecology of terrestrial plants and animals. University of Chicago Press, Chicago.Google Scholar
Yemane, T., Vondra, C. F., and Aronson, J. 1996. Provenance, tectonics, and climate change; Kada Hadar Member, Hadar Formation, Ethiopia. Geological Society of America Abstracts with Programs 28(6):72.Google Scholar
Zar, J. H. 1984. Biostatistical analysis, 2d ed. Prentice-Hall, Englewood Cliffs, N.J. Google Scholar