Skip to main content Accessibility help
×
Hostname: page-component-7bb8b95d7b-dtkg6 Total loading time: 0 Render date: 2024-09-22T23:39:24.921Z Has data issue: false hasContentIssue false

Volume References

Published online by Cambridge University Press:  19 May 2022

Sally C. Reynolds
Affiliation:
Bournemouth University
René Bobe
Affiliation:
University of Oxford
Get access

Summary

Image of the first page of this content. For PDF version, please use the ‘Save PDF’ preceeding this image.'
Type
Chapter
Information
Publisher: Cambridge University Press
Print publication year: 2022

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

Volume References

Abbate, E., Albianelli, A., Azzaroli, A., et al. (1998). A one-million-year-old Homo cranium from the Danakil (Afar) Depression of Eritrea. Nature 393(6684), 458460.CrossRefGoogle ScholarPubMed
Abbate, E., Woldehaimanot, B., Bruni, P., et al. (2004). Geology of the Homo-bearing Pleistocene Dandiero Basin (Buia region, Eritrean Danakil Depression). Rivista Italiana di Paleontologia e Stratigrafia 10, 534.Google Scholar
Abbate, E., Albianelli, A., Awad, A., et al. (2010). Pleistocene environments and human presence in the middle Atbara valley (Khashm El Girba, Eastern Sudan). Palaeogeography, Palaeoclimatology, Palaeoecology 292, 1234.Google Scholar
Abbate, E., Bruni, P., Landucci, F. and Pellicanò, G. (2012). Unusual ichnofossils in Homo erectus-bearing beds of the Pleistocene lake deposits in central-eastern Eritrea, East Africa. Palaios 27, 97104.Google Scholar
Adams, J.W. (2012a). A revised listing of fossil mammals from the Haasgat cave system ex-situ deposits (HGD), South Africa. Palaeontologia Electronica, 15(3), 188.Google Scholar
Adams, J.W. (2012b). Stable carbon isotope analysis of fauna from the Gondolin GD 2 fossil assemblage, South Africa. Annals of the Ditsong National Museum of Natural History, 2(1), 15.Google Scholar
Adams, J.W. (2012c). Craniodental and postcranial remains of the extinct porcupine Hystrix makapanensis Greenwood, 1958 (Rodentia: Hystricidae) from Gondolin, South Africa. Annals of the Ditsong National Museum of Natural History, 2(1), 717.Google Scholar
Adams, J.W. and Rovinsky, D.S. (2018). Taphonomic interpretations of the Haasgat HGD assemblage: A case study in the impact of sampling and preparation methods on reconstructing South African karstic assemblage formation. Quaternary International, 495, 418.Google Scholar
Adams, J.W., Hemingway, J., Kegley, A.D.T. and Thackeray, J.F. (2007). Luleche, a new paleontological site in the Cradle of Humankind, North-West Province, South Africa. Journal of Human Evolution, 53(6), 751754.Google Scholar
Adams, J.W., Herries, A.I.R., Hemingway, J., et al (2010). Initial fossil discoveries from Hoogland, a new Pliocene primate-bearing karstic system in Gauteng Province, South Africa. Journal of Human Evolution, 59(6), 685.CrossRefGoogle Scholar
Adams, J.W., Kegley, A.D. and Krigbaum, J. (2013). New faunal stable carbon isotope data from the Haasgat HGD assemblage, South Africa, including the first reported values for Papio angusticeps and Cercopithecoides haasgati. Journal of Human Evolution, 64(6), 693.CrossRefGoogle ScholarPubMed
Adams, J.W., Olah, A., McCurry, M., et al. (2015). Newly discovered in situ primates and mammals from the early Pleistocene Haasgat deposits, South Africa. American Journal of Physical Anthropology, 156, 66.Google Scholar
Adams, J.W., Rovinsky, D.S., Herries, A.I. and Menter, C.G. (2016). Macromammalian faunas, biochronology and palaeoecology of the early Pleistocene Main Quarry hominin-bearing deposits of the Drimolen Palaeocave System, South Africa. PeerJ, 4, e1941.CrossRefGoogle ScholarPubMed
Adelsberger, K.A., Wirth, K.R., Mabulla, A.Z.P. and Bowman, D.C. (2011). Geochemical and mineralogical characterization of Middle Stone Age tools of Laetoli, Tanzania and comparisons with possible source materials. In: Harrison, T. (Ed.), Paleontology and Geology of Laetoli, Tanzania: Human Evolution in Context. Volume 1: Geology, Geochronology, Paleoecology and Paleoenvironment. Dordrecht: Springer, pp. 143165.CrossRefGoogle Scholar
Agnew, N. and Demas, M. (1998). Preserving the Laetoli footprints. Scientific American 279, 4455.CrossRefGoogle Scholar
Agnew, N., Demas, M. and Leakey, M.D. (1996). The Laetoli footprints. Science 271, 16511652.CrossRefGoogle ScholarPubMed
Agrawal, R., Imieliński, T. and Swami, A. (1993). Mining association rules between sets of items in large databases. In: Proceedings of the 1993 ACM SIGMOD International Conference on Management of Data. New York: ACM, pp. 207216.CrossRefGoogle Scholar
Aguirre, E. and Leakey, P. (1974). Nakali: nueva fauna de Hipparion del Rift Valley de Kenya. Estudios Geológicos 30, 219227.Google Scholar
Agustí, J., Garcés, M. and Krijgsman, W. (2006). Evidence for African–Iberian exchanges during the Messinian in the Spanish mammalian record. Palaeogeography, Palaeoclimatology, Palaeoecology, 238(1), 514.Google Scholar
Agustí, J., Cabrera, L. and Garcés, M. (2013). The Vallesian mammal turnover: A late Miocene record of decoupled land-ocean evolution. Geobios, 46(1), 151157.CrossRefGoogle Scholar
Aiello, L.C. and Wheeler, P. (1995). The expensive-tissue hypothesis: the brain and the digestive system in human and primate evolution. Current Anthropology 36, 199221.Google Scholar
Alba, D.M., Delson, E., Carnevale, G., et al. (2014). First joint record of Mesopithecus and cf. Macaca in the Miocene of Europe. Journal of Human Evolution 67, 118.Google Scholar
Albert, R.M. and Bamford, M.K. (2012). Vegetation during UMBI and deposition of Tuff IF at Olduvai Gorge, Tanzania (ca. 1.8 Ma), based on phytoliths and plant remains. Journal of Human Evolution 63, 342350.CrossRefGoogle ScholarPubMed
Albert, R.M., Bamford, M.K., Stanistreet, I., et al. (2015). Vegetation landscape at DK locality, Olduvai Gorge, Tanzania. Palaeogeography, Palaeoclimatology, Palaeoecology 426, 3445.CrossRefGoogle Scholar
Alemseged, Z. (1998). L’hominidé Omo-323: sa position phylétique et son environnement dans le cadre de l’évolution des communautés de mammiféres du Plio-Pléistocène dans la basse vallée de l’Omo (Ethiopie). PhD dissertation, Anthropologie, Muséum National d’Histoire Naturelle, Paris.Google Scholar
Alemseged, Z. (2003). An integrated approach to taphonomy and faunal change in the Shungura Formation (Ethiopia) and its implication for hominid evolution. Journal of Human Evolution 44(4), 451478.Google Scholar
Alemseged, , and Geraads, D. (1998). Theropithecus atlanticus (Cercopithecidae, Mammalia) from the late Pliocene of Ahl al Oughlam, Casablanca, Morocco. Journal of Human Evolution 34, 609621.Google Scholar
Alemseged, Z. and Geraads, D. (2000). A new Middle Pleistocene fauna from the Busidima-Telalak region of the Afar, Ethiopia. Comptes Rendu de l’Académie des Sciences Paris 331, 459556.Google Scholar
Alemseged, Z., Coppens, Y. and Geraads, D. (2002). Hominid cranium from Omo: description and taxonomy of Omo-323–1976-896. American Journal of Physical Anthropology, 117, 103112.Google Scholar
Alemseged, Z., Wynn, J. G., Kimbel, W. H., et al. (2005). A new hominin from the Basal Member of the Hadar Formation, Dikika, Ethiopia, and its geological context. Journal of Human Evolution 49(4), 499514.CrossRefGoogle ScholarPubMed
Alemseged, Z., Spoor, F., Kimbel, W.H., et al. (2006). A juvenile early hominin skeleton from Dikika, Ethiopia. Nature 443, 296301.Google Scholar
Alemseged, Z., Bobe, R. and Geraads, D. (2007). Comparability of fossil data and its significance for the interpretation of hominin environments. A case study in the lower Omo Valley, Ethiopia. In: Behrensmeyer, A.K., Bobe, R. and Alemseged, Z. (Eds.), Hominin Environments in the East African Pliocene: An Assessment of the Faunal Evidence. Dordrecht: Springer, pp. 159181.Google Scholar
Alemseged, Z., Wynn, J.G., Geraads, D., et al. (2020). Fossils from Mille-Logya, Afar, Ethiopia, elucidate the link between Pliocene environmental changes and Homo origins. Nature Communications 11, 2480.Google Scholar
Albianelli, A. and Napoleone, G. (2004). Magnetostratigraphy of the Homo-bearing Pleistocene Dandiero Basin (Danakil Depression, Eritrea). Rivista Italiana di Stratigrafia e Paleontologia 110(Supplement), 3544.Google Scholar
Alroy, J. (1998). Equilibrial diversity dynamics in North American mammals. In: McKinney, M.L. and Drake, J. (Eds.), Biodiversity Dynamics: Turnover of Populations, Taxa, and Communities. New York: Columbia University Press, pp. 232287.Google Scholar
Alroy, J. (2000). New methods for quantifying macroevolutionary patterns and processes. Paleobiology 26, 707733.2.0.CO;2>CrossRefGoogle Scholar
Almécija, S., Tallman, M., Alba, D.M., et al. (2013) The femur of Orrorin tugenensis exhibits morphometric affinities with both Miocene apes and later hominins. Nature Communications, 4(1), 112.Google Scholar
Altamura, F., Gaudzinski-Windeuser, S., Melis, R.T. and Mussi, M. (2019). Reassessing hominin skills at an early middle Pleistocene hippo butchery site: Gombore II-2 (Melka Kunture, Upper Awash valley, Ethiopia). Journal of Paleolithic Archaeology 3, 132. https://doi.org/10.1007/s41982-019-00046-0CrossRefGoogle Scholar
Altamura, F., Bennett, M.R., Marchetti, L., et al. (2020). Ichnological and archaeological evidence from Gombore II OAM, Melka Kunture, Ethiopia: an integrated approach to reconstruct local environments and biological presences between 1.2–0.85 Ma. Quaternary Science Reviews 244, 106506. https://doi.org/10.1016/j.quascirev.2020.106506Google Scholar
Amani, F. and Geraads, D. (1993). Le gisement moustérien du Djebel Irhoud, Maroc: précisions sur la faune et la biochronologie, et description d’un nouveau reste humain. Comptes Rendu de l’Académie des Sciences Paris 316, 847852.Google Scholar
Amani, F. and Geraads, D. (1998). Le gisement Moustérien du Djebel Irhoud, Maroc: précisions sur la faune et la paléoecologie. Bulletin d’Archéologie Marocaine 18, 1117.Google Scholar
Ambrose, S.H. (1998). Chronology of the Later Stone Age and food production in East Africa. Journal of Archaeological Science 25, 377392.Google Scholar
Ambrose, S.H., Hlusko, L.J., Kyule, D., Deino, A. and Williams, M. (2003). Lemudong’o: a new 6 Ma paleontological site near Narok, Kenya Rift Valley. Journal of Human Evolution 44, 737742.Google Scholar
Ambrose, S.H., Bell, C.J., Bernor, R.L., et al. (2007a). The paleoecology and paleogeographic context of Lemudong’o Locality 1, a late Miocene terrestrial fossil site in southern Kenya. Kirtlandia 56, 3852.Google Scholar
Ambrose, S.H., Nyamai, C.M., Mathu, E.M. and Williams, M.A. (2007b). Geology, geochemistry, and stratigraphy of the Lemudong’o Formation, Kenya Rift Valley. Kirtlandia 56, 5364.Google Scholar
Ambrose, S.H., WoldeGabriel, G., White, T.D. and Suwa, G. (2011). The role of paleosol carbon isotopes in reconstructing the Aramis Ardipithecus ramidus habitat: woodland or grassland. PaleoAnthropology 2011, A1.Google Scholar
Ameur, R., Jaeger, J.-J. and Michaux, J. (1976). Radiometric age of early Hipparion in North-West Africa. Nature 261, 3839.Google Scholar
Andrews, P. (1983). Small faunal diversity at Olduvai Gorge. In: Clutton-Brock, J. and Grigson, C. (Eds.), Animals and Archaeology: 1. Hunters and their Prey. Oxford: BAR International Series 163.Google Scholar
Andrews, P. (1989). Palaeoecology of Laetoli. Journal of Human Evolution 18, 173181.Google Scholar
Andrews, P. (1990). Owls, Caves, and Fossils. London: Natural History Museum.Google Scholar
Andrews, P. (2006). Taphonomic effects of faunal impoverishment and faunal mixing. Palaeogeography, Palaeoclimatology, Palaeoecology 241, 572589.Google Scholar
Andrews, P. and Bamford, M. (2008). Past and present ecology of Laetoli, Tanzania. Journal of Human Evolution 58, 7898.CrossRefGoogle Scholar
Andrews, P. and Evans, E.N. (1979). The environment of Ramapithecus in Africa. Paleobiology 5(1), 2230.CrossRefGoogle Scholar
Andrews, P. and Hixson, S. (2014). Taxon-free methods of palaeoecology. Annales Zoologici Fennici 51(1–2), 269284.CrossRefGoogle Scholar
Andrews, P.J. and Humphrey, L. (1999). African Miocene environments and the transition to early hominines. In: Bromage, T. and Schrenk, F. (Eds.), African Biogeography, Climate Change and Human Evolution. Oxford: Oxford University Press, pp. 282315.Google Scholar
Andrews, P. and O’Brien, E. (2000). Climate, vegetation, and predictable gradients in mammal species richness in southern Africa. Journal of Zoology 251, 205231.CrossRefGoogle Scholar
Andrews, P. and O’Brien, E. (2010). Mammal species richness in Africa. In Werdelin, L. and Sanders, W. (Eds.), Cenozoic Mammals of Africa. New York: Columbia University Press, pp. 929947.Google Scholar
Andrews, P., Lord, J.M. and Evans, E.M.N. (1979). Patterns of ecological diversity in fossil and modern mammalian faunas. Biological Journal of the Linnean Society 11, 177205.Google Scholar
Andrews, P., Bamford, M.K., Njau, E.-F. and Leliyo, G. (2011). The ecology and biogeography of the Endulen–Laetoli area in northern Tanzania. In: Harrison, T. (Ed.), Paleontology and Geology of Laetoli: Human Evolution in Context: Volume 1: Geology, Geochronology, Paleoecology and Paleoenvironment. Dordrecht: Springer, pp. 167200.Google Scholar
Antón, S.C. (2012). Early Homo: who, when, and where. Current Anthropology 53(S6), S278S298.Google Scholar
Antón, S.C. (2013). Homo erectus and related taxa. In: Begun, D.R. (Ed.), A Companion to Paleoanthropology. Pondicherry: Blackwell Publishing Ltd, pp. 497516.Google Scholar
Antón, S.C., Potts, R. and Aiello, L.C. (2014). Evolution of early Homo: an integrated biological perspective. Science 345(6192), 1236828.Google Scholar
Aouadi, N., Dridi, Y. and Dhia, W.B. (2014). Holocene environment and subsistence patterns from Capsian and Neolithic sites in Tunisia. Quaternary International 320, 314.CrossRefGoogle Scholar
Aouraghe, H. (2000). Les carnivores fossiles d’El Harhoura 1, Temara, Maroc. L’anthropologie 104, 147171.CrossRefGoogle Scholar
Aouraghe, H. (2004). Les populations de mammifères atériens d’El Harhoura 1 (Témara, Maroc). Bulletin d’Archéologie Marocaine 20, 83104.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. 2. ptie. Paléontologie. Paris: Éditions du Muséum.Google Scholar
Arambourg, C. (1948). Contribution à l’étude géologique et paléontologique du bassin du lac Rudolphe et de la basse vallée de l’Omo. Paléontologie. Mission scientifique de l’Omo, Fasc. 3 (Géologie). Paris: Éditions du Muséum, pp. 231562.Google Scholar
Arambourg, C. (1949). Sur la présence, dans le Villafranchien d’Algérie, de vestiges éventuels d’industrie humaine. Comptes-rendus de l’Académie des Sciences 229, 6667.Google Scholar
Arambourg, C. (1954). L’Hominien fossile de Ternifine (Algérie). Comptes-rendus de l’Académie des Sciences 239, 893895.Google Scholar
Arambourg, C. (1959). Vertébrés continentaux du Miocène supérieur de l’Afrique du Nord. Publications du Service Géologique de l’Algérie, NS, Paléontologie 4, 1159.Google Scholar
Arambourg, C. (1961). Note préliminaire sur quelques Vertébrés nouveaux du Burdigalien de Libye. Comptes rendus sommaires des séances de la Société géologique de France 4, 107109.Google Scholar
Arambourg, C. (1963a). Le gisement de Ternifine. L’Atlanthropus mauritanicus. Archives de l’Institut de Paléontologie Humaine 32, 37190.Google Scholar
Arambourg, C. (1963b). Continental vertebrate faunas of the Tertiary of North Africa. African Ecology and Human Evolution 36, 55.Google Scholar
Arambourg, C. (1967). Appendix A. Observations sur la faune des Grottes d’Hercule près de Tanger, Maroc. In: Howe, B. (Ed.), The Palaeolithic of Tangier, Morocco: Excavations at Cape Ashakara, 1939–1947. Cambridge, MA: The Peabody Museum, pp. 181186.Google Scholar
Arambourg, C. (1970). Les vertébrés du Pléistocène de l’Afrique du Nord. Archives du Muséum National d’Histoire Naturelle, ser. 7, 10, 1126.Google Scholar
Arambourg, C. (1979). Vertébrés villafranchiens d’Afrique du Nord (Artiodactyles, Carnivores, Primates, Reptiles, Oiseaux). Paris: Fondation Singer-Polignac, 141 pp.Google Scholar
Arambourg, C. and Biberson, P. (1955). Découverte de vestiges humains acheuléens dans la carrière de Sidi Abderrahman, près Casablanca. Comptes-rendus de l’Académie des Sciences 240, 16611663.Google Scholar
Arambourg, C. and Coppens, Y. (1968). Découverte d’un australopithécien nouveau dans les gisements de l’Omo (Éthiopie). South African Journal of Science, 64, 5859.Google Scholar
Arambourg, C. and Hoffstetter, R. (1963). Le gisement de Ternifine. Historique et géologie. Archives de l’Institut de Paléontogie Humaine, 32, 936.Google Scholar
Archer, W., Aldeias, V. and McPherron, S.P. (2020). What is ‘in situ’? A reply to Harmand et al. (2015). Journal of Human Evolution 142, 102740.Google Scholar
Arènes, J. and Depape, G. (1953). Etude paléobotanique. Contribution à l’étude des flores fossiles de l’Afrique du Nord. Archives du Muséum National d’Histoire Naturelle, ser. 7 2, 185.Google Scholar
Armour-Chelu, M. and Bernor, R.L. (2011). Equidae. In: Harrison, T. (Ed.), Paleontology and Geology of Laetoli: Human Evolution in Context. Vol. 2: Fossil Hominins and the Associated Fauna. Dordrecht: Springer, pp. 295326.CrossRefGoogle Scholar
Armour-Chelu, M., Bernor, R.L. and Mittman, H.-W. (2006). Hooijer’s hypodigm for “Hipparion” cf. ethiopicum (Equidae, Hipparioninae) from Olduvai, Tanzania and comparative material from the East African Plio-Pleistocene. Beiträge zur Paläontologie 30, 1524.Google Scholar
Arnason, U. and Janke, A. (2002). Mitogenomic analyses of eutherian relationships. Cytogenetic and Genome Research 96(1–4), 2032.CrossRefGoogle ScholarPubMed
Aronson, J.L. and Hailemichael, M. (2010). Reply to “Is the Pliocene Ethiopian Monsoon extinct? A comment on Aronson et al. (2008)”. Journal of Human Evolution 59, 139142.Google Scholar
Aronson, J.L. and Taieb, M. (1981). Geology and paleogeography of the Hadar hominid site, Ethiopia. In: Rapp, G. and Vondra, C.F. (Eds.), Hominid Sites: Their Geologic Setting. Boulder: Westview Press, pp. 165195.Google Scholar
Aronson, J.L., Hailemichael, M. and Savin, S.M. (2008). Hominid environments at Hadar from paleosol studies in a framework of Ethiopian climate change. Journal of Human Evolution 55, 532550.Google Scholar
Asfaw, B. (1987). The Belohdelie frontal: new evidence of early hominid cranial morphology from the Afar of Ethiopia. Journal of Human Evolution 16(7–8), 611624.Google Scholar
Asfaw, B., Beyene, Y., Semaw, S., et al. (1991). Fejej: a new paleoanthropological research area in Ethiopia. Journal of Human Evolution 21, 137143.Google Scholar
Asfaw, B., Beyene, Y., Suwa, G., et al. (1992). The earliest Acheulean from Konso-Gardula. Nature 360, 732735.CrossRefGoogle ScholarPubMed
Asfaw, B., Beyene, Y., Semaw, S., et al. (1993). Tephra from Fejej, Ethiopia – a reply. Journal of Human Evolution 25, 519521.CrossRefGoogle Scholar
Asfaw, B., White, T., Lovejoy, O., et al. (1999). Australopithecus garhi: a new species of early hominid from Ethiopia. Science 284(5414), 629635.CrossRefGoogle ScholarPubMed
Asfaw, B., Gilbert, W.H., Beyene, Y., et al. (2002). Remains of Homo erectus from Bouri, Middle Awash, Ethiopia. Nature 416(6878), 317320.Google Scholar
Ashley, G.M. (2020). Paleo-Critical Zones, windows into the changing life and landscapes during the Quaternary Period. Quaternary Research 96, 5365.CrossRefGoogle Scholar
Ashley, G. and Hay, R.L. (2002). Sedimentation patterns in a Plio-Pleistocene volcaniclastic rift-platform basin, Olduvai Gorge, Tanzania. In: Renaut, R.W. and Ashley, G.M. (Eds.), Sedimentation in Continental Rifts. Tulsa: SEPM (Society for Sedimentary Geology), pp. 107122.CrossRefGoogle Scholar
Ashley, G.M. and Liutkus, C.M. (2002). Tracks, trails and trampling by large vertebrates in a rift valley paleo-wetland, lowermost Bed II, Olduvai Gorge, Tanzania. Ichnos 9(1–2), 2332.Google Scholar
Ashley, G.M., Tactikos, J.C. and Owen, R.B. (2009). Hominin use of springs and wetlands: paleoclimate and archaeological records from Olduvai Gorge (1.79–1.74 Ma). Palaeogeography, Palaeoclimatology, Palaeoecology 272, 116.Google Scholar
Ashley, G.M., Barboni, D., Dominguez-Rodrigo, M., et al. (2010). A spring and wooded habitat at FLK Zinj and their relevance to origins of human behavior. Quaternary Research 74, 304314.Google Scholar
Ashley, G.M., Bunn, H.T., Delaney, J.S., et al. (2014). Paleoclimatic and paleoenvironmental framework of FLK North archaeological site, Olduvai Gorge, Tanzania. Quaternary International 322–323, 5465.Google Scholar
Ashley, G.M., de Wet, C.B., Barboni, D. and Magill, C.R. (2016). Subtle signatures of seeps: record of groundwater in a Dryland, DK, Olduvai Gorge, Tanzania. The Depositional Record 2, 421.Google Scholar
Ashley, G.M., Barboni, D., Dominguez-Rodrigo, M., et al. (2017). Paleoenvironmental and paleoecological reconstruction of a freshwater oasis in savannah grassland at FLK North, Olduvai Gorge, Tanzania. Quaternary Research 74, 333343.CrossRefGoogle Scholar
Assefa, Z. (2006). Faunal remains from Porc-Epic: paleoecological and zooarchaeological investigations from a Middle Stone Age site in southeastern Ethiopia. Journal of Human Evolution 51, 5075.Google Scholar
Assefa, Z., Yirga, S. and Reed, K.E. (2008). The large-mammal fauna from the Kibish Formation. Journal of Human Evolution 55, 501512.Google Scholar
Au ffenberg, W. (1981). The fossil turtles of Olduvai Gorge, Tanzania, Africa. Copeia 1981, 509522.CrossRefGoogle Scholar
Avery, D.M. (1981). Holocene micromammalian faunas from the Northern Cape, South Africa. South African Journal of Science 77, 265273.Google Scholar
Avery, D.M. (1982a). Micromammals as palaeoenvironmental indicators and an interpretation of the late quaternary in the Southern Cape Province, South Africa. Annals of the South African Museum 85, 183374.Google Scholar
Avery, D.M. (1982b). The micromammalian fauna from Border Cave, Kwazulu, South Africa. Journal of Archaeological Science 9(2), 187204.CrossRefGoogle Scholar
Avery, D.M. (1987a). Micromammalian evidence for natural vegetation and the introduction of farming during the Holocene in the Magaliesberg, Transvaal. South African Journal of Science 83, 221225.Google Scholar
Avery, D.M. (1987b). Late Pleistocene coastal environment of the Southern Cape province of South Africa: micromammals from Klasies river mouth. Journal of Archaeological Science 14(4), 405421.Google Scholar
Avery, D.M. (1990). Holocene climatic change in Southern Africa: the contribution of micromammals to its study. South African Journal of Science 86, 407412.Google Scholar
Avery, D.M. (1991). Micromammals, owls and vegetation change in the Eastern Cape Midlands, South Africa, during the last millennium. Journal of Arid Environments 20, 357369.CrossRefGoogle Scholar
Avery, D.M. (1992). The environment of early modern humans at Border Cave, South Africa: micromammalian evidence. Palaeogeography, Palaeoclimatology, Palaeoecology 91, 7187.Google Scholar
Avery, D.M. (1995a). Southern savannahs and Pleistocene hominid adaptations: the micromammalian perspective. In: Vrba, E., Denton, G., Partridge, T. and Burckle, L. (Eds.), Paleoclimate and Evolution with Emphasis on Human Origins. New Haven: Yale University Press, pp. 459478.Google Scholar
Avery, D.M. (1995b). A preliminary assessment of the micromammalian remains from Gladysvale Cave, South Africa. Palaeontologia Africana 32, 110.Google Scholar
Avery, D.M. (1997). Micromammals and the Holocene environment of Rose Cottage Cave. South African Journal of Science 93, 445448.Google Scholar
Avery, D.M. (1998). An assessment of the lower Pleistocene micromammalian fauna from Swartkrans Members 1–3, Gauteng, South Africa. Geobios 31, 393414.Google Scholar
Avery, D.M. (2001). The Plio-Pleistocene vegetation and climate of Sterkfontein and Swartkrans, South Africa, based on micromammals. Journal of Human Evolution 41, 113132.Google Scholar
Avery, D.M. (2007). Pleistocene micromammals from Wonderwerk Cave, South Africa: practical issues. Journal of Archaeological Science 34, 613625.Google Scholar
Avery, D.M. (2019). A Fossil History of Southern African Land Mammals. Cambridge: Cambridge University Press.Google Scholar
Avery, D.M. (2022). Rodents and other micromammals from the Pleistocene strata in Excavation 1 at Wonderwerk Cave, South Africa: a work in progress. Quaternary International 614, 23–36.CrossRefGoogle Scholar
Avery, D.M. and Avery, G. (2011). Micromammals in the Northern Cape Province of South Africa, past and present. African Natural History 7, 339.Google Scholar
Avery, S.T. and Tebbs, E.J. (2018). Lake Turkana, major Omo River developments, associated hydrological cycle change and consequent lake physical and ecological change. Journal of Great Lakes Research 44, 11641182.CrossRefGoogle Scholar
Avery, D.M., Stratford, D.J. and Sénégas, F. (2010). Micromammals and the formation of the Name Chamber, at Sterkfontein, South Africa. Geobios, 43(4), 379387.Google Scholar
Avery, G., Kandel, A.W., Klein, R.G., Conard, N.J. and Cruz-Uribe, K. (2004). Tortoises as food and taphonomic elements in palaeo ‘landscapes’. In: Brugal, J.-P. and Desse, J. (Eds.), Petits animaux et sociétés humaines: du complément alimentaire aux ressources utilitaires. Antibes: Editions APDCA, pp. 147161.Google Scholar
Azaïs, F.B. and Chambard, R. (1931). Cinq années de recherches archéologiques en Ethiopie. Paris: Librairie Orientaliste Paul Geuthner.Google Scholar
Ba, K., Granjon, L., Hutterer, R., and Duplantier, J.-M. (2000). Les micromammifères du Djoudj (Delta du Sénégal) par l’analyse du régime alimentaire de la chouette effraie, Tyto alba. Bonner zoologische Beiträge 49, 3138.Google Scholar
Bache, F., Popescu, S.M., Rabineau, M., et al. (2012). A two‐step process for the reflooding of the Mediterranean after the Messinian Salinity Crisis. Basin Research 24(2), 125153.Google Scholar
Bachiri Taoufiq, N., Barhoun, N. and Suc, J.-P. (2008). Les environnements continentaux du corridor rifain (Maroc) au Miocène supérieur d’après la palynologie. Geodiversitas 30, 4158.Google Scholar
Backwell, L.R. and d’Errico, F. (2001). Evidence of termite foraging by Swartkrans early hominids. Proceedings of the National Academy of Sciences 98(4), 13581363.Google Scholar
Backwell, L.R. and d’Errico, F. (2004). Swartkrans; a cave’s chronicle of early man. Additional evidence of early hominid bone tools from Swartkrans. Transvaal Museum Monographs 8(1).Google Scholar
Backwell, L. and d’Errico, F. (2008). Early hominid bone tools from Drimolen, South Africa. Journal of Archaeological Science 35(11), 28802894.Google Scholar
Backwell, L., Pickering, R., Brothwell, D., et al. (2009). Probable human hair found in a fossil hyaena coprolite from Gladysvale cave, South Africa. Journal of Archaeological Science 36(6), 12691276.Google Scholar
Backwell, L.R., Parkinson, A.H., Roberts, E.M., d’Errico, F. and Huchet, J.-B. (2012). Criteria for identifying bone modification by termites in the fossil record. Paleogeography, Paleoclimatology, Paleoecology 337–338, 7287.Google Scholar
Backwell, L., Huchet, J.B., Jashashvili, T., Dirks, P.H. and Berger, L.R. (2020). Termites and necrophagous insects associated with early Pleistocene (Gelasian) Australopithecus sediba at Malapa, South Africa. Palaeogeography, Palaeoclimatology, Palaeoecology 560, 109989.Google Scholar
Badenhorst, S. and Plug, I. (2012). The faunal remains from the Middle Stone Age levels of Bushman Rock Shelter in South Africa. South African Archaeological Bulletin 67, 1631.Google Scholar
Badenhorst, S. and Steininger, C.M. (2019). The Equidae from Cooper’s D, an early Pleistocene fossil locality in Gauteng, South Africa. PeerJ 7, e6909.Google Scholar
Badenhorst, S., Senegas, F., Gommery, D., et al. (2011). Pleistocene faunal remains from Garage Ravine Cave on Bolt’s Farm in the Cradle of Humankind, South Africa. Annals of the Ditsong National Museum of Natural History 1(1), 3340.Google Scholar
Badenhorst, S., van Niekerk, K.L. and Henshilwood, C.S. (2014). Rock hyraxes (Procavia capensis) from Middle Stone Age levels at Blombos Cave, South Africa. African Archaeological Review 31(1), 2543.Google Scholar
Bagley, B. and Nyblade, A.A. (2013). Seismic anisotropy in eastern Africa, mantle flow, and the African superplume. Geophysical Research Letters 40, 15001505.Google Scholar
Badgley, C. (1986). Taphonomy of mammalian fossil remains from Siwalik rocks of Pakistan. Paleobiology 12, 119142.Google Scholar
Bagtache, B., Hadjouis, D. and Eisenmann, V. (1984). Présence d’un Equus caballin (E. algericus n. sp.) et d’une autre espèce nouvelle d’Equus (E. melkiensis n. sp.) dans l’Atérien des Allobroges, Algérie. Comptes Rendus de l’Académie des Sciences 14, 609612.Google Scholar
Bailey, G.N. and King, G.C.P. (2011). Dynamic landscapes and human dispersal patterns: tectonics, coastlines, and the reconstruction of human habitats. Quaternary Science Review 30, 15331553.Google Scholar
Bailey, G.N., Reynolds, S.C. and King, G.C. (2011). Landscapes of human evolon: models and methods of tectonic geomorphology and the reconstruction of hominin landscapes. Journal of Human Evolution 60(3), 257280.CrossRefGoogle Scholar
Bailloud, G. (1965). Les gisements paléolithiques de Melka-Konture. Cahiers de l’Institut éthiopien d’archéologie 1, 137.Google Scholar
Bailón, S. (2000). Amphibiens et reptiles du Pliocène terminal d’Ahl al Oughlam (Casablanca, Maroc). Geodiversitas 22(4), 539558.Google Scholar
Bailón, S. (2016). Les Squamates. In: J.-P. Raynal and A. Mohib (Eds.), Préhistoire de Casablanca. 1 ‒ La Grotte des Rhinocéros (fouilles 1991 et 1996). Villes et Sites Archéologiques du Maroc, volume 6. Rabat: Ministère de la Culture, INSAP, p. 87.Google Scholar
Bajo, P., Drysdale, R.N., Woodhead, J.D., et al. (2020). Persistent influence of obliquity on ice age terminations since the Middle Pleistocene transition. Science 367, 1235.CrossRefGoogle ScholarPubMed
Baker, B.H. (1958). Geology of the Magadi area. Report No. 42, Geological Survey of Kenya.Google Scholar
Baker, B.H. (1963). Geology of the area south of Magadi. Degree Sheet 58, N.W. Quarter. Report No. 61, Geological Survey of Kenya.Google Scholar
Bakker, E.M.V.Z. and Mercer, J.H. (1986). Major late Cainozoic climatic events and palaeoenvironmental changes in Africa viewed in a world wide context. Palaeogeography, Palaeoclimatology, Palaeoecology 56(3), 217235.Google Scholar
Bamford, M. (1999). Pliocene fossil woods from an early hominid cave deposit, Sterkfontein, South Africa. South African Journal of Science 95, 231237.Google Scholar
Bamford, M.K. (2000). Cenozoic macro-plants. In: Partridge, T.C. and Maud, R.R. (Eds.), The Cenozoic of Southern Africa. Oxford: Oxford Monographs on Geology and Geophysics, 40, pp. 351356.Google Scholar
Bamford, M.K. (2005). Early Pleistocene fossil wood from Olduvai Gorge, Tanzania. Quaternary International 129, 1522.Google Scholar
Bamford, M. (2011a). Fossil woods. In: Harrison, T. (Ed.), Paleontology and Geology of Laetoli: Human Evolution in Context. Vol. 1: Geology, Geochronology, Paleoecology, and Paleoenvironment. Dordrecht: Springer, pp. 217233.Google Scholar
Bamford, M. (2011b). Fossil leaves, fruits and seeds. In: Harrison, T. (Ed.), Paleontology and Geology of Laetoli: Human Evolution in Context. Vol. 1: Geology, Geochronology, Paleoecology, and Paleoenvironment. Dordrecht: Springer, pp. 235252.Google Scholar
Bamford, M.K. (2012). Fossil sedges, macroplants and roots from Olduvai Gorge, Tanzania. Journal of Human Evolution 63, 351363.CrossRefGoogle ScholarPubMed
Bamford, M.K. (2015a). Macrobotanical remains from Wonderwerk Cave (Excavation 1), Oldowan to Late Pleistocene (2 Ma to 14ka BP), South Africa. African Archaeological Review 32, 813838.Google Scholar
Bamford, M.K. (2015b). Charcoal from pre-Holocene Stratum 5, Wonderwerk Cave, South Africa. Palaeoecology of Africa 33, 153174.Google Scholar
Bamford, M. (2017). Pleistocene fossil woods from the Okote Member, site FwJj 14 in the Ileret region, Koobi Fora Formation, northern Kenya. Journal of Human Evolution 112, 134147.Google Scholar
Bamford, M.K. and Henderson, Z.L. (2003). A reassessment of the wooden fragment from Florisbad, South Africa. Journal of Archaeological Science 30(6), 637650.Google Scholar
Bamford, M., Albert, R.M. and Cabanes, D. (2006). Plio-Pleistocene macroplant fossil remains and phytoliths from lowermost Bed II in the eastern palaeolake margin of Olduvai Gorge, Tanzania. Quaternary International 148, 95112.Google Scholar
Bamford, M.K., Stanistreet, I. G., Stollhofen, H. and Albert, R.M. (2008). Late Pliocene grassland from Olduvai Gorge, Tanzania. Palaeogeography, Palaeoclimatology, Palaeoecology 257, 280293.Google Scholar
Bamford, M.K., Neumann, F.H., Pereira, L.M., et al. (2010). Botanical remains from a coprolite from the Pleistocene hominin site of Malapa, Sterkfontein Valley, South Africa. Palaeontologica Africana 45, 2328.Google Scholar
Bamford, M.K., Senut, B. and Pickford, M. (2013). Fossil leaves from Lukeino, a 6-million-year-old formation in the Baringo Basin, Kenya. Geobios 46(4), 253272.CrossRefGoogle Scholar
Barboni, D. (2014). Vegetation of northern Tanzania during the Plio-Pleistocene: a synthesis of the paleobotanical evidences from Laetoli, Olduvai, and Peninj hominin sites. Quaternary International 322–323, 264276.Google Scholar
Barboni, D., Ashley, G.M., Dominguez-Rodrigo, M., et al. (2010). Phytoliths infer locally dense and heterogeneous paleovegetation at FLK North and surrounding localities during upper Bed I time, Olduvai Gorge, Tanzania. Quaternary Research 74, 344354.Google Scholar
Barbour, G.B. (1949a). Makapansgat. Science Monthly 69, 141147.Google Scholar
Barbour, G.B. (1949b). Ape or man? An incomplete chapter of human ancestry from South Africa. Ohio Journal of Science 49, 129145.Google Scholar
Bardin, G., Raynal, J.-P. and Kieffer, G. (2004). Drainage patterns and regional morphostructure at Melka Kunture (Upper Awash, Ethiopia). In: Chavaillon, J. and Piperno, M. (Eds.), Studies on the Early Paleolithic site of Melka Kunture, Ethiopia – 2004. Florence: Istituto Italiano di Preistoria e Protostoria, pp. 8392.Google Scholar
Barker, D.S. and Milliken, K.L. (2008). Cementation of the Footprint Tuff, Laetoli, Tanzania. The Canadian Mineralogist 46, 831841.Google Scholar
Bar-Matthews, M., Marean, C.W., Jacobs, Z., et al. (2010). A high resolution and continuous isotopic speleothem record of paleoclimate and paleoenvironment from 90 to 53 ka from Pinnacle Point on the south coast of South Africa. Quaternary Science Reviews 29(17–18), 21312145.Google Scholar
Barnosky, A.D. (2001). Distinguishing the effects of the red queen and court jester on Miocene mammal evolution in the northern Rocky Mountains. Journal of Vertebrate Paleontology 21, 172185.Google Scholar
Barnosky, A. (2005). Effects of Quaternary climatic change on speciation in mammals. Journal of Mammalian Evolution 12(1), 247264. doi: 10.1007/s10914-005-4858-8Google Scholar
Barnosky, A.D., Matzke, N., Tomiya, S., et al. (2011). Has the earth’s sixth mass extinction already arrived? Nature 471, 5157.CrossRefGoogle ScholarPubMed
Barr, W.A. (2015). Paleoenvironments of the Shungura Formation (Plio-Pleistocene: Ethiopia) based on ecomorphology of the bovid Astragalus. Journal of Human Evolution 88, 97107.CrossRefGoogle ScholarPubMed
Barr, W.A. (2017). Bovid locomotor functional trait distributions reflect land cover and annual precipitation in sub-Saharan Africa. Evolutionary Ecology Research 18(3), 253269.Google Scholar
Barrett, L., Gaynor, D., Rendall, D., Mitchell, D. and Henzi, S.P. (2004). Habitual cave use and thermoregulation in chacma baboons (Papio hamadryas ursinus). Journal of Human Evolution 46(2), 215222.Google Scholar
Barsky, D., Chapon-Sao, C., Bahain, J.-J., et al. (2011). The early Oldowan stone-tool assemblage from Fejej FJ-1A, Ethiopia. Journal of African Archaeology 9, 207224.CrossRefGoogle Scholar
Barthelme, J. (1991). Lenderut: a new Acheulian site in the southern Kenya rift. Nyame Akuma 35, 2124.Google Scholar
Basu, C., Falkingham, P.L. and Hutchinson, J.R. (2016). The extinct, giant giraffid Sivatherium giganteum: skeletal reconstruction and body mass estimation. Biology Letters 12.Google Scholar
Beadle, L.C. (1974). The Inland Waters of Tropical Africa. London: Longman.Google Scholar
Beaumont, P.B. (1982). Aspects of the Northern Cape Pleistocene project. In: Coetzee, J.A. and van Zinderen Bakker, E.M. (Eds.), Palaeoecology of Africa and the Surrounding Islands vol. 15. Cape Town: A.A. Balkema, pp. 4144.Google Scholar
Beaumont, P.B. (1990). Wonderwerk Cave. In: Beaumont, P. and Morris, D. (Eds.), Guide to Archaeological Sites in the Northern Cape. Kimberly: McGregor Museum, pp. 101134.Google Scholar
Beaumont, P.B. (2004). Wonderwerk Cave. In: Beaumont, P. and Morris, D. (Eds.), Archaeology in the Northern Cape: Some Key Sites. Kimberly: McGregor Museum, pp. 3136.Google Scholar
Beaumont, P.B. (2011). The edge: more on fire-making by about 1.7 million years ago at Wonderwerk Cave in South Africa. Current Anthropology 52(4), 585595.Google Scholar
Beaumont, P.B. and Vogel, J.C. (2006). On a timescale for the past million years of human history in central South Africa. South African Journal of Science 102, 217228.Google Scholar
Beauvilain, A. (2008). The contexts of discovery of Australopithecus bahrelghazali (Abel) and of Sahelanthropus tchadensis (Tounami): unearthed, embedded in sandstone, or surface collected? South African Journal of Science 104, 165168.Google Scholar
Beauvilain, A. and Le Guellec, Y. (2004). Further details concerning fossils attributed to Sahelanthropus tchadensis (Toumaï). South African Journal of Science 100, 142144.Google Scholar
Beck, C.C., Feibel, C.S., Wright, J.D. and Mortlock, R.A. (2019). Onset of the African Humid Period by 13.9 kyr BP at Kabua Gorge, Turkana Basin, Kenya. The Holocene 29, 10111019.Google Scholar
Bedaso, Z.K. (2011). Stable isotope studies of paleoenvironment and paleoclimate from Afar, Ethiopia. PhD dissertation, University of South Florida.Google Scholar
Bedaso, Z.K., Wynn, J.G., Alemseged, Z. and Geraads, D. (2010). Paleoenvironmental reconstruction of the Asbole fauna (Busidima Formation, Afar, Ethiopia) using stable isotopes. Geobios 43, 165177.Google Scholar
Bedaso, Z.K., Wynn, J.G., Alemseged, Z. and Geraads, D. (2013). Dietary and paleoenvironmental reconstruction using stable isotopes of herbivore tooth enamel from middle Pliocene Dikika, Ethiopia: implication for Australopithecus afarensis habitat and food resources. Journal of Human Evolution 64, 2138.Google Scholar
Beden, M. (1976). Proboscideans from Omo group deposits. In: Coppens, Y., Howell, F.C., Isaac, G.L.I. and Leakey, R.E.F. (Eds.), Earliest Man and Environments in the Lake Rudolf Basin. Chicago: University of Chicago Press, 193208.Google Scholar
Beden, M. (1979). Les éléphants (Loxodonta et Elephas) d’Afrique Orientale: systématique, phylogénie, intérêt biochronologique. Thèse, Université de Poitiers, France.Google Scholar
Beden, M. (1987). Les faunes Plio-Pléistocène de la basse vallée de ‘omo (Éthiopie) Tome 2, Les éléphantidés (Mammalia: Proboscidea). Cahiers de paléontologie. Paris: Editions du CNRS.Google Scholar
Begon, M., Townsend, C.R. and Harper, J.L. (2006). Ecology: From Individuals to Ecosystems, 4th ed. Oxford: Blackwell.Google Scholar
Begun, D.R. (2015). The Real Planet of the Apes: A New Story of Human Origins. Princeton: Princeton University Press.Google Scholar
Behrensmeyer, A.K. (1975). The taphonomy and paleoecology of Plio-Pleistocene vertebrate assemblages East of Lake Rudolf, Kenya. Bulletin of the Museum of Comparative Zoology 146, 473578.Google Scholar
Behrensmeyer, A.K. (1976). Lothagam Hill, Kanapoi, and Ekora: A general summary of stratigraphy and faunas. In: Coppens, Y., Howell, F.C., Isaac, G.L. and Leakey, R.E. (Eds.), Earliest Man and Environments in the Lake Rudolf Basin. Chicago: University of Chicago Press, pp. 163170.Google Scholar
Behrensmeyer, A.K. (1978a). The habitat of Plio-Pleistocene hominids in East Africa: taphonomic and microstratigraphic evidence. In:Jolly, C.J. (Ed.), Early Hominids of Africa. London: Duckworth, pp. 165189.Google Scholar
Behrensmeyer, A.K. (1978b). Taphonomic and ecologic information from bone weathering. Paleobiology 4, 150162.Google Scholar
Behrensmeyer, A.K. (1982). Time resolution in fluvial vertebrate assemblages. Paleobiology 8(3), 211227.CrossRefGoogle Scholar
Behrensmeyer, A.K. (1985). Taphonomy and the paleoecologic reconstruction of hominid habitats in the Koobi Fora Formation. In: Coppens, Y. (Ed.), L’Environnement des Hominidés au Plio-Pléistocène. Paris: Masson, pp. 309324.Google Scholar
Behrensmeyer, A.K. (1988). Vertebrate preservation in fluvial channels. Palaeogeography, Palaeoclimatolology, Palaeoecology 63(1–3), 183199.Google Scholar
Behrensmeyer, A.K. (1991). Terrestrial vertebrate accumulations. In: Allison, P. and Briggs, D.E.G. (Eds.), Taphonomy: Releasing the Data Locked in the Fossil Record. New York: Plenum, pp. 291335.Google Scholar
Behrensmeyer, A.K. (2006). Climate change and human evolution. Science 311, 476478.Google Scholar
Behrensmeyer, A.K. and Barry, J.C. (2005). Biostratigraphic surveys in the Siwaliks of Pakistan: a method for standardized surface sampling of the vertebrate fossil record. Palaeontologia Electronica 8, 124.Google Scholar
Behrensmeyer, A.K. and Boaz, D.E.D. (1980). The recent bones of Amboseli Park, Kenya in relation to East African paleoecology. In: Behrensmeyer, A.K. and Hill, A. (Eds.), Fossils in the Making: Vertebrate Taphonomy and Paleoecology. Chicago: University of Chicago Press, pp. 7293.Google Scholar
Behrensmeyer, A.K. and Boaz, D.E.D. (1981). Late Pleistocene geology and paleontology of Amboseli National Park, Kenya. In: Coetzee, J.A. (Ed.), Palaeoecology of Africa and the Surrounding Islands Vol. 13. Rotterdam: Balkema, pp. 175188.Google Scholar
Behrensmeyer, A.K. and Cooke, H.B.S. (1985). Paleoenvironments, stratigraphy, and taphonomy in the African Pliocene and early Pleistocene. In: Delson, E. (Ed.), Ancestors: The Hard Evidence. New York: Alan R. Liss, pp. 6062.Google Scholar
Behrensmeyer, A.K. and Isaac, G.L. (1997). Geological context and paleoenvironments. In: Isaac, G.L. (Ed.), Koobi Fora Research Project Volume 5: Plio-Pleistocene Archaeology. Oxford: Clarendon Press, pp. 1270.Google Scholar
Behrensmeyer, A.K. and Kidwell, S.M. (1985). Taphonomy’s contributions to paleobiology. Paleobiology, 11(1), 105119.Google Scholar
Behrensmeyer, A.K. and Laporte, L.F. (1981). Footprints of a Pleistocene hominid in northern Kenya. Nature 289(5794), 167169.Google Scholar
Behrensmeyer, A.K. and Reed, K.E. (2013). Reconstructing the habitats of Australopithecus: Paleoenvironments, site taphonomy, and faunas. In: Reed, K., Fleagle, J. and Leakey, R. (Eds.), The Paleobiology of Australopithecus. New York: Springer, pp. 4160.Google Scholar
Behrensmeyer, A.K. and Western, D. (2009). Bone burial in land surface assemblages and its impact on the vertebrate fossil record. Journal of Vertebrate Paleontology 29(Supplement to #3), 61.Google Scholar
Behrensmeyer, A.K., Western, D. and Boaz, D.E.D. (1979). New perspectives in vertebrate paleoecology from a recent bone assemblage. Paleobiology 5, 1221.Google Scholar
Behrensmeyer, A.K., Damuth, J.D., DiMichele, W.A., et al. (1992). Terrestrial Ecosystems through Time: Evolutionary Paleoecology of Terrestrial Plants and Animals. Chicago: University of Chicago Press.Google Scholar
Behrensmeyer, A.K., Potts, R., Plummer, T., et al. (1995). The Pleistocene locality of Kanjera, Western Kenya: stratigraphy, chronology and paleoenvironments. Journal of Human Evolution, 29, 247274.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
Behrensmeyer, A.K., Kidwell, S.M. and Gastaldo, R.A. (2000). Taphonomy and paleobiology. Paleobiology 26(S4), 103147.Google Scholar
Behrensmeyer, A.K., Potts, R., Deino, A. and Ditchfield, P. (2002). Olorgesailie, Kenya: a million years in the life of a rift basin. In: Renaut, R.W. and Ashley, G.M. (Eds.), Sedimentation in Continental Rifts, SEPM Special Publication 73. Broken Arrow, OK: Society for Sedimentary Geology, pp. 97106.CrossRefGoogle Scholar
Behrensmeyer, A.K., Harmon, E.H. and Kimbel, W.H. (2003). Environmental context and taphonomy of the A.L. 333 Locality, Hadar, Ethiopia. PaleoAnthropology PAS 2003 abstracts, A2.Google Scholar
Behrensmeyer, A.K., Bobe, R. and Alemseged, Z. (2007a). Approaches to the analysis of faunal change during the East African Pliocene. In: Bobe, R., Alemseged, Z. and Behrensmeyer, A.K. (Eds.), Hominin Environments in the East African Pliocene: An Assessment of the Faunal Evidence. Berlin: Springer, 124.Google Scholar
Behrensmeyer, A.K., Alemseged, Z., Bobe, R. (2007b). Finale and future. Investigating faunal evidence for hominin paleoecology in East Africa. In: Bobe, R., Alemseged, Z. and Behrensmeyer, A.K. (Eds.), Hominin Environments in the East African Pliocene: An Assessment of the Faunal Evidence. Berlin: Springer Verlag, pp. 333345.Google Scholar
Behrensmeyer, A.K., Potts, R. and Deino, A. (2018). The Oltulelei Formation of the southern Kenyan Rift Valley: a chronicle of rapid landscape transformation over the last 500 ky. Bulletin of the Geological Society of America, 130(9–10), 14741492.Google Scholar
Belmaker, M. (2006). Large mammalian community structure through time: ‘Ubeidiya’ as a case study. PhD dissertation, The Hebrew University of Jerusalem, Jerusalem, Israel.Google Scholar
Belmaker, M. (2010). Early Pleistocene faunal connections between Africa and Eurasia: an ecological perspective. In: Fleagle, J.G., Shea, J.J., Grine, F.E., Baden, A.L. and Leakey, R.E. (Eds.), Out of Africa I. Dordrecht: Springer, pp. 183205.Google Scholar
Belmaker, M. (2018). Insight from carnivore community composition on the paleoecology of early Pleistocene Eurasian sites: implications for the dispersal of hominins out of Africa. Quaternary International 464(A), 317. DOI: 10.1016/j.quaint.2017.02.017Google Scholar
Belyea, L.R. and Lancaster, J. (1999). Assembly rules within a contingent ecology. Oikos 86, 402416.Google Scholar
Ben Moktar, N. and Mannaï-Tayech, B. (2015). Palynological reconstruction of climate and paleoenvironment during the lower to middle Miocene in northeast Tunisia. Arabian Journal of Geosciences 8, 1114911159.Google Scholar
Benammi, M., Calvo, M., Prévot, M. and Jaeger, J.-J. (1996). Magnetostratigraphy and paleontology of Aït Kandoula basin (High Atlas, Morocco) and the African-European late Miocene terrestrial fauna exchanges. Earth and Planetary Science Letters 145(1), 1529.Google Scholar
Bender, P.A. (1990). A reconsideration of the fossil Suidae of the Makapansgat Limeworks, Potgietersrus, Northern Transvaal. Unpublished MSc dissertation. University of the Witwatersrand, Johannesburg.Google Scholar
Bender, P.A. (1992). A reconsideration of the fossil Suid, Potamochoeroides shawi, from the Makapansgat Limeworks, Potgietersrus, Northern Transvaal. Navorsinge van die Nasionale Museum, Bloemfontein 8, 167.Google Scholar
Bender, P.A. and Brink, J.S. (1992). A preliminary report on new large mammal fossil finds from the Cornelia-Uitzoek site. South African Journal of Science 88, 512515.Google Scholar
Benefit, B.R. and McCrossin, M.L. (1990). Diet, species diversity, and distribution of African fossil baboons. Kroeber Anthropological Society Papers 71–72, 7993.Google Scholar
Benefit, B.R. and Pickford, M. (1986). Miocene fossil cercopithecoids from Kenya. American Journal of Physical Anthropology 69, 441464.Google Scholar
Benito-Calvo, A. and de la Torre, I. (2011). Analysis of orientation patterns in Olduvai Bed I assemblages using GIS techniques: Implications for site formation processes. Journal of Human Evolution 61, 5060.Google Scholar
Bennett, K.D. (1997). Evolution and Ecology: The Pace of Life. Cambridge: Cambridge University Press.Google Scholar
Bennett, M.R. and Reynolds, S.C. (2021). Inferences from footprints: archaeological best practice. In: Pastoors, A. and Lenssen-Erz, T. (Eds.), Reading Prehistoric Human Tracks. Cham: Springer, pp. 1539.Google Scholar
Bennett, M.R., Harris, J.W.K., Richmond, B.G., et al. (2009). Early hominin foot morphology based on 1.5-million-year-old footprints from Ileret, Kenya. Science 323, 11971201.Google Scholar
Bennington, J.B., Dimichele, W.A., Badgley, C., et al. (2009). Critical Issues of scale in paleoecology. Palaios 24(1), 14.Google Scholar
Benton, M.J. and Storrs, G.W. (1994). Testing the quality of the fossil record: paleontological knowledge is improving. Geology 22(2), 111114.Google Scholar
Berge, C., Penin, X. and Pellé, É. (2006). New interpretation of Laetoli footprints using an experimental approach and Procrustes analysis: preliminary results. Comptes Rendus Palevol 5, 561569.Google Scholar
Berger, L.R. (2006). Brief communication: Predatory bird damage to the Taung type-skull of Australopithecus africanus Dart 1925. American Journal of Physical Anthropology 131, 166168.Google Scholar
Berger, L. and Brink, J. (2007). An atlas of Southern African mammalian fossil bearing sites – Late Miocene to Late Pleistocene. Online at: www.profleeberger.com/files/An_Atlas_of_southern_African_Fossil_Bearing_Sites.pdf.Google Scholar
Berger, L.R. and Clarke, R.J. (1995). Eagle involvement in accumulation of the Taung child fauna. Journal of Human Evolution 29, 275299.Google Scholar
Berger, L.R. and Lacruz, R. (2003). Preliminary report on the first excavations at the new fossil site of Motsetse, Gauteng, South Africa. South African Journal of Science, 99, 279282.Google Scholar
Berger, L.R. and McGraw, W.S. (2007). Further evidence for eagle predation of, and feeding damage on, the Taung child. South African Journal of Science 103, 496498.Google Scholar
Berger, L.R. and Parkington, J.E. (1995). A new Pleistocene hominid‐bearing locality at Hoedjiespunt, South Africa. American Journal of Physical Anthropology 98(4), 601609.Google Scholar
Berger, L.R. and Tobias, P.V. (1994). New discoveries at the early hominid site of Gladysvale, South Africa. South African Journal of Science, 90(4), 223226.Google Scholar
Berger, L.R., Keyser, A.W. and Tobias, P.V. (1993). Gladysvale: first early hominid site discovered in South Africa since 1948. American Journal of Physical Anthropology 92(1), 107111.Google Scholar
Berger, L.R., Pickford, M. and Thackeray, J.F. (1995). A Plio-Pleistocene hominin upper central incisor from the Cooper’s site, South Africa. South African Journal of Science 91, 341342.Google Scholar
Berger, L.R., Lacruz, R. and de Ruiter, D.J. (2002). Brief communication: Revised age estimates of Australopithecus-bearing deposits at Sterkfontein, South Africa. American Journal of Physical Anthropology 119, 192197.CrossRefGoogle Scholar
Berger, L.R., de Ruiter, D.J., Steininger, C.M. and Hancox, P.J. (2003). Preliminary results of excavations at the newly investigated Coopers D deposit, Gauteng, South Africa. South African Journal of Science 99, 276278.Google Scholar
Berger, L.R., Lacruz, R., Hall, G., et al. (2006). An Acheulean handaxe from Gladysvale Cave Site, Gauteng, South Africa: research in action. South African Journal of Science 102(3), 103105.Google Scholar
Berger, L.R., Pickering, R., Kuhn, B., et al. (2009). A Mid-Pleistocene in situ fossil brown hyaena (Parahyaena brunnea) latrine from Gladysvale Cave, South Africa. Palaeogeography, Palaeoclimatology, Palaeoecology 279(3–4), 131136.Google Scholar
Berger, L.R., de Ruiter, D.J., Churchill, S.E., et al. (2010). Australopithecus sediba: a new species of Homo-like australopith from South Africa. Science 328, 195204.CrossRefGoogle ScholarPubMed
Berger, L.R., Hawks, J., de Ruiter, D.J., et al. (2015). Homo naledi, a new species of the genus Homo from the Dinaledi Chamber, South Africa. elife 4, e09560.Google Scholar
Berna, F., Goldberg, P., Horwitz, L.K., et al. (2012). Microstratigraphic evidence of in situ fire in the Acheulean strata of Wonderwerk Cave, Northern Cape province, South Africa. Proceedings of the National Academy of Sciences 109(20), E1215E1220.Google Scholar
Bernor, R.L. (2007). The latest Miocene hipparionine (Equidae) from Lemudong’o, Kenya. Kirtlandia 56, 148151.Google Scholar
Bernor, R.L. and Harris, J.M. (2003). Systematics and evolutionary biology of the late Miocene and early Pliocene hipparionine equids from Lothagam, Kenya. In: Leakey, M.G. and Harris, J.M. (Eds.), Lothagam: The Dawn of Humanity in Eastern Africa. New York: Columbia University Press, pp. 387438.Google Scholar
Bernor, R.L. and Kaiser, T.M. (2006). Systematics and paleoecology of the earliest Pliocene equid, Eurygnathohippus hooijeri n. sp. from Langebaanweg, South Africa. Mitteilungen aus dem Hamburgischen zoologischen Museum und Institut 103, 149186.Google Scholar
Bernor, R.L. and Sanders, W.J. (1990). Fossil Equidae from Plio-Pleistocene strata of the upper Semliki, Zaire. In:Boaz, N.T. (Ed.), Evolution of Environments and Hominidae in the African Western Rift Valley. Martinsville: Virginia Museum of Natural History, pp. 189196.Google Scholar
Bernor, R.L., Kaiser, T.M. and Nelson, S.V. (2004). The Oldest Ethiopian Hipparion (Equinae, Perissodactyl from Chorora: Systematics, Paleodiet and Paleoclimate. Courier-Forschungsinstitut Senckenberg, 246, 213226.Google Scholar
Bernor, R.L., Scott, R.S. and Haile-Selassie, Y. (2005). A contribution to the evolutionary history of Ethiopian hipparionine horses (Mammalia, Equidae): morphometric evidence from the postcranial skeleton. Geodiversitas 27, 133158.Google Scholar
Bernor, R., Rook, L. and Haile-Selassie, Y. (2009). Paleobiogeography. Ardipithecus kadabba: Late Miocene Evidence from the Middle Awash, Ethiopia. Berkeley: University of California Press, pp. 549563.Google Scholar
Bernor, R.L., Armour-Chelu, M., Gilbert, W.H., Kaiser, T.M. and Schulz, E. (2010). Equidae. In: Werdelin, L. and Sanders, W.J. (Eds.), Cenozoic Mammals of Africa. Berkeley: University of California Press, pp. 685721.Google Scholar
Bernor, R.L., Gilbert, H., Semprebon, G.M., Simpson, S. and Semaw, S. (2013). Eurygnathohippus woldegabrieli, sp. nov. (Perissodactyla, Mammalia), from the Middle Pliocene of Aramis, Ethiopia. Journal of Vertebrate Paleontology 33, 14721485.Google Scholar
Bernoussi, R. (1997). Contribution à l’étude paléontologique et observations archéozoologiques pour deux sites du Pléistocène moyen du Maroc atlantique, la grotte à Hominidés de la carrière Thomas I et la Grotte des Rhinocéros de la carrière Oulad Hamida 1 (Casablanca, Maroc). Thèse Univ. Bordeaux I, no. 1711, 265 pp.Google Scholar
Besse, J. and Courtillot, V. (2002). Apparent and true polar wander and the geometry of the geomagnetic field over the last 200 Myr. Journal of Geophysical Research: Solid Earth 107(B11), EPM-6.Google Scholar
Betzler, C. and Ring, U. (1995). Sedimentology of the Malawi Rift: facies and stratigraphy of the Chiwondo Beds, northern Malawi. Journal of Human Evolution 28(1), 2335.Google Scholar
Betzler, C., Eberli, G. P., Kroon, D., et al. (2016). The abrupt onset of the modern South Asian Monsoon winds. Scientific Reports 6, 29838.Google Scholar
Beyene, Y., Katoh, S., WoldeGabriel, G., et al. (2013). The characteristics and chronology of the earliest Acheulean at Konso, Ethiopia. Proceedings of the National Academy of Sciences of the USA, 110, 15841591.Google Scholar
Beyene, Y., Asfaw, B., Sano, K. and Suwa, G. (2015). The Konso-Gardula Research Project, Volume 2. Archaeological Collections: Background and the Early Acheulean Assemblages. Bulletin 48. Tokyo: The University Museum, The University of Tokyo, pp. 1103.Google Scholar
Biberson, P. (1961a). Le cadre paléogéographique de la préhistoire du Maroc atlantique. Publications du Service des Antiquités du Maroc 16, 1235.Google Scholar
Biberson, P. (1961b). Le paléolithique inférieur du Maroc atlantique. Publications du Service des Antiquités du Maroc 17, 1544.Google Scholar
Bibi, F. (2011a). Tragelaphus nakuae: evolutionary change, biochronology, and turnover in the African Plio‐Pleistocene. Zoological Journal of the Linnean Society London, 162, 699711.Google Scholar
Bibi, F. (2011b). Mio-Pliocene faunal exchanges and african biogeography: the record of fossil bovids. PLoS ONE, 6(2), e16688.Google Scholar
Bibi, F. and Kiessling, W. (2015). Continuous evolutionary change in Plio-Pleistocene mammals of eastern Africa. Proceedings of the National Academy of Sciences of the USA 112, 1062310628.Google Scholar
Bibi, F., Hill, A., Beech, M. and Yasin, W. (2008). A river fauna from the Arabian desert: late Miocene fossils from the United Arab Emirates. Journal of Vertebrate Paleontology 28(3), 53A53A.Google Scholar
Bibi, F., Souron, A., Bocherens, H., Uno, K. and Boisserie, J.-R.. (2013). Ecological change in the lower Omo Valley around 2.8 Ma. Biology Letters 9: 20120890.Google Scholar
Bibi, F., Rowan, J. and Reed, K. (2017). Late Pliocene Bovidae from Ledi-Geraru (Lower Awash Valley, Ethiopia) and their implications for Afar paleoecology. Journal of Vertebrate Paleontology, 37, e1337639.Google Scholar
Bibi, F., Pante, M., Souron, A., et al. (2018). Paleoecology of the Serengeti during the Oldowan–Acheulean transition at Olduvai Gorge, Tanzania: the mammal and fish evidence. Journal of Human Evolution 120, 4875.CrossRefGoogle ScholarPubMed
Bigazzi, G., Balestrieri, M.L., Norelli, P. and Oddone, M. (2004). Fission-track dating of a tephra layer in the Alat Formation of the Dandero Group (Danakil Depression, Eritrea). Rivista Italiana di Stratigrafia e Paleontologia 110(Supplement), 4549.Google Scholar
Binford, L.R. (1981). Bones, Ancient Men and Modern Myths. New York: New York Academic Press.Google Scholar
Binford, L.R. (1983). In Pursuit of the Past. New York: Thames and Hudson.Google Scholar
Binford, L.R. (1986). Comment on “Systematic butchery by Plio-Pleistocene hominids at Olduvai Gorge.” Current Anthropology 27, 444446.Google Scholar
Binford, L.R., Bunn, H.T. and Kroll, E.M. (1988a). Fact and fiction about the Zinjanthropus floor: data, arguments, and interpretations. Current Anthropology 29, 123135.Google Scholar
Binford, L.R., Mills, M. and Stone, N.M. (1988b). Hyena scavenging behavior and its implications for the interpretation of faunal assemblages from FLK 22 (the Zinj floor) at Olduvai Gorge. Journal of Anthropological Archaeology 7, 99135.Google Scholar
Binetti, K.M. (2011). Early Pliocene hominid paleoenvironments in the Tugen Hills, Kenya. PhD dissertation, Yale University.Google Scholar
Birchette, M. (1981). Postcranial remains of Cercopithecoides. American Journal of Physical Anthropology 54, 201.Google Scholar
Birkenfeld, M., Avery, D.M. and Horwitz, L.K. (2015). GIS virtual reconstructions of the temporal and spatial relations of fossil deposits at Wonderwerk Cave (South Africa). African Archaeological Review 32, 857876.Google Scholar
Bishop, L.C. (1994). Pigs and the ancestors: hominids, suids and environments during the Plio-Pleistocene of East Africa. PhD dissertation, Yale University. Ann Arbor, MI: University Microforms International. ProQuest publication number 9430258.Google Scholar
Bishop, L.C. (1997). Fossil suids from the Manonga Valley, Tanzania. In: Harrison, T. (Ed.), Neogene Paleontology of the Manonga Valley, Tanzania: A Window into the Evolutionary History of East Africa. New York: Springer, pp. 191217.Google Scholar
Bishop, L.C. (1999). Suid paleoecology and habitat preferences at African Pliocene and Pleistocene hominid localities. In: Bromage, T.G. and Schrenk, F. (Eds.), African Biogeography, Climate Change, and Human Evolution. Oxford: Oxford University Press, pp. 216225.Google Scholar
Bishop, L.C. (2010). Suoidea. In: Werdelin, L. and Sanders, W.J. (Eds.), Cenozoic Mammals of Africa. Berkeley: University of California Press, pp. 821842.Google Scholar
Bishop, L.C. (2011). Suidae. In: Harrison, T. (Ed.), Paleontology and Geology of Laetoli: Human Evolution in Context. Vol. 2: Fossil Hominins and the Associated Fauna. Dordrecht: Springer, pp. 327337.Google Scholar
Bishop, L. and Hill, A. (1999). Fossil Suidae from the Baynunah Formation, Emirate of Abu Dhabi, United Arab Emirates. In: Whybrow, P.J. and Hill, A.P. (Eds.), Fossil Vertebrates of Arabia, with Emphasis on the Late Miocene Faunas, Geology, and Palaeoenvironments of the Emirate of Abu Dhabi, United Arab Emirates. New Haven: Yale University Press, pp. 254270Google Scholar
Bishop, L.C. and Reynolds, S.C. (2000). Fauna from Twin Rivers. In: Barham, L. (Ed.), The Middle Stone Age of Zambia, South Central Africa. Bristol: Western Academic & Specialist Press, pp. 217222.Google Scholar
Bishop, L.C., Hill, A. and Kingston, J.D. (1999a). Palaeoecology of Suidae from the Tugen Hills, Baringo, Kenya. In: Andrews, P. and Banham, P. (Eds.), Late Cenozoic Environments and Hominid Evolution: A tribute to Bill Bishop. London: Geological Society of London, pp. 99111.Google Scholar
Bishop, L.C., Pickering, T.R., Plummer, T. and Thackeray, J.F. (1999b). Paleoenvironment setting for the Oldowan Industry at Sterkfontein (abstract). Paper for presentation at the XV International Union for Quaternary Research, 3–11 August 1999, Durban, South Africa.Google Scholar
Bishop, L.C., King, T., Hill, A. and Wood, B. (2006a). Palaeoecology of Kolpochoerus heseloni (= K. limnetes): a multiproxy approach. Proceedings of the Royal Society of South Africa 61, 8188.Google Scholar
Bishop, L.C., Plummer, T.W., Ferraro, J.V., et al. (2006b). Recent research into oldowan hominin activities at Kanjera South, Western Kenya. African Archaeological Review 23, 31.Google Scholar
Bishop, L.C., Plummer, T.W., Hertel, F. and Kovarovic, K. (2011). Paleoenvironments of Laetoli, Tanzania as determined by antelope habitat preferences. In: Harrison, T. (Ed.), Paleontology and Geology of Laetoli: Human Evolution in Context. Vol. 1: Geology, Geochronology, Paleoecology, and Paleoenvironment. Dordrecht: Springer, pp. 355366.Google Scholar
Bishop, W.W. (1968). The evolution of fossil environments in East Africa. Transactions of the Leicester Literary and Philosophical Society 62, 2244.Google Scholar
Bishop, W.W. (1971). The late Cenozoic history of East Afrca in relation to hominoid evolution. In: Turekian, K.K. (Ed.), The Late Cenozoic Glacial Ages. New Haven: Yale University Press, pp. 493527.Google Scholar
Bishop, W.W. and Chapman, G.R. (1970). Early Pliocene sediments and fossils from the northern Kenya Rift Valley. Nature 226, 914918.Google Scholar
Bishop, W.W. and Pickford, M.H.L. (1975). Geology, fauna and palaeoenvironments of the Ngorora Formation, Kenya Rift Valley. Nature 254, 185192.Google Scholar
Bishop, W.W. and Whyte, F. (1962). Tertiary mammalian faunas and sediments in Karamoja and Kavirondo, East Africa. Nature 4861, 12831287.Google Scholar
Bishop, W.W., Chapman, G.R., Hill, A. and Miller, J.A. (1971). Succession of Cainozoic vertebrate assemblages from Northern Kenya Rift Valley. Nature 233(5319), 389394.Google Scholar
Bishop, W., Hill, A. and Pickford, M. (1978). Chesowanja: a revised geological interpretation. In: Bishop, W.W. (Ed.), Geological Background to Fossil Man. Edinburgh: Scottish Academic Press, pp. 309327.Google Scholar
Black, C.C. and Krishtalka, L. (1986). Rodents, bats, and insectivores from the Plio-Pleistocene sediments to the east of Lake Turkana, Kenya. Contributions in Science 372, 115.Google Scholar
Blois, J.L. and Hadly, E.A. (2009). Mammalian responses to Cenozoic climatic change. Annual Review of Earth and Planetary Sciences 37, 181208.Google Scholar
Blondel, C., Merceron, G., Andossa, L., et al. (2010). Dental mesowear analysis of the late Miocene Bovidae from Toros-Menalla (Chad) and early hominid habitats in Central Africa. Palaeogeography, Palaeoclimatology, Palaeoecology 292, 184191.Google Scholar
Blondel, C., Rowan, J., Merceron, G., et al. (2018). Feeding ecology of Tragelaphini (Bovidae) from the Shungura Formation, Omo Valley, Ethiopia: contribution of dental wear analyses. Palaeogeography, Palaeoclimatology, Palaeoecology 496, 103120.Google Scholar
Blome, M.W., Cohen, A.S., Tryon, C.A., Brooks, A.S. and Russell, J. (2012). The environmental context for the origins of modern human diversity: a synthesis of regional variability in African climate 150,000–30,000 years ago. Journal of Human Evolution 62, 563592.Google Scholar
Blum, M.G.B. and Jakobsson, M. (2010). Deep divergences of human gene trees and models of human origins. Molecular Biology and Evolution 28, 889898.Google Scholar
Blumenschine, R.J. (1986). Early Hominid Scavenging Opportunities: Implications of Carcass Availability in the Serengeti and Ngorongoro Ecosystems. British Archaeological Reports International Series 283.Google Scholar
Blumenschine, R.J. (1987). Characteristics of an early hominid scavenging niche. Current Anthropology 29, 382407.Google Scholar
Blumenschine, R.J. (1988). An experimental model of the timing of hominid and carnivore influence on archaeological bone assemblages. Journal of Archaeological Science 15, 483502.Google Scholar
Blumenschine, R.J. (1995). Percussion marks, tooth marks, and experimental determinations of the timing of hominid and carnivore access to long bones at FLK Zinjanthropus, Olduvai Gorge, Tanzania. Journal of Human Evolution 29, 2151.Google Scholar
Blumenschine, R.J. and Masao, F.T. (1991). Living sites at Olduvai Gorge, Tanzania? Preliminary landscape archaeology results in the basal Bed II lake margin zone. Journal of Human Evolution 21, 451462.Google Scholar
Blumenschine, R.J. and Peters, C.R. (1998). Archaeological predictions for hominid land use in the paleo-Olduvai Basin, Tanzania, during lowermost Bed II times. Journal of Human Evolution 34, 565607.Google Scholar
Blumenschine, R.J. and Pobiner, B.L. (2007). Zooarchaeology and the ecology of Oldowan hominin carnivory. In: Ungar, P.S. (Ed.), Evolution of the Human Diet. New York: Oxford University Press, pp. 167190.Google Scholar
Blumenschine, R.J., Marean, C.W. and Capaldo, S.D. (1996). Blind tests of inter-analyst correspondence and accuracy in the identification of cut marks, percussion marks, and carnivore tooth marks on bone surfaces. Journal of Archaeological Science 23, 493507.Google Scholar
Blumenschine, R.J., Peters, C.R., Masao, F.T., et al. (2003). Late Pliocene Homo and hominid land use from Western Olduvai Gorge, Tanzania. Science 299(5610), 12171221.Google Scholar
Blumenschine, R.J., Stanistreet, I.G. and Masao, F.T. (2012a). Olduvai Gorge and the Olduvai Landscape Paleoanthropology Project. Journal of Human Evolution 63, 247250.Google Scholar
Blumenschine, R.J., Stanistreet, I.G., Njau, J.K., et al. (2012b). Environments and hominin activities across the FLK Peninsula during Zinjanthropus times (1.84 Ma), Olduvai Gorge, Tanzania. Journal of Human Evolution 63, 364383.Google Scholar
Blumenthal, S.A., Levin, N.E., Brown, F.H., et al. (2017). Aridity and hominin environments. Proceedings of the National Academy of Sciences 114, 73317336.Google Scholar
Boaz, N.T. (1990). Evolution of Environments and Hominidae in the African Western Rift Valley. Memoir Number 1. Martinsville: Virginia Museum of Natural History, p. 356.Google Scholar
Boaz, N.T., El-Arnauti, A., Gaziry, A.W., de Heinzelin, J. and Boaz, D.D. (Eds.) (1987). Neogene Paleontology and Geology of As Sahabi. New York: Alan R. Liss.Google Scholar
Boaz, N.T., El-Arnauti, A., Pavlakis, P., Salem, M.J. (Eds.) (2008a). Circum-Mediterranean Geology and Biotic Evolution During the Neogene Period: The Perspective from Libya. Garyounis Scientific Bulletin, Special Issue 5. Benghazi: University of Garyounis. 308 pp.Google Scholar
Boaz, N., El-Arnauti, A., Agustí, J., et al. (2008b). Temporal, lithostratigraphic, and biochronologic setting of the Sahabi Formation, North-Central Libya. Geology of East Libya 3, 959972.Google Scholar
Bobe, R. (1997). Hominid environments in the Pliocene: an analysis of fossil mammals from the lower Omo valley, Ethiopia. PhD dissertation, Anthropology, University of Washington.Google Scholar
Bobe, R. (2006). The evolution of arid ecosystems in eastern Africa. Journal of Arid Environments 66(3), 564584.Google Scholar
Bobe, R. (2011). Fossil mammals and paleoenvironments in the Omo-Turkana Basin. Evolutionary Anthropology 20, 254263.Google Scholar
Bobe, R. and Behrensmeyer, A.K. (2004). The expansion of grassland ecosystems in Africa in relation to mammalian evolution and the origin of the genus Homo. Palaeogeography, Palaeoclimatology, Palaeoecology 207, 399420.Google Scholar
Bobe, R. and Carvalho, S. (2019). Hominin diversity and high environmental variability in the Okote Member, Koobi Fora Formation, Kenya. Journal of Human Evolution 126, 91105.Google Scholar
Bobe, R. and Eck, G.G. (2001). Responses of African bovids to Pliocene climatic change. Paleobiology 27, 147.Google Scholar
Bobe, R. and Leakey, M.G. (2009). Ecology of Plio-Pleistocene mammals in the Omo – Turkana Basin and the emergence of Homo. In: Grine, F.E., Fleagle, J.G. and Leakey, R.E. (Eds.), The First Humans – Origin and Early Evolution of the Genus Homo. Dordrecht, Springer, pp. 173184.Google Scholar
Bobe, R., Behrensmeyer, A.K. and Chapman, R.E. (2002). Faunal change, environmental variability, and late Pliocene hominin evolution. Journal of Human Evolution 42, 475497.Google Scholar
Bobe, R., Alemseged, Z. and Behrensmeyer, A.K. (Eds.) (2007a). Hominin Environments in the East African Pliocene: An Assessment of the Faunal Evidences. Dordrecht: Springer.Google Scholar
Bobe, R., Behrensmeyer, A.K., Eck, G. and Harris, J.M. (2007b). Patterns of abundance and diversity in late Cenozoic bovids from the Turkana and Hadar Basins, Kenya and Ethiopia. In: Bobe, R., Alemseged, Z. and Behrensmeyer, A.K. (Eds.), Hominin Environments in the East African Pliocene: An Assessment of the Faunal Evidence. Dordrecht: Springer, pp. 129157.Google Scholar
Bobe, R., Assefa, Z., and Lepre, C.J. (2009). The Akaki-Fanta prehistoric site, Addis Ababa, Ethiopia. Report to the National Geographic Society. Washington, DC: National Geographic Society, p. 12.Google Scholar
Bobe, R., Manthi, F.K., Ward, C.V., Plavcan, J.M., and Carvalho, S. (2020a). The ecology of Australopithecus anamensis in the early Pliocene of Kanapoi, Kenya. Journal of Human Evolution 140, 102717.Google Scholar
Bobe, R., Martínez, F.I. and Carvalho, S. (2020b). Primate adaptations and evolution in the southern African Rift Valley. Evolutionary Anthropology 29, 94101.Google Scholar
Bocherens, H., Koch, P., Mariotti, A., Geraads, D. and Jaeger, J.-J. (1996). Isotopic biogeochemistry (13C, 18O) of mammalian enamel from African Pleistocene Hominid sites. Palaios 11, 306318.Google Scholar
Boisserie, J.R. (2005). The phylogeny and taxonomy of Hippopotamidae (Mammalia: Artiodactyla): a review based on morphology and cladistic analysis. Zoological Journal of the Linnean Society 143(1), 126.Google Scholar
Boisserie, J.-R. (2020). Hippopotamidae (Cetartiodactyla, Hippopotamoidea) from Kanapoi, Kenya, and the taxonomic status of the late early Pliocene hippopotamids from the Turkana Basin. Journal of Human Evolution 140, 102377.Google Scholar
Boisserie, J.R. and Merceron, G. (2011). Correlating the success of Hippopotaminae with the C4 grass expansion in Africa: relationship and diet of early Pliocene hippopotamids from Langebaanweg, South Africa. Palaeogeography, Palaeoclimatology, Palaeoecology 308(3–4), 350361.Google Scholar
Boisserie, J.-R. and White, T.D. (2004). A new species of Pliocene Hippopotamidae from the Middle Awash, Ethiopia. Journal of Vertebrate Paleontology 24, 464473.Google Scholar
Boisserie, J.-R., Brunet, M., Andossa, L. and Vignaud, P. (2003). Hippopotamids from the Djurab Pliocene faunas, Chad, Central Africa. Journal of African Earth Sciences 36, 1527.Google Scholar
Boisserie, J.R., Zazzo, A., Merceron, G., et al. (2005a). Diets of modern and late Miocene hippopotamids: evidence from carbon isotope composition and micro-wear of tooth enamel. Palaeogeography, Palaeoclimatology, Palaeoecology, 221(1–2), 153174. doi:10.1016/j.palaeo.2005.02.010Google Scholar
Boisserie, J.-R., Likius, A., Vignaud, P. and Brunet, M. (2005b) A new late Miocene hippopotamid from Toros-Menalla, Chad. Journal of Vertebrate Paleontology 25(3), 665673.Google Scholar
Boisserie, J.R., Guy, F., Delagnes, A., et al. (2008). New palaeoanthropological research in the Plio-Pleistocene Omo Group, Lower Omo Valley, SNNPR (Southern Nations, Nationalities and People Regions), Ethiopia. Comptes Rendus Palevol 7(7), 429439.Google Scholar
Boisserie, J.-R., Souron, A., Mackaye, H.T., et al. (2014). A new species of Nyanzachoerus (Cetartiodactyla: Suidae) from the Late Miocene Toros-Ménalla, Chad, Central Africa. PLoS ONE 9, e103221.Google Scholar
Boisserie, J.-R., Suwa, G., Asfaw, B., et al. (2017). Basal hippopotamines from the upper Miocene of Chorora, Ethiopia. Journal of Vertebrate Paleontology 37, e1297718.Google Scholar
Bolter, D.R., Elliott, M.C., Hawks, J. and Berger, L.R. (2020). Immature remains and the first partial skeleton of a juvenile Homo naledi, a late Middle Pleistocene hominin from South Africa. PLoS ONE 15(4), e0230440.Google Scholar
Bondioli, L., Coppa, A., Frayer, D.W., et al. (2006). A one million year old human pubic symphysis. Journal of Human Evolution 50, 479483.Google Scholar
Boné, E.L. and Dart, R.A. (1955). A catalog of the Australopithecine fossils found at the Limeworks, Makapansgat. American Journal of Physical Anthropology 13, 621624.Google Scholar
Bonnefille, R. (1968). Contribution à l’étude de la flore d’un niveau pléistocène de la haute vallée de l’Aouache (Ethiopie). Comptes Rendu de l’Académie des Sciences Paris 266, 12291232.Google Scholar
Bonnefille, R. (1969). Indication sur la paléoflore d’un niveau du Quaternaire moyen du site de Melka Kontouré (Ethiopie). Comptes rendu sommaire des Séances de la Société géologique de France 7, 238239.Google Scholar
Bonnefille, R. (1971). Atlas des pollens d’Ethiopie. Principales espèces des forêts de montagne. Pollens et spores 13, 1572.Google Scholar
Bonnefille, R. (1972). Associations polliniques actuelles et quaternaires en Ethiopie (Vallées de l’Awash et de l’Omo). Thèse de Doctorat ès Sciences, Université Paris VI, France.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. (1984a). Palynological research at Olduvai Gorge. National Geographic Society Research Papers 17, 227243.Google Scholar
Bonnefille, R. (1984b). Cenozoic vegetation and environments of early hominidsin East Africa. In: White, R.O. (Ed.), The Evolution of the East Asian Environment. Hong Kong: University of Hong Kong, pp.579612.Google Scholar
Bonnefille, R. (2010). Cenozoic vegetation, climate changes and hominid evolution in tropical Africa. Global and Planetary Change 72, 390411.Google Scholar
Bonnefille, R. and Dechamps, R. (1983). Data on fossil flora. In: de Heinzelin, J. (Ed.), The Omo Group. Tervuren: Musée Royale de l’Afrique Central, pp. 191207.Google Scholar
Bonnefille, R. and Letouzey, R. (1976). Fruits fossiles d’Antrocaryon dans la vallée de l’Omo (Éthiopie). Adansonia, Ser. 2, 16, 6582.Google Scholar
Bonnefille, R. and Riollet, G. (1980). Palynologie, vegetation et climats de Bed I et Bed II à Olduvai, Tanzanie. In: Leakey, R.E. and Ogot, B.A. (Eds.), Proceedings of the Eighth Pan-African Congress of Prehistory and Quaternary Studies. Nairobi: The International Louis Leakey Memorial Institute for African Prehistory, pp. 123127.Google Scholar
Bonnefille, R. and Riollet, G. (1987). Palynological spectra from the Upper Laetolil Beds. In: Leakey, M.D. and Harris, J.M. (Eds.), Laetoli: A Pliocene Site in Northern Tanzania. Oxford: Clarendon Press, pp. 2347.Google Scholar
Bonnefille, R., Vincens, A. and Buchet, G. (1987). Palynology, stratigraphy, and paleoenvironments of a Pliocene hominid site (2.9–3.3 M.Y.) at Hadar, Ethiopia. Palaeogeography, Palaeoclimatology, Palaeoecology 60, 249281.Google Scholar
Bonnefille, R., Potts, R., Chalié, F., Jolly, D. and Peyron, O. (2004). High-resolution vegetation and climate change associated with Pliocene Australopithecus afarensis. Proceedings of the National Academy of Sciences, 101, 1212512129.Google Scholar
Bonnefille, R., Melis, R.T. and Mussi, M. (2018). Variability in the mountain environment at Melka Kunture archaeological site, Ethiopia, during the Early Pleistocene (~1.7 Ma) and the Middle Pleistocene transition (0.9–0.6 Ma). In: Gallotti, R. and Mussi, M. (Eds.), The Emergence of the Acheulean in East Africa and Beyond: Contributions in Honor of Jean Chavaillon. Cham: Springer International Publishing, pp. 93114.Google Scholar
Boshoff, A.F. and Kerley, G.I.H. (2001). Potential distributions of the medium- to large sized mammals in the Cape Floristic Region, based on historical accounts and habitat requirements. African Zoology 36, 245e273.Google Scholar
Boshoff, A., Landman, M. and Kerley, G. (2016). Filling the gaps on the maps: historical distribution patterns of some larger mammals in part of southern Africa. Transactions of the Royal Society of South Africa 71, 2387.Google Scholar
Boswell, P.G.H. (1935). Human remains from Kanam and Kanjera, Kenya Colony. Nature 135, 371371.Google Scholar
Bougariane, B., Zouhri, S., Ouchaou, B., Oujaa, A. and Boudad, L. (2010). Large mammals from the Upper Pleistocene at Tamaris I ‘Grotte des gazelles’ (Casablanca, Morocco): paleoecological and biochronological implications. Historical Biology 22, 295302.Google Scholar
Boulton, S.J., Van De Velde, J.H., and Grimes, S.T. (2019). Palaeoenvironmental and tectonic significance ofMiocene lacustrine and palustrine carbonates (Aït Kandoula Formation) in the Ouarzazate Foreland Basin, Morocco. Sedimentary Geology 383, 195215.Google Scholar
Bowen, B.E. and Von dra, C.F. (1973). Stratigraphical relationships of the Plio-Pleistocene deposits, East Rudolf, Kenya. Nature 242, 391393.Google Scholar
Brachert, T.C., Brügmann, G.B., Mertz, D.F., et al. (2010). Stable isotope variation in tooth enamel from Neogene hippopotamids: monitor of meso and global climate and rift dynamics on the Albertine Rift, Uganda. International Journal of Earth Sciences 99, 16631675.Google Scholar
Bradley, B.J. (2008). Reconstructing phylogenies and phenotypes: a molecular view of human evolution. Journal of Anatomy 212(4), 337353.Google Scholar
Braga, J. and Thackeray, J.F. (2003). Early Homo at Kromdraai B: probabilistic and morphological analysis of the lower dentition. Comptes Rendu Palevol 2, 269279.Google Scholar
Braga, J. and Thackeray, J.F. (2016). Kromdraai, a Birthplace of Paranthropus in the Cradle of Humankind. Stellenbosch: African Sun Media Metro, 113 pp.Google Scholar
Braga, J., Thackeray, J.F., Dumoncel, J., et al. (2013). A new partial temporal bone of a juvenile hominin from the site of Kromdraai B (South Africa). Journal of Human Evolution 65, 447456.Google Scholar
Braga, J., Thackeray, J.F., Bruxelles, L., Dumoncel, J. and Fourvel, J.-B. (2017). Stretching the time span of hominin evolution at Kromdraai (Gauteng, South Africa): recent discoveries. Comptes Rendu Palevol 16(1), 5870.Google Scholar
Brain, C.K. (1958). The Transvaal Ape-man-bearing Cave Deposits. Transvaal Museum Memoir No. 11. Pretoria: Transvaal Museum.Google Scholar
Brain, C.K. (1967). Hottentot food remains and their bearing on the interpretation of fossil bone assemblages. Scientific Papers of theNamib Desert Research Station 32, 111.Google Scholar
Brain, C.K. (1969a). The contribution of Namib Desert Hottentots to an understanding of australopithecine bone accumulations. Scientific Papers of the Namib Desert Research Station 39, 1322.Google Scholar
Brain, C.K. (1969b). Faunal remains from the Bushman Rock Shelter, Eastern Transvaal. South African Archaeological Bulletin 24, 5255.Google Scholar
Brain, C.K. (1970). New finds at the Swartkrans australopithecine site. Nature 225, 11121119.Google Scholar
Brain, C.K. (1972). An attempt to reconstruct the behaviour of australopithecines: the evidence for interpersonal violence. Zoologica Africana 7, 379401.Google Scholar
Brain, C.K. (1974). Some suggested procedures in the analysis of bone accumulations from southern African Quaternary sites. Annals of the Transvaal Museum 29, 18.Google Scholar
Brain, C.K. (1975a). An introduction to the South African australopithecine bone accumulations. In: Clason, A.T. (Ed.), Archaeological Studies. Amsterdam: North Holland, pp. 109119.Google Scholar
Brain, C.K. (1975b). An interpretation of the bone assemblage from the Kromdraai australopithecine site, South Africa. In: Tuttle, R.H. (Ed.), Palaeontology, Morphology and Palaeoecology. The Hague: Mouton, pp. 225243.Google Scholar
Brain, C.K. (1976). A re-interpretation of the Swartkrans site and its remains. South African Journal of Science 72, 141146.Google Scholar
Brain, C.K. (1978). Some aspects of the South African australopithecine sites and their bone accumulations. In: Jolly, C. (Ed.), African Hominidae of the Plio-Pleistocene. London: Duckworth, pp. 131161.Google Scholar
Brain, C.K. (1980). Some criteria for the recognition of bone-collecting agencies in African caves. In:Behrensmeyer, A.K. and Hill, A.P. (Eds.), Fossils in the Making. Chicago: University of Chicago Press, pp. 107130.Google Scholar
Brain, C.K. (1981). The Hunters or the Hunted? An Introduction to African Cave Taphonomy. Chicago: The University of Chicago Press.Google Scholar
Brain, C.K. (1993a). Structure and stratigraphy of the Swartkrans Cave in the light of the new excavations. In: Brain, C.K. (Ed.), Swartkrans: A Cave’s Chronicle of Early Man. Transvaal Museum Monograph No. 8. Pretoria: Transvaal Museum, pp.722.Google Scholar
Brain, C.K. (1993b). A taphonomic overview of the Swartkrans fossil assemblages. In: Brain, C.K. (Ed.), Swartkrans: A Cave’s Chronicle of Early Man. Transvaal Museum Monograph No. 8. Pretoria: Transvaal Museum, pp.257264.Google Scholar
Brain, C.K. (1993c). The occurrence of burnt bones at Swartkrans and their implications for the control of fire by early hominids. In: Brain, C.K. (Ed.), Swartkrans: A Cave’s Chronicle of Early Man. Transvaal Museum Monograph No. 8. Pretoria: Transvaal Museum, pp.229242.Google Scholar
Brain, C.K. (1994). Swartkrans: A Cave’s Chronicle of Early Man. Transvaal Museum Monograph. Pretoria: Transvaal Museum.Google Scholar
Brain, C.K. and Shipman, P. (1993). The Swartkrans bone tools. In: Brain, C.K. (Ed.), Swartkrans: A Cave’s Chronicle of Early Man. Transvaal Museum Monograph No. 8. Pretoria: Transvaal Museum, pp.195215.Google Scholar
Brain, C.K. and Sillen, A. (1988). Evidence from the Swartkrans cave for the earliest use of fire. Nature 336, 464466.Google Scholar
Brain, C.K., Lowe, C.V.R. and Dart, R.A. (1955). Kafuan stone artefacts in the post-australopithecine breccia at Makapansgat. Nature 175(4444), 1618.Google Scholar
Brain, C.K., Churcher, R.S., Clark, J.D., et al. (1988). New evidence of early hominids, their culture and environment from the Swartkrans Cave, South Africa. South African Journal of Science 84, 828835.Google Scholar
Bramble, D.M. and Lieberman, D.E. (2004). Endurance running and the evolution of Homo. Nature 432(7015), 345352.Google Scholar
Brandy, L.-D., Sabatier, M. and Jaeger, J.-J. (1980). Implications phylogénétiques et biogéographiques des dernières découvertes de Muridae en Afghanistan, au Pakistan et en Ethiopie. Geobios 13, 639643.Google Scholar
Brantingham, P.J. (2003). A neutral model of stone raw material procurement. American Antiquity 68(3), 487509.Google Scholar
Braun, D.R. (2006). The ecology of Oldowan technology: perspectives from Koobi Fora and Kanjera South. Doctoral dissertation, Rutgers University.Google Scholar
Braun, D. and Harris, J.W.K. (2009). Plio-Pleistocene technological variation: a view from the KBS Mb., Koobi Fora Formation. In: Schick, K. and Toth, N. (Eds.), The Cutting Edge: New Approaches to the Archaeology of Human Origins. Gosport, IN: The Stone Age Institute, pp. 1731.Google Scholar
Braun, D., Plummer, T.W., Ditchfield, P., et al. (2008). Oldowan behavior and raw material transport: perspectives from the Kanjera Formation. Journal of Archaeological Science 35, 23292345.Google Scholar
Braun, D.R., Plummer, T., Ferraro, J.V., Ditchfield, P. and Bishop, L.C. (2009). Raw material quality and Oldowan hominin toolstone preferences: evidence from Kanjera South, Kenya. Journal of Archaeological Science 36, 16051614.Google Scholar
Braun, D.R., Harris, J.W.K., Levin, N.E., et al. (2010). Early hominin diet included diverse terrestrial and aquatic animals 1.95 Ma in East Turkana, Kenya. Proceedings of the National Academy of Sciences, 107(22), 1000210007.Google Scholar
Braun, D.R., Levin, N.E., Stynder, D., et al. (2013). Mid-Pleistocene Hominin occupation at Elandsfontein, Western Cape, South Africa. Quaternary Science Reviews 82, 145166.Google Scholar
Braun, K., Bar-Matthews, M., Matthews, A., et al. (2019a). Late Pleistocene records of speleothem stable isotopic compositions from Pinnacle Point on the South African south coast. Quaternary Research, 91, 265288.Google Scholar
Braun, D.R., Aldeias, V., Archer, W., et al. (2019b). Earliest known Oldowan artifacts at >2.58 Ma from Ledi-Geraru, Ethiopia, highlight early technological diversity. Proceedings of the National Academy of Sciences 116, 11712.Google Scholar
Brickhill, J.A.J. (1976). Some fossil felids from Makapansgat Limeworks. Unpublished BSc dissertation, University of the Witwatersrand.Google Scholar
Brink, J.S. (1987). The archaeozoology of Florisbad, Orange Free State. Memoirs of the National Museum, Bloemfontein 24, 1151.Google Scholar
Brink, J.S. (1988). The taphonomy and palaeoecology of the Florisbad spring fauna. Palaeoecology of Africa 19, 169179.Google Scholar
Brink, J.S. (1994). An ass, Equus (Asinus) sp., from the late Quaternary mammalian assemblages of Florisbad and Vlakkraal, central southern Africa. South African Journal of Science 90, 497500.Google Scholar
Brink, J.S. (1999). Preliminary report on a caprine from the Cape mountains, South Africa. Archaeozoologia 10, 1126.Google Scholar
Brink, J.S. (2004). The taphonomy of an early/middle Pleistocene hyaena burrow at Cornelia-Uitzoek, South Africa. Revue de Palaéobiologie, 23, 731740.Google Scholar
Brink, J.S. (2005). The evolution of the black wildebeest, Connochaetes gnou, and modern large mammal faunas in central South Africa. PhD thesis, University of Stellenbosch, South Africa.Google Scholar
Brink, J.S. (2016). Faunal evidence for mid- and late Quaternary environmental change in southern Africa. In: Knight, J. and Grab, S.W. (Eds.), Quaternary Environmental Change in Southern Africa: Physical and Human Dimensions. Cambridge: Cambridge University Press.Google Scholar
Brink, J.S. and Deacon, H.J. (1982). A study of a Last Interglacial shell midden and bone accumulation at Herolds Bay, Cape Province, South Africa. Palaeoecology of Africa 15, 3139.Google Scholar
Brink, J.S. and Lee-Thorp, J.A. (1992). The feeding niche of an extinct springbok, Antidorcas bondi (Antelopini, Bovidae), and its paleoenvironmental meaning. South African Journal of Science 88, 227229.Google Scholar
Brink, J.S. and Rossouw, L. (2000). New trial excavations at the Cornelia-Uitzoek type locality. Navorsinge van die Nasionale Museum 16, 141156.Google Scholar
Brink, J.S., Herries, A.I.R., Moggi-Cecchi, J., et al. (2012). First hominine remains from a ~1.0 million year old bone bed at Cornelia-Uitzoek, Free State Province, South Africa. Journal of Human Evolution 63, 527535.Google Scholar
Brink, J.S., de Beer, F.C., Hoffman, J. and Bam, L. (2013). The evolutionary meaning of Raphicerus-like morphology in the dentitions and postcrania of Antidorcas bondi (Antilopini). Zitteliana 31, 21.Google Scholar
Brink, J.S., Bousman, C.B. and Grün, R. (2015a). A reconstruction of the skull of Megalotragus priscus (Broom, 1909), based on a find from Erfkroon, Modder River, South Africa, with notes on the chronology and biogeography of the species. Palaeoecology of Africa 33, 7194.Google Scholar
Brink, J., Holt, S. and Horwitz, L.K. (2015b). Preliminary findings on macro-faunal taxonomy, taphonomy, biochronology and palaeoecology from the basal layers of Wonderwerk Cave, South Africa. In: Thiaw, I. and Bocoum, H. (Eds.), Preserving African Cultural Heritage. (Proceedings of the 13th Congress of the Panafrican Archaeological Association for Prehistory and Related Studies – PAA and of the 20th Meeting of the Society of Africanist Archaeologists – Safa). Dakar: Memoires de IIFAN – C. A. DIOP, no. 93, pp. 137147.Google Scholar
Brink, J.S., Holt, S. and Horwitz, L.K. (2016). The Oldowan and early Acheulean mammalian fauna of Wonderwerk Cave (Northern Cape Province, South Africa). African Archaeological Review 33, 223250.Google Scholar
Broadley, D.G. (1962). Some fossil chelonian fragments from Makapansgat. Nature 194(4830), 791.Google Scholar
Broadley, D.G. (1997). A new species of Psammobates (Reptilia: Testudinidae) from the early Pleistocene of South Africa. Palaeontologica Africana 34, 111114.Google Scholar
Broccoli, A.J., Dahl, K.A. and Stouffer, R.J. (2006). Response of the ITCZ to Northern Hemisphere cooling. Geophysical Research Letters 33(1).Google Scholar
Brochu, C.A. (2020). Pliocene crocodiles from Kanapoi, Turkana Basin, Kenya. Journal of Human Evolution 140, 102410.Google Scholar
Brochu, C.A. and Storrs, G.W. (2012). A giant crocodile from the Plio-Pleistocene of Kenya, the phylogenetic relationships of Neogene African crocodylines, and the antiquity of Crocodylus in Africa. Journal of Vertebrate Paleontology 32, 587602.Google Scholar
Brochu, C.A., Njau, J., Blumenschine, R.J. and Densmore, Ll.D. (2010). New horned crocodile from the Plio-Pleistocene hominid sites at Olduvai Gorge, Tanzania. PLoS ONE 5(2), e9333. doi:10.1371/journal.pone.0009333Google Scholar
Brock, A., McFadden, P.L. and Partridge, T.C. (1977). Preliminary palaeomagnetic results from Makapansgat and Swartkrans. Nature 266(5599), 249250.Google Scholar
Brocklehurst, N., Upchurch, P., Mannion, P.D. and O’Connor, J. (2012). The completeness of the fossil record of Mesozoic birds: implications for early avian evolution. PLoS ONE 7(6), e30956.Google Scholar
Bromage, T.G. and Schrenk, F. (1987). A cercopithecoid tooth from the Pliocene in Malawi. Journal of Human Evolution 15, 497500.Google Scholar
Bromage, T.G., Schrenk, F. and Kaufulu, Z. (1985). Plio-Pleistocene deposits in Malawi’s hominid corridor. National Geographic Society Research Report 21, 5152.Google Scholar
Bromage, T.G., Schrenk, F. and Juwayeyi, Y.M. (1995). Paleobiogeography of the Malawi Rift: age and vertebrate paleontology of the Chiwondo Beds, northern Malawi. Journal of Human Evolution 28, 3757.Google Scholar
Brooks, A.C. (1961). A study of the Thomson’s gazelle (Gazella thomsonii Gunther) in Tanganyika. London: Colonial Research Publications (25).Google Scholar
Brooks, A.S. and Yellen, J.E. (1987). The preservation of activity areas in the archaeological record: ethnoarchaeological and archaeological work in northwest Ngamiland, Botswana. In: Kent, S. (Ed.), Method and Theory of Activity Area Research: An Ethnoarchaeologial Approach. New York: Columbia University Press, pp. 63106.Google Scholar
Brooks, A.S., Nevell, L., Yellen, J.E. and Hartman, G. (2006). Projectile technologies of the African MSA. In: Hovers, E. and Kuhn, S.L. (Eds.), Transitions Before the Transition: Evolution and Stability in the Middle Paleolithic and Middle Stone Age. New York: Springer, pp. 233256.Google Scholar
Brooks, A.S., Yellen, J.E., Potts, R., et al. (2018). Long-distance stone transport and pigment use in the earliest Middle Stone Age. Science 360, 9094.Google Scholar
Brooks, T., Balmford, A., Burgess, N., et al. (2001). Toward a blueprint for conservation in Africa. Bioscience 51, 613624.Google Scholar
Broom, R. (1913). Man contemporaenous with extinct animals in South Africa. Annals of the South African Museum XII, 1316.Google Scholar
Broom, R. (1925a). Some notes on the Taungs skull. Nature 115, 569.Google Scholar
Broom, R. (1925b). On the newly discovered South African man-ape. Natural History 25, 409418.Google Scholar
Broom, R. (1929). Note on the milk dentition of Australopithecus. Proceedings of the Zoological Society of London 99, 8588.Google Scholar
Broom, R. (1930). The age of Australopithecus. Nature 125, 814.Google Scholar
Broom, R. (1934). On the fossil remains associated with Australopithecus africanus. South African Journal of Science 31, 471480.Google Scholar
Broom, R. (1936). New fossil anthropoid skull from South Africa. Nature 138, 486488.Google Scholar
Broom, R. (1937). Notices of a few more new fossil mammals from the caves of the Transvaal. Annals of the Transvaal Museum 20(10), 509514.Google Scholar
Broom, R. (1938a). The Pleistocene anthropoid apes of South Africa. Nature 142, 377379.Google Scholar
Broom, R. (1938b). Further evidence on the structure of the South African Pleistocene anthropoids. Nature 142(3603), 897899.Google Scholar
Broom, R. (1939). The fossil rodents of the limestone cave at Taungs. Annals of the Transvaal Museum 19, 315317.Google Scholar
Broom, R. (1941). Mandible of a young Paranthropus child. Nature 147, 607608.Google Scholar
Broom, R. (1942). The hand of the ape-man, Paranthropus robustus. Nature 149, 513514.Google Scholar
Broom, R. (1943). An ankle-bone of the Ape-man, Paranthropus robustus. Nature 152, 689690.Google Scholar
Broom, R. (1943). South Africa’s part in the solution of the problem of the origin of man. South African Journal of Science 40, 6880.Google Scholar
Broom, R. (1945). Age of the South African ape-man. Nature 155, 389390.Google Scholar
Broom, R. (1946). The occurrence and general structure of the South African ape-men. Transvaal Museum Memoirs 2, 7153.Google Scholar
Broom, R. (1948a). Some South African Pliocene and Pleistocene mammals. Annals of the Transvaal Museum 21, 138.Google Scholar
Broom, R. (1948b). The giant rodent mole, Gypsorhychus. Annals of the Transvaal Museum 21, 4749.Google Scholar
Broom, R. (1949). Jaw of the ape-man Paranthropus crassidens. Nature 163, 903.Google Scholar
Broom, R. (1950). The genera and species of the South African fossil ape‐men. American Journal of Physical Anthropology 8(1), 114.Google Scholar
Broom, R. and Robinson, J.T. (1949). A new type of fossil man. Nature 164(4164), 322323.Google Scholar
Broom, R. and Robinson, J.T. (1952). Swartkrans Ape-Man, ‘Paranthropus crassidens. Transvaal Museum Memoire 2, Pretoria.Google Scholar
Broom, R. and Schepers, G.W.H. (1946). The South African Fossil Ape-Men: The Australopithecinae. Transvaal Museum Memoir No. 2. Pretoria: Transvaal Museum,Google Scholar
Broom, R., Robinson, J.T. and Schepers, G.W.H. (1950). Sterkfontein Ape-Man Plesianthropus (Transvaalensis). Transvaal Museum Memoir No. 4. Pretoria: Transvaal Museum.Google Scholar
Brophy, J.K. (2011). Reconstructing the habitat mosaic associated with Australopithecus robustus: evidence from quantitative morphological analysis of bovid teeth. Doctoral dissertation, Texas A&M University.Google Scholar
Brophy, J.K., de Ruiter, D.J., Athreya, S. and DeWitt, T.J. (2014). Quantitative morphological analysis of bovid teeth and implications for paleoenvironmental reconstruction of Plovers Lake, Gauteng Province, South Africa. Journal of Archaeological Science 41, 376388.Google Scholar
Brown, A. and Verhagen, B. (1985). Two Antidorcas bondi individuals from the Late Stone Age site of Kruger Cave 35/83, Olifantshoek, Rustenburg District, South Africa. South African Journal of Science 81, 102.Google Scholar
Brown, B., Brown, F.H. and Walker, A. (2001). New hominids from the Lake Turkana Basin, Kenya. Journal of Human Evolution 41, 2944.Google Scholar
Brown, F.H. (1969). Observations of the stratigraphy and radiometric age of the ‘Omo Beds’, lower Omo basin, southern Ethiopia. Quaternaria 11, 714.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. (1995). The potential of the Turkana Basin for paleoclimatic reconstruction in east Africa. In: Vrba, E.S., Denton, G.H., Partridge, T.C. and Burkle, L.H. (Ed.), Paleoclimate and Evolution with Emphasis on Human Origins. New Haven: Yale University Press, pp. 319330.Google Scholar
Brown, F.H. and Cerling, T.E. (1982). The stratigraphic significance of the Tulu Bor Tuff of the Koobi Fora Formation. Nature 299, 212215.CrossRefGoogle Scholar
Brown, F.H. and de Heinzelin, J. (1983). The lower Omo Basin. In: de Heinzelin, J. (Ed.), The Omo Group. Tervuren: Musee Royale de l’Afrique Centrale, pp. 724.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 of London 143, 297310.Google Scholar
Brown, F.H. and Feibel, C.S. (1991). Stratigraphy, depositional environments and palaeogeography of the Koobi Fora Formation. In: Harris, J.M. (Ed.), Koobi Fora Research Project, Volume 3: The Fossil Ungulates: Geology, Fossil Artiodactyls, and Paleoenvironments. Oxford: Clarendon Press, pp. 130.Google Scholar
Brown, F.H. and Fuller, C.R. (2008). Stratigraphy and tephra of the Kibish Formation, southwestern Ethiopia. Journal of Human Evolution 55, 366403.Google Scholar
Brown, F.H. and McDougall, I. (2011). Geochronology of the Turkana Depression of Northern Kenya and Southern Ethiopia. Evolutionary Anthropology 20, 217227.Google Scholar
Brown, F.H. and Nash, W.P. (1976). Radiometric dating and tuff mineralogy of Omo Group deposits. In: Coppens, Y., Howell, F.C., Isaac, G.L.I. and Leakey, R.E.F. (Eds.), Earliest Man and Environments in the Lake Rudolf Basin. Chicago: University of Chicago Press, pp. 5063.Google Scholar
Brown, F.H., de Heinzelin, J. and Howell, F.C. (1970). Plio/Pleistocene formations in the lower Omo Basin, southern Ethiopia. Quaternaria 8, 247268.Google Scholar
Brown, F.H., Shuey, R.T. and Croes, M.K. (1978). Magnetostratigraphy of the Shungura and Usno Formations, southwestern Ethiopia: new data and comprehensive reanalysis. Geophysical Journal International 54(3), 519538.Google Scholar
Brown, F.H., McDougall, I., Davies, T. and Maier, R. (1985). An integrated Plio-Pleistocene chronology for the Turkana Basin. In: Delson, E. (Ed.), Ancestors: The Hard Evidence. New York: Alan R. Liss, pp. 8290.Google Scholar
Brown, F.H., Haileab, B. and McDougall, I. (2006). Sequence of tuffs between the KBS Tuff and the Chari Tuff in the Turkana Basin, Kenya and Ethiopia. Journal of the Geological Society of London 163, 185204.Google Scholar
Brown, K.S., Marean, C.W., Herries, A.I.R., et al. (2009). Fire as an engineering tool of early modern humans. Science 325, 859862.Google Scholar
Bruch, A.A., Sievers, C. and Wadley, L. (2012). Quantification of climate and vegetation from southern African Middle Stone Age sites – an application using Late Pleistocene plant material from Sibudu, South Africa. Quaternary Science Reviews, 45, 717.Google Scholar
Brugal, J.P. (Ed.) (2017). Classes de taille chez les grands mammifères. In:Taphonomies. Paris: Edition des Archives Contemporaines, Coll. Sciences Archéologiques, pp. 472476.Google Scholar
Brugal, J.-P. and Denys, C. (1989). Vertébrés du site acheuléen d’Isenya (Kenya, district de Kajiado). Implications paléoécologiques et paléobiogéographiques. Comptes Rendu de l’Académie des Sciences Paris 308, 15031508.Google Scholar
Brugal, J.P., Roche, H. and Kibunjia, M. (2003). Faunes et paléoenvironnements des principaux sites archéologiques plio-pléistocènes de la formation de Nachukui (Ouest-Turkana, Kenya). Comptes Rendus Palevol 2(8), 675684.Google Scholar
Bruhn, R.L., Brown, F.H., Gathogo, P.N. and Haileab, B. (2011). Pliocene volcano-tectonics and paleogeography of the Turkana Basin, Kenya and Ethiopia. Journal of African Earth Sciences 59, 295312.Google Scholar
Brumfitt, I.M., Chinsamy, A. and Compton, J.S. (2013). Depositional environment and bone diagenesis of the Mio/Pliocene Langebaanweg bonebed, South Africa. South African Journal of Geology 116(2), 241258.Google Scholar
Brun, A., Guerin, C., Levy, A., Riser, J. and Rognon, P. (1988). Steppic environments at the end of the Upper Pleistocene in southern Tunisia (Oued el Akarit). Journal of African Earth Sciences 7, 969980.Google Scholar
Bruner, E., Bondioli, L., Coppa, A., et al. (2011a). A preliminary paleoneurological survey of the endocast from Buia (UA-31). PaleoAnthropology Journal 2011, 5 (abstract).Google Scholar
Bruner, E., de la Cuétara, J.M. and Holloway, R.L. (2011b). A bivariate approach to the variation of the parietal curvature in the genus Homo. Anatomic Record 294, 15481556.Google Scholar
Bruner, E., Bondioli, L., Coppa, A., et al. (2016). The endocast of the one-million-year-old human cranium from Buia (UA 31), Danakil Eritrea. American Journal of Physical Anthropology 160, 458468.Google Scholar
Brunet, M. (2010). Two new Mio-Pliocene Chadian hominids enlighten Charles Darwin’s 1871 prediction. Philosophical Transactions of the Royal Society B: Biological Sciences 365(1556), 33153321.Google Scholar
Brunet, M. and White, T.D. (2001). Deux nouvelles espèces de Suini (Mammalia, Suidae) du continent Africain (Ethiopie; Tchad). Comptes Rendus de l’Academie des Sciences, Paris 332, 5157.Google Scholar
Brunet, M., Beauvilain, A., Coppens, Y., et al. (1995). The first australopithecine 2,500 kilometres west of the Rift Valley (Chad). Nature 378(6554), 273275.Google Scholar
Brunet, M., Beauvilain, A., Coppens, Y., et al. (1996). Australopithecus bahrelghazali, une nouvelle espèce d’Hominidé ancien de la région de Koro Toro (Tchad). Comptes Rendus de l’Académie de Sciences 322, 907913.Google Scholar
Brunet, M., Beauvilain, A., Geraads, D., et al. (1998). Tchad: découverte d’une faune de mammifères du Pliocène inférieur. Comptes Rendus de l’Académie des Sciences – Series IIA – Earth and Planetary Science 326, 153158.Google Scholar
Brunet, M., Beauvilain, A., Billiou, D., et al. (2000). Chad: discovery of a vertebrate fauna close to the Mio-Pliocene boundary. Journal of Vertebrate Paleontology 20, 205209.Google Scholar
Brunet, M., Guy, F., Pilbeam, D., et al. (2002). A new hominid from the Upper Miocene of Chad, Central Africa. Nature 418(6894), 145151.Google Scholar
Brunet, M., Guy, F., Pilbeam, D., et al. (2005). New material of the earliest hominid from the Upper Miocene of Chad. Nature 434, 752755.Google Scholar
Bruxelles, L., Clarke, R.J., Maire, R., Ortega, R. and Stratford, D. (2014). Stratigraphic analysis of the Sterkfontein StW 573 Australopithecus skeleton and implications for its age. Journal of Human Evolution 70, 3648.Google Scholar
Bryant, J.D. and Froelich, P.N. (1995). A model of oxygen isotope fractionation in body water of large mammals. Geochimica et Cosmochimica Acta 59(21), 45234537.Google Scholar
Bulpin, T.V. (1980). Discovering South Africa (2nd ed.). Cape Town: T.V. Bulpin Publications (Pty) Ltd.Google Scholar
Bunn, H.T. (1981). Archaeological evidence for meat-eating by Plio-Pleistocene hominids from Koobi Fora and Olduvai Gorge. Nature 291, 574577.Google Scholar
Bunn, H. (1982). Meat-eating and human evolution: studies on the diet and subsistence patterns of Plio-Pleistocene hominids in East Africa. Doctoral Dissertation, University of California, Berkeley.Google Scholar
Bunn, H.T. (1997). The bone assemblages from the excavated sites. In: Isaac, G.L. (Ed.), Koobi Fora Research Project Volume 5: Plio-Pleistocene Archaeology. Oxford: Clarendon Press, pp. 402458.Google Scholar
Bunn, H.T. (2007). Meat made us human. In: Ungar, P.S. (Ed.), Evolution of the Human Diet. New York: Oxford University Press, pp.191211.Google Scholar
Bunn, H.T. and Kroll, E.M. (1986). Systematic butchery by Plio/Pleistocene hominids at Olduvai Gorge, Tanzania. Current Anthropology 27, 431451.Google Scholar
Bunn, H.T. and Kroll, E.M. (1988). Fact and fiction about the Zinjanthropus floor: data, arguments, and interpretations. Current Anthropology 29, 135149.Google Scholar
Bunn, H.T. and Pickering, T.R. (2010). Bovid mortality profiles in paleoecological context falsify hypotheses of endurance running–hunting and passive scavenging by early Pleistocene hominins. Quaternary Research 74(3), 395404.Google Scholar
Bush, A.M., Powell, M.G., Arnold, W.S., Bert, T.M. and Daley, G.M. (2002). Time-averaging, evolution, and morphologic variation. Paleobiology 28(1), 925.Google Scholar
Bush, A.M., Markey, M.J. and Marshall, C.R. (2004). Removing bias from diversity curves: the effects of spatially organized biodiversity on sampling-standardization. Paleobiology 30(4), 666686.Google Scholar
Butler, P. (1952). The milk molars of Perissodactyla, with remarks on molar occlusion. Proceedings of the Zoological Society of London 121, 777817.Google Scholar
Butler, P.M. and Greenwood, M. (1976). Elephant-shrews (Macroscelididae) from Olduvai and Makapansgat. In: Savage, R.J.G. and Coryndon, S.C. (Eds.), Fossil Vertebrates of Africa. London: Academic Press, pp. 155.Google Scholar
Butler, R.J., Benson, R.B.J., Carrano, M.T., Mannion, P.D. and Upchurch, P. (2011). Sea level, dinosaur diversity and sampling biases: investigating the ‘common cause’ hypothesis in the terrestrial realm. Proceedings of the Royal Society B: Biological Sciences 278(1709), 11651170.Google Scholar
Butzer, K.W. (1971). Another look at the australopithecine cave breccias of the Transvaal 1. American Anthropologist 73(5), 11971201.Google Scholar
Butzer, K.W. (1974). Geology of the Cornelia Beds. Memoir – Nasionale Museum, Bloemfontein 9, 732.Google Scholar
Butzer, K.W. (1976). The Mursi, Nkalabong, and Kibish formations, lower Omo Basin, Ethiopia. In: Coppens, Y., Howell, F.C., Isaac, G.L. and Leakey, R.E.F. (Eds.), Earliest Man and Environments in the Lake Rudolf Basin. Chicago: University of Chicago Press, pp. 2449.Google Scholar
Butzer, K.W. (1980). Palaeoecology of the South African Australopithecines. In Leakey, R. and Ogot, B. (Eds.), Proceedings of the Pan-African Congress of Prehistory and Quaternary Studies, Nairobi, 1977. Nairobi, pp. 131132.Google Scholar
Butzer, K.W. (1984a). Archaeology and Quaternary environment in the interior of southern Africa. In: Klein, R.G. (Ed.), Southern African Prehistory and Paleoenvironments. Rotterdam: Balkema, pp. 164.Google Scholar
Butzer, K.W. (1984b). Late Quaternary environments in South Africa. In: Vogel, J.C. (Ed.), Late Cenozoic Palaeoclimates of the Southern Hemisphere. Rotterdam: Balkema, pp. 235264.Google Scholar
Butzer, K.W. and Thurber, D.L. (1969). Some late Cenozoic sedimentary formations of the Lower Omo Basin. Nature 222, 11381143.Google Scholar
Butzer, K.W., Beaumont, P.B. and Vogel, J.C. (1978a). Lithostratigraphy of Border cave, KwaZulu, south Africa: a middle stone age sequence beginning c. 195,000 BP. Journal of Archaeological Science 5(4), 317341.Google Scholar
Butzer, K.W., Stuckenrath, R., Bruzewicz, A.J. and Helgren, D.M. (1978b). Late Cenozoic paleoclimates of the Gaap Escarpment, Kalahari margin, South Africa. Quaternary Research 10(3), 310339.Google Scholar
Butzer, K.W., Stuckenrath, R. and Vogel, J.C. (1979, April). The geo-archaeological sequence of Wonderwerk Cave, South Africa. In: Society of Africanist Archaeologists Meeting Calgary Abstracts.Google Scholar
Cabanes, D., Shahack-Gross, R. (2015). Understanding fossil phytolith preservation: the role of partial dissolution in paleoecology and archaeology. PLoS ONE 10, e0125532.Google Scholar
Cabido, M., Pons, E., Cantero, J.J., Lewis, J.P. and Anton, A. (2008). Photosynthetic pathway variation among C4 grasses along a precipitation gradient in Argentina. Journal of Biogeography 35, 131140.Google Scholar
Cadman, A. and Rayner, R.J. (1989). Climatic change and the appearance of Australopithecus africanus in the Makapansgat sediments. Journal of Human Evolution 18(2), 107113.Google Scholar
Calandra, I. and Merceron, G. (2016). Dental microwear texture analysis in mammalian ecology. Mammal Review 46(3), 215228.Google Scholar
Caley, T., Extier, T., Collins, J.A., et al. (2018). A two-million-year-long hydroclimatic context for hominin evolution in southeastern Africa. Nature, 560(7716), 7679.Google Scholar
Callaway, E. (2018). Femur findings remain a secret. Nature 553, 391392.Google Scholar
Calow, P.P. (Ed.). (1999). Blackwell’s Concise Encyclopedia of Ecology. Oxford: Blackwell.Google Scholar
Camp, C.L. (1948). University of California African expedition – southern section. Science 108, 550552.Google Scholar
Campbell, T.L., Lewis, P.J. and Williams, J.K. (2011). Analysis of the modern distribution of South African Gerbilliscus (Rodentia: Gerbillinae) with implications for Plio-Pleistocene palaeoenvironmental reconstruction. South African Journal of Science 107(1/2), Art. #497.Google Scholar
Campbell, M.C. and Tishkoff, S.A. (2010). The evolution of human genetic and phenotypic variation in Africa. Current Biology 20, R166R173.Google Scholar
Campisano, C.J. (2007). Tephrostratigraphy and hominin paleoenvironments of the Hadar Formation, Afar depression, Ethiopia. PhD dissertation, Rutgers University.Google Scholar
Campisano, C.J. (2012). Geological summary of the Busidima Formation (Plio-Pleistocene) at the Hadar paleoanthropological site, Afar Depression, Ethiopia. Journal of Human Evolution 62, 338352.Google Scholar
Campisano, C.J. and Feibel, C.S. (2007). Connecting local environmental sequences to global climate patterns: evidence from the hominin-bearing Hadar Formation, Ethiopia. Journal of Human Evolution 53(5), 515527.Google Scholar
Campisano, C.J. and Feibel, C.S. (2008). Depositional environments and stratigraphic summary of the Pliocene Hadar formation at Hadar, Afar depression, Ethiopia. In: Quade, J. and Wynn, J. (Eds.), The Geology of Early Humans in the Horn of Africa. Boulder: The Geological Society of America, pp.179201.Google Scholar
Campisano, C.J. and Reed, K.E. (2007). Spatial and temporal patterns of Australopithecus afarensis habitats at Hadar, Ethiopia. PaleoAnthropology PAS 2007 abstracts, A6.Google Scholar
Campisano, C. J., Cohen, A. S., Arrowsmith, J. R., et al. (2017). The Hominin Sites and Paleolakes Drilling Project: high-resolution paleoclimate records from the East African Rift System and their implications for understanding the environmental context of hominin evolution. Paleoanthropology 143.Google Scholar
Cande, S.C. and Kent, D.V. (1995). Revised calibration of the geomagnetic polarity time scale for the Late Cretaceous and Cenozoic. Journal of Geophysical Research 100, 60936095.Google Scholar
Cane, M.A. and Molnar, P. (2001). Closing the Indonesian seaway as a precursor to east African aridification around 3–4 million years ago. Nature 411, 157162.Google Scholar
Cannon, M.D. (2001). Archaeofaunal relative abundance, sample size, and statistical methods. Journal of Archaeological Science 28(2), 185195.Google Scholar
Capaldo, S.D. (1997). Experimental determinations of carcass processing by Plio-Pleistocene hominids and carnivores at FLK 22 (Zinjanthropus), Olduvai Gorge, Tanzania. Journal of Human Evolution 33, 555597.Google Scholar
Capaldo, S.D. and Peters, C.R. (1995). Skeletal inventories from wildebeest drownings at lakes Masek and Ndutu in the Serengeti ecosystem. Journal of Archaeological Science 22, 385408.Google Scholar
Carney, J., Hill, A., Miller, J.A. and Walker, A. (1971). Late australopithecine from Baringo District, Kenya. Nature 230, 509514.Google Scholar
Carrión, J.S. and Scott, L. (1999). The challenge of pollen analysis in palaeoenvironmental studies of hominid beds: the record from Sterkfontein caves. Journal of Human Evolution 36(4), 401408.Google Scholar
Cartwright, C.R. (2013). Identifying the woody resources of Diepkloof Rock Shelter (South Africa) using scanning electron microscopy of the MSA wood charcoal assemblages. Journal of Archaeological Science 40(9), 34633474.Google Scholar
Cartwright, C. and Parkington, J. (1997). The wood charcoal assemblages from Elands Bay Cave, southwestern Cape: principles, procedures and preliminary interpretation. The South African Archaeological Bulletin 52, 5972.Google Scholar
Caruana, M.V. (2017). Lithic production strategies in the Oldowan assemblages from Sterkfontein Member 5 and Swartkrans Member 1, Gauteng Province, South Africa. Journal of African Archaeology 15(1), 119.Google Scholar
Castañeda, I.S., Mulitza, S., Schefuß, E., et al. (2009). Wet phases in the Sahara/Sahel region and human migration patterns in North Africa. Proceedings of the National Academy of Sciences of the USA 106, 2015920163.Google Scholar
Cerdeño, E. (1998). Diversity and evolutionary trends of the Family Rhinocerotidae (Perissodactyla). Palaeogeography, Palaeoclimatology, Palaeoecology, 141(1), 1334.Google Scholar
Cerling, T.E. (1992). Development of grasslands and savannas in East Africa during the Neogene. Palaeogeography, Palaeoclimatology, Palaeoecology (Global and Planetary Change Section) 97, 241247.Google Scholar
Cerling, T.E. (2003). Diets of east African bovidae based on stable isotope analysis. Journal of Mammalogy 84(2), 456470.Google Scholar
Cerling, T.E. and Hay, R.L. (1986). An isotopic study of paleosol carbonates from Olduvai Gorge. Quaternary Research 25, 6378.Google Scholar
Cerling, T.E. and Viehl, K. (2004). Seasonal diet changes of the forest hog (Hylochoerus meinertzhageni Thomas) based on the carbon isotopic composition of hair. African Journal of Ecology 42(2), 8892.Google Scholar
Cerling, T.E., Bowman, J.R. and O’Neil, J.R. (1988). An isotopic study of a fluvial–lacustrine sequence: the Plio-Pleistocene Koobi For a Formation, East Africa. Palaeogeography, Palaeoclimatology, Palaeoecology 63, 335356.Google Scholar
Cerling, T.E., Quade, J., Wang, Y. and Bowman, J.R. (1989). Carbon isotopes in soils and palaeosols as ecology and palaeoecology indicators. Nature, 341, 138139.Google Scholar
Cerling, T.E., Wang, Y. and Quade, J. (1993). Expansion of C4 ecosystems as an indicator of global ecological change in the Late Miocene. Nature 361, 344345.Google Scholar
Cerling, T.E., Harris, J.M., MacFadden, B.J., et al. (1997). Global vegetation change through the Miocene–Pliocene boundary. Nature 389, 153158.Google Scholar
Cerling, T.E., Harris, J.M. and Leakey, M.G. (1999). Browsing and grazing in elephants: the isotope record of modern and fossil proboscideans. Oecologia 120, 364374.Google Scholar
Cerling, T.E., Harris, J.M., Leakey, M.G. and Mudida, N. (2003a). Stable isotope ecology of Northern Kenya, with emphasis on the Turkana Basin, Kenya. In: Leakey, M.G. and Harris, J.M. (Eds.), Lothagam: the Dawn of Humanity in Africa. New York: Columbia University Press, pp. 583603.Google Scholar
Cerling, T., Harris, J. and Leakey, M. (2003b). Isotope paleoecology of the Nawata and Nachukui Formations at Lothagam, Turkana Basin, Kenya. In: Leakey, M.G. and Harris, J.M. (Eds.), Lothagam: the Dawn of Humanity in Africa. New York: Columbia University Press, pp. 605624.Google Scholar
Cerling, T.E., Harris, J.M. and Passey, B. H. (2003c). Diets of east African bovidae based on stable isotope analysis. Journal of Mammalogy 84, 456470.Google Scholar
Cerling, T., Harris, J.M. and Leakey, M. (2005). Environmentally driven dietary adaptations in African mammals. In: Baldwin, I.T., Caldwell, M.M., Heldmaier, G., et al. (Eds.), A History of Atmospheric CO2; and Its Effects on Plants, Animals, and Ecosystems. New York: Springer, Vol. 177, pp. 258272.Google Scholar
Cerling, T.E., Levin, N.E., Quade, J., et al. (2010). Comment on the paleoenvironment of Ardipithecus ramidus. Science 328, 1105-d.Google Scholar
Cerling, T.E., Mbua, E., Kirera, F.M., et al. (2011a). Diet of Paranthropus boisei in the early Pleistocene of East Africa. Proceedings of the National Academy of Sciences 108, 93379341.Google Scholar
Cerling, T.E., Wynn, J.G., Andanje, S.A., et al. (2011b). Woody cover and hominin environments in the past 6 million years. Nature 476(7358), 5156.Google Scholar
Cerling, T.E., Manthi, F.K., Mbua, E.N., et al. (2013a). Stable isotope-based diet reconstructions of Turkana Basin hominins. Proceedings of the National Academy of Sciences 110(26), 1050110506.Google Scholar
Cerling, T.E., Chritz, K.L., Jablonski, N.G., Leakey, M.G. and Manthi, F.K. (2013b). Diet of Theropithecus from 4 to 1 Ma in Kenya. Proceedings of the National Academy of Sciences of the United States of America 110, 1050710512.Google Scholar
Cerling, T.E., Brown, F.H. and Wynn, J.G. (2014). On the environment of Aramis. Current Anthropology 55, 469470.Google Scholar
Cerling, T.E., Andanje, S.A., Blumenthal, S.A., et al. (2015a). Dietary changes of large herbivores in the Turkana Basin, Kenya from 4 to 1 Ma. Proceedings of the National Academy of Sciences 112, 1146711472.Google Scholar
Cerling, T.E., Brown, F.H. and Wynn, J.G. (2015b). On the environment of Aramis: concerning comment and replies of August 2014. Current Anthropology 56, 445446.Google Scholar
Chaïd-Saoudi, Y., Geraads, D. and Raynal, J.-P. (2006). The fauna and associated artefacts from the Lower Pleistocene site of Mansourah (Constantine, Algeria). Comptes-Rendus Palevol 5, 963971.Google Scholar
Chao, A. and Jost, L. (2012). Coverage‐based rarefaction and extrapolation: standardizing samples by completeness rather than size. Ecology 93(12), 25332547.Google Scholar
Chao, A., Gotelli, N.J., Hsieh, T.C., et al. (2014). Rarefaction and extrapolation with Hill numbers: a framework for sampling and estimation in species diversity studies. Ecological Monographs 84(1), 4567.Google Scholar
Chapman, C.A., Gautier-Hion, A., Oates, J.F. and Onderdonk, P.A. (1999). African primate communities: determinants of structure and threats to survival. In: Fleagle, J.G., Janson, C.H. and Reed, K.E. (Eds.), Primate Communities. Cambridge: Cambridge University Press, pp. 137.Google Scholar
Chapman, G.R. (1971). The geological evolution of the Northern Kamasia Hills Baringo District, Kenya. PhD dissertation, University of London.Google Scholar
Chapman, G.R. and Brook, M. (1978). Chronostratigraphy of the Baringo Basin, Kenya. In: Bishop, W.W. (Ed.), Geological Background to Fossil Man. Edinburgh: Scottish Academic Press, pp. 207223.Google Scholar
Chapman, G.R., Lippard, S.J. and Martyn, J.E. (1978). The stratigraphy and structure of the Kamasia Range, Kenya Rift Valley. Journal of the Geologicical Society of London 135, 265281.Google Scholar
Chapon, C., Bahain, J.-J., de Lumley, H., Perrenoud, C. and Van Damme, D. (2005). Stratigraphie, sédimentologie et magnétostratigraphie du site oldowayen de Fejej FJ-1, sud-Omo, Ethiopie: premiers résultats. Quaternaire 16, 143152.Google Scholar
Chapon, C., Bahain, J.-J., Beyene, Y., et al. (2011). Geochemistry of the Fejej tuffs (South Omo, Ethiopia), their tephrostratigraphical correlation with Plio-Pleistocene formations in the Omo-Turkana Basin. Comptes Rendus Palevol 10, 251258.Google Scholar
Charles-Dominique, T., Davies, T.J., Hempson, G.P., et al. (2016). Spiny plants, mammal browsers, and the origin of African savannas. Proceedings of the National Academy of Sciences 113(38), E5572E5579.Google Scholar
Charteris, J., Wall, J.C. and Nottrodt, J.W. (1981). Functional reconstruction of gait from the Pliocene hominid footprints at Laetoli, northern Tanzania. Nature 290, 496498.Google Scholar
Charteris, J., Wall, J.C. and Nottrodt, J.W. (1982). Pliocene hominid gait: new interpretations based on available footprint data from Laetoli. American Journal of Physical Anthropology 58, 133144.Google Scholar
Chase, B.M. and Meadows, M.E. (2007). Late Quaternary dynamics of southern Africa’s winter rainfall zone. Earth Science Reviews 84, 103138.Google Scholar
Chavaillon, J. (1973). Chronologie des niveaux paléolithiques de Melka-Kunturé. Comptes Rendu de l’Académie des Sciences (D) 276, 15331536.Google Scholar
Chavaillon, J. (1976). Evidence for the technical practices of early pleistocene hominids, Shungura Formation, Lower Omo Valley, Ethiopia. In:Coppens, Y., Howell, F.C., Isaac, G.L. and Leakey, R.E.F. (Eds.), Earliest Man and Environments in the Lake Rudolf Basin. Chicago: University of Chicago Press, pp. 565573.Google Scholar
Chavaillon, J. (1979). Stratigraphie du site archéologique de Melka-Kunturé (Ethiopie). Bulletin de la Société géologique de France (sér. 7) 21, 227232.Google Scholar
Chavaillon, J. (1980). Chronologie archéologique de Melka-Kunturé (Ethiopie). In: Leakey, R.E. and Ogot, B.A. (Eds.), Proceedings of the 8th Panafrican Congress of Prehistory and Quaternary Studies, 1977. Nairobi: The International Louis Leakey Memorial Institute for African Prehistory, pp. 200201.Google Scholar
Chavaillon, J. (1982). Les habitats paléolithiques de Melka-Kunturé. Proceedings of the Xth IUSPP, Mexico, 19–24 October 1981, pp. 149168.Google Scholar
Chavaillon, J. and Berthelet, A. (2004). The archaeological sites of Melka Kunture. In: Chavaillon, J. and Piperno, M. (Eds.), Studies on the Early Paleolithic site of Melka Kunture, Ethiopia. Florence: Istituto Italiano di Preistoria e Protostoria, pp. 2580.Google Scholar
Chavaillon, J. and Chavaillon, N. (1971). Présence éventuelle d’un abri oldowayen dans le gisement de Melka-Kunturé (Ethiopie). Comptes rendus de l’Académie des Sciences 273, 623625.Google Scholar
Chavaillon, J. and Coppens, Y. (1975). Découverte d’Hominidé dans un site acheuléen de Melka-Kunturé. Bulletins et mémoires de la Société d’anthropologie Paris 2, 125128.Google Scholar
Chavaillon, J. and Coppens, Y. (1986). Nouvelle découverte d’Homo erectus à Melka-Kunturé. Comptes Rendu de l’Académie des Sciences (II) 1, 99104.Google Scholar
Chavaillon, J. and Piperno, M. (2004a). History of excavations at Melka Kunture. In: Chavaillon, J. and Piperno, M. (Eds.), Studies on the Early Paleolithic site of Melka Kunture, Ethiopia. Florence: Istituto Italiano di Preistoria e Protostoria, pp. 323.Google Scholar
Chavaillon, J. and Piperno, M. (Eds.) (2004b). Studies on the Early Paleolithic site of Melka Kunture, Ethiopia. Florence: Istituto Italiano di Preistoria e Protostoria.Google Scholar
Chavaillon, J. and Taieb, M. (1968). Stratigraphie du Quaternaire de Melka-Kunturé, vallée de l’Awash, Ethiopie. Premiers résultats. Comptes Rendu de l’Académie des Sciences (D) 266, 12101212.Google Scholar
Chavaillon, J., Brahimi, C. and Coppens, Y. (1974). Première découverte d’Hominidé dans l’un des sites acheuléens de Melka-Kunturé (Ethiopie), Comptes Rendu de l’Académie des Sciences (D) 278, 32993302.Google Scholar
Chavaillon, J., Chavaillon, N., Coppens, Y. and Senut, B. (1977). Présence d’Hominidé dans le site oldowayen de Gomboré I à Melka-Kunturé. Comptes Rendu de l’Académie des Sciences (D) 285, 961963.Google Scholar
Chavaillon, J., Chavaillon, N., Hours, F. and Piperno, M. (1979). From the Oldowan to the Middle Stone Age at Melka Kunture (Ethiopia). Understanding cultural changes. Quaternaria 21, 126.Google Scholar
Chavaillon, J., Hours, F. and Coppens, Y. (1987). Découverte de restes humains fossiles associés à un outillage acheuléen final à Melka-Kunturé (Ethiopie). Comptes Rendu de l’Académie des Sciences II 10, 539542.Google Scholar
Chazan, M. (2015a). The Earlier Stone Age lithic assemblage from Wonderwerk Cave, Northern Cape Province, South Africa. In: Thiaw, I. and Bocoum, H. (Eds.), Preserving African Cultural Heritage. Proceedings of the 13th Congress of the Panafrican Archaeological Association for Prehistory and Related Studies – PAA and of the 20th Meeting of the Society of Africanist Archaeologists – Safa. Dakar: Memoires de IIFAN – C. A. DIOP, no. 93, pp. 247252.Google Scholar
Chazan, M. (2015b). Technological trends in the Acheulean of Wonderwerk Cave, South Africa. African Archaeological Review 32, 701728.Google Scholar
Chazan, M. and Horwitz, L.K. (2015). An overview of recent research at Wonderwerk Cave, South Africa. In: Thiaw, I. and Bocoum, H. (Eds.), Preserving African Cultural Heritage. Proceedings of the 13th Congress of the Panafrican Archaeological Association for Prehistory and Related Studies – PAA and of the 20th Meeting of the Society of Africanist Archaeologists – Safa. Dakar: Memoires de IIFAN – C. A. DIOP, no. 93, pp. 253261.Google Scholar
Chazan, M., Ron, H., Matmon, A., et al. (2008). Radiometric dating of the Earlier Stone Age sequence in Excavation I at Wonderwerk Cave, South Africa: preliminary results. Journal of Human Evolution 55(1), 111.Google Scholar
Chazan, M., Avery, M.D., Bamford, M.K., et al. (2012). The Oldowan horizon in Wonderwerk Cave (South Africa): archaeological, geological, paleontological and paleoclimatic evidence. Journal of Human Evolution 63, 859866.Google Scholar
Chazan, M., Porat, N., Sumner, T.A. and Horwitz, L.K. (2013). The use of OSL dating in unstructured sands: the archaeology and chronology of the Hutton Sands at Canteen Kopje (Northern Cape Province, South Africa). Archaeological and Anthropological Sciences 5(4), 351363.Google Scholar
Chazan, M., Horwitz, L.K., Ecker, M., et al. (2017). Renewed excavations at Wonderwerk Cave, South Africa. Evolutionary Anthropology 26(6), 258260.Google Scholar
Chazan, M., Berna, F., Brink, J., et al. (2020). Archeology, environment, and chronology of the early middle stone age component of Wonderwerk cave. Journal of Paleolithic Archaeology 3(3), 302335.Google Scholar
Chen, F.-C. and Li, W.-H. (2001). Genomic divergences between humans and other hominoids and the effective population size of the common ancestor of humans and chimpanzees. The American Journal of Human Genetics 68(2), 444456.Google Scholar
Cheney, D.L. and Seyfarth, R.M. (1981). Selective forces affecting the predator alarm calls of vervet monkeys. Behaviour 76, 2561.Google Scholar
Chernet, T., Hart, W.K., Aronson, J.L. and Walter, R.C. (1998). New age constraints on the timing of volcanism and tectonism in the northern Main Ethiopian Rift–southern Afar transition zone (Ethiopia). Journal of Volcanology and Geothermal Research 80, 267280.Google Scholar
Chorowicz, J. (2005). The East African rift system. Journal of African Earth Sciences, 43, 379410.Google Scholar
Churcher, C.S. (1956). The fossil Hyracoidea of the Transvaal and Taungs deposits. Annals of the Transvaal Museum 22, 477501.Google Scholar
Churcher, C.S. (2000). Extinct equids from Limeworks Cave and Cave of Hearths, Makapansgat, Northern Province, and a consideration of variation in the cheek teeth of Equus capensis Broom. Palaeontologica Africana 36, 97117.Google Scholar
Churcher, C.S. (2006). Distribution and history of the Cape zebra (Equus capensis) in the Quaternary of Africa. Transactions of the Royal Society of South Africa 61, 8995.Google Scholar
Churcher, C.S. and Hooijer, D.A. (1980). The Olduvai zebra (Equus oldowayensis) from the later Omo beds, Ethiopia. Zoologische Mededelingen 55(22), 265281.Google Scholar
Churcher, C.S., Kleindienst, M.R. and Schwarcz, H.P. (1999). Faunal remains from a Middle Pleistocene lacustrine marl in Dakhleh Oasis, Egypt: palaeoenvironmental reconstructions. Palaeogeography, Palaeoclimatology, Palaeoecology 154, 301312.Google Scholar
Churcher, C.S., Kleindienst, M.R., Wiseman, M.F. and McDonald, M.M.A. (2008). The Quaternary faunas of Dakhleh Oasis, Western Desert of Egypt. In: Wiseman, M.F. (Ed.), The Oasis Papers 2. Proceedings of the Second International Conference on the Dakhleh Oasis Project. Oxford: Oxbow Books, pp. 324.Google Scholar
Churchill, S.E., Brink, J.S., Berger, L.R., et al. (2000a). Erfkroon: a new Florisian fossil locality from fluvial contexts in the western Free State, South Africa. South African Journal of Science 96, 161163.Google Scholar
Churchill, S.E., Berger, L.R. and Parkington, J.E. (2000b). A Middle Pleistocene human tibia from Hoedjiespunt, Western Cape, South Africa. South African Journal of Science 96(7), 367369.Google Scholar
Cipriani, L. (1928). Skull “Australopithecus africanus” Dart. Archivio per l’antropologia e la etnologia LVII, 129.Google Scholar
Clark, J.D. (1974). The stone artefacts from Cornelia, O.F.S., South Africa. Memoir – Nasionale Museum, Bloemfontein 9, 3362.Google Scholar
Clark, J.D. (1979). Hominid occupation of the East-Central Highlands of Ethiopia in the Plio-Pleistocene. Nature 282, 3339.Google Scholar
Clark, J.D. (1987). Transitions: Homo erectus and the Acheulian: the Ethiopian sites of Gadeb and the Middle Awash. Journal of Human Evolution 16, 809826.Google Scholar
Clark, J.D. (1993). Stone artefact assemblages from Members 1–3, Swartkrans Cave. In: Brain, C.K. (Ed.), Swartkrans: A Cave’s Chronicle of Early Man. Transvaal Museum Monograph No. 8. Pretoria: Transvaal Museum, pp. 167194.Google Scholar
Clark, J.D. and Haynes, C.V. (1970). An elephant butchery site at Mwanganda’s Village, Karonga, Malawi and its relevance to palaeolithic archaeology. World Archaeology I, 390411.Google Scholar
Clark, J.D., Oakley, K.P., Wells, L.H. and McCleland, J.A.C. (1947). New studies on Rhodesian Man. Journal of the Royal Anthropological Institute of Great Britain and Ireland 77, 732.Google Scholar
Clark, J.D., Stephens, E.A. and Coryndon, S.C. (1966). Pleistocene fossiliferous lake beds of the Malawi (Nyasa) Rift: a preliminary report. American Anthropologist 68, 4649.Google Scholar
Clark, J.D., Haynes, C.V., Mawby, F.E. and Gauthier, A. (1970). Interim report on palaeo-anthropological investigations in the Lake Malawi Rift. Quaternaria 13, 305354.Google Scholar
Clark, J.D., Asfaw, B., Assefa, G., et al. (1984). Palaeoanthropological discoveries in the Middle Awash Valley, Ethiopia. Nature 307(5950), 423428.Google Scholar
Clark, J.D., de Heinzelin, J., Schick, D., et al. (1994). African Homo erectus: old radiometric ages and young Oldowan assemblages in the Middle Awash Valley, Ethiopia. Science 264, 19071910.Google Scholar
Clark, J.D., Beyene, Y., WoldeGabriel, G., et al. (2003). Stratigraphic, chronological and behavioural contexts of Pleistocene Homo sapiens from Middle Awash, Ethiopia. Nature 423(6941), 747.Google Scholar
Clark, J.L. and Plug, I. (2008). Animal exploitation strategies during the South African Middle Stone Age: Howiesons Poort and post-Howiesons Poort fauna from Sibudu Cave. Journal of Human Evolution 54, 886898.Google Scholar
Clarke, R.J. (1977). The cranium of the Swartkrans hominid, SK 847 and its relevance to human origins. PhD thesis, University of the Witwatersrand, Johannesburg.Google Scholar
Clarke, R.J. (1985). Australopithecus and early Homo in southern Africa. In: Delson, E. (Ed.), Ancestors: the Hard Evidence. New York: Alan R. Liss, pp. 171177.Google Scholar
Clarke, R.J. (1988). A new Australopithecus cranium from Sterkfontein and its bearing on the ancestry of Paranthropus. In: Grine, F.E. (Ed.), Evolutionary History of the ‘Robust’ Australopithecines. New York: Routledge, pp.285292.Google Scholar
Clarke, R.J. (1994). On some new interpretations of Sterkfontein stratigraphy. South African Journal of Science 90, 211214.Google Scholar
Clarke, R.J. (1998). The first ever discovery of a well-preserved skull and associated skeleton of Australopithecus. South African Journal of Science 94, 460463.Google Scholar
Clarke, R.J. (1999). Discovery of complete arm and hand of the 3.3 million-year-old Australopithecus skeleton from Sterkfontein. South African Journal of Science 95, 477480.Google Scholar
Clarke, R.J. (2002a). Newly revealed information on the Sterkfontein Member 2 Australopithecus skeleton. South African Journal of Science 98, 523526.Google Scholar
Clarke, R.J. (2002b). On the unrealistic ‘revised age estimates’ for Sterkfontein. South African Journal of Science 98, 415419.Google Scholar
Clarke, R.J. (2006). A deeper understanding of the stratigraphy of the Sterkfontein fossil hominid site. Transactions of the Royal Society of Africa 61, 111120.Google Scholar
Clarke, R.J. (2012). A brief review of history and results of 40 years of Sterkfontein excavations. African Genesis: Perspectives on Hominin Evolution, 62, 120.Google Scholar
Clarke, R.J. and Howell, F.C. (1972). Affinities of the Swartkrans 847 hominin cranium. American Journal of Physical Anthropologie 37, 319336.Google Scholar
Clarke, R.J. and Kuman, K. (2019). The skull of StW 573, a 3.67 Ma Australopithecus prometheus skeleton from Sterkfontein caves, South Africa. Journal of Human Evolution 134, 102634.Google Scholar
Clarke, R.J., Howell, F.C. and Brain, C.K. (1970). More evidence of an advanced hominid at Swartkrans. Nature 225, 12191222.Google Scholar
Clauzon, G., Suc, J.-P., Gautier, F., Berger, A. and Loutre, M.-F. (1996). Alternate interpretation of the Messinian salinity crisis: controversy resolved? Geology 24(4), 363366.Google Scholar
Clementz, M.T. and Koch, P.L. (2001). Differentiating aquatic mammal habitat and foraging ecology with stable isotopes in tooth enamel. Oecologia 129(3), 461472.Google Scholar
Clementz, M.T., Holroyd, P.A. and Koch, P.L. (2008). Identifying aquatic habits of herbivorous mammals through stable isotope analysis. Palaios 23(9), 574585.Google Scholar
Close, R.A., Evers, S.W., Alroy, J. and Butler, R.J. (2018). How should we estimate diversity in the fossil record? Testing richness estimators using sampling‐standardised discovery curves. Methods in Ecology and Evolution 9(6), 13861400.Google Scholar
Codron, D., Codron, J., Lee-Thorp, J., et al. (2007). Stable isotope characterization of mammalian predator–prey relationships in a South African savanna. European Journal of Wildlife Research 53(3), 161170.Google Scholar
Codron, D., Brink, J.S., Rossouw, L. and Clauss, M. (2008). The evolution of ecological specialization in southern African ungulates: competition or physical environmental turnover? Oikos 117(3), 344353.Google Scholar
Coe, M. (1980). The role of modern ecological studies in the reconstruction of paleoenvironments in sub-Saharan Africa. In: Behrensmeyer, A.K. and Hill, A.P. (Eds.), Fossils in the Making: Vertebrate Taphonomy and Paleoecology. Chicago: The University of Chicago Press, pp. 5567.Google Scholar
Coetzee, J.A. (1993). African flora since the terminal Jurassic. In: Goldblatt, P. (Ed.), Biological Relationships between Africa and South America. New Haven: Yale University Press.Google Scholar
Coetzee, L. and Brink, J.S. (2003). Fossil oribatid mites (Acari, Oribatida) from the Florisbad Quaternary deposits, South Africa. Quaternary Research 59(2), 246254.Google Scholar
Coffing, K., Feibel, C.S., Leakey, M.G. and Walker, A. (1994). Four-million-year old hominids from East Lake Turkana, Kenya. American Journal of Physical Anthropology 93, 5565.Google Scholar
Cohen, A., Arrowsmith, R., Behrensmeyer, A.K., et al. (2009). Understanding paleoclimate and human evolution through the Hominin Sites and Paleolakes Drilling Project. Scientific Drilling 8, 6065.Google Scholar
Cohen, B., O’Regan, H. and Steininger, C. (2019). Mongoose manor: Herpestidae remains from the Early Pleistocene Cooper’s D locality in the Cradle of Humankind, Gauteng, South Africa. Palaeontologia Africana, 53.Google Scholar
Cohen, K.M. and Gibbard, P. (2011). Global Chronostratigraphical Correlation Table for the last 2.7 Million Years. Cambridge: Subcommission on Quaternary Stratigraphy (International Commission on Stratigraphy).Google Scholar
Colcord, D.E., Shilling, A.M., Sauer, P.E., et al. (2018). Sub-Milankovitch paleoclimatic and paleoenvironmental variability in East Africa recorded by Pleistocene lacustrine sediments from Olduvai Gorge, Tanzania. Palaeogeography, Palaeoclimatology, Palaeoecology 495, 284291.Google Scholar
Colin, C., Siani, G., Segueni, F., et al. (2008). Restitution de l’histoire de la mousson nord-africaine entre 6,2 et 4,9 Ma et relations possibles avec les évènements tardi-Miocène. Comptes Rendus Geoscience 340, 749760.Google Scholar
Colin, C., Siani, G., Liu, Z., et al. (2014). Late Miocene to early Pliocene climate variability off NW Africa (ODP Site 659). Palaeogeography, Palaeoclimatology, Palaeoecology 401, 8195.Google Scholar
Collins, J.A., Carr, A.S., Schefuß, E., Boom, A. and Sealy, J. (2017). Investigation of organic matter and biomarkers from Diepkloof Rock Shelter, South Africa: insights into Middle Stone Age site usage and palaeoclimate. Journal of Archaeological Science 85, 5165.Google Scholar
Collinson, M.E. and Hooker, J.J. (1991). Fossil evidence of interactions between plants and plant-eating mammals. Philosophical Transactions of the Royal Society of London 333, 197208.Google Scholar
Collinson, M.E. and Hooker, J.J. (2003). Paleogene vegetation of Eurasia: framework for mammalian faunas. Deinsea 10, 4183.Google Scholar
Compton, J.S. (2011). Pleistocene sea-level fluctuations and human evolution on the southern coastal plain of South Africa. Quaternary Science Reviews 30, 506527.Google Scholar
Conroy, G.C. (1996). The cave breccias of Berg Aukas, Namibia: a clustering approach to mine dump paleontology. Journal of Human Evolution 30(4), 349355.Google Scholar
Conroy, G.C., Jolly, C.J., Cramer, D. and Kalb, J.E. (1978). Newly discovered fossil hominid skull from the Afar depression, Ethiopia. Nature 276, 6770.Google Scholar
Conroy, G.C., Senut, B., Van Couvering, J. and Mein, P. (1992). Otavipithecus namibiensis, first Miocene hominoid from southern Africa. Nature, 356(6365), 144148.Google Scholar
Conroy, G.C., Pickford, M., Senut, B. and Mein, P. (1993). Diamonds in the desert: the discovery of Otavipithecus namibiensis. Evolutionary Anthropology: Issues, News, and Reviews 2(2), 4652.Google Scholar
Conroy, G.C., Senut, B., Gommery, D., Pickford, M. and Mein, P. (1996). Brief communication: new primate remains from the Miocene of Namibia, southern Africa. American Journal of Physical Anthropology 99(3), 487492.Google Scholar
Cooke, H.B.S. (1963). Pleistocene mammal faunas of Africa, with particular reference to southern Africa. In: Howell, F.C. and Bourlière, F. (Eds.), African Ecology and Human Evolution. Chicago: Aldine, pp. 65116.Google Scholar
Cooke, H.B.S. (1967). The Pleistocene sequence in South Africa and problems of correlation. In: Bishop, W.W. and Clark, J.D. (Eds.), Background to Evolution in Africa. Chicago: University of Chicago Press, pp. 175184.Google Scholar
Cooke, H.B.S. (1974). The fossil mammals of Cornelia, OFS, South Africa. Memoires of the National Museum, Bloemfontein 9, 6384.Google Scholar
Cooke, H.B.S. (1976). Suidae from Plio-Pleistocene strata of the Rudolf Basin. In: Coppens, Y., Howell, F.C., Isaac, G.L. and Leakey, R.E.F. (Eds.), Earliest Man and Environments in the Lake Rudolf Basin. Chicago: University of Chicago Press, pp. 251263.Google Scholar
Cooke, H.B.S. (1978a). Suid evolution and correlation of African hominid localities: an alternative taxonomy. Science 201, 460463.Google Scholar
Cooke, H.B.S. (1978b). Plio-Pleistocene Suidae from Hadar, Ethiopia. Kirtlandia 29, 163.Google Scholar
Cooke, H.B.S. (1978c). Faunal evidence for the biotic setting of early African hominids. In Jolly, C.J. (Ed.), Early Hominids of Africa. New York: St. Martin’s Press, pp. 267284.Google Scholar
Cooke, H.B.S. (1982). Phacochoerus modestus from Bed I, Olduvai Gorge, Tanzania. Zeitschrift der Geologischen Wissenschaften 10, 899908.Google Scholar
Cooke, H.B.S. (1983). Human evolution: the geological framework. Canadian Journal of Anthropology 3, 143161.Google Scholar
Cooke, H.B.S. (1988). The larger mammals from Cave of Hearths. In: Mason, R. (Ed.), Cave of Hearths, Makapansgat, Transvaal. Johannesburg: University of the Witwatersrand Archaeological Research Unit, pp. 507534.Google Scholar
Cooke, H.B.S. (1990a). Taung fossils in the University of California collections. In Sperber, G.H. (Ed.), From Apes to Angels. New York: Wiley-Liss, pp. 119134.Google Scholar
Cooke, H.B.S. (1990b). Suid remains from the upper Semliki area, Zaire. In: Boaz, N.T. (Ed.), Evolution of Environments and Hominidae in the African Western Rift Valley. Martinsville: Virginia Museum of Natural History, pp. 197201.Google Scholar
Cooke, H.B.S. (1994). Phacochoerus modestus from Sterkfontein Member 5. South African Journal of Science 90, 99100.Google Scholar
Cooke, H.B.S. (1997). The status of the African fossil suids Kolpochoerus limnetes (Hopwood, 1926), K. phacochoeroides (Thomas, 1884) and K. afarensis (Cooke, 1978). Geobios 30, 121126.Google Scholar
Cooke, H.B.S. (2005). Makapansgat suids and Metridiochoerus. Paleontologica Africana 41, 131140.Google Scholar
Cooke, H.B.S. (2007). Stratigraphic variation in Suidae from the Shungura Formation and some coeval deposits. In: Bobe, R., Alemseged, Z. and Behrensmeyer, A.K. (Eds.), Hominin Environments in the East African Pliocene: An Assessment of the Faunal Evidence. New York: Springer, pp. 107127.Google Scholar
Cooke, H.B.S. and Ewer, R.F. (1972). Fossil Suidae from Kanapoi and Lothagam, northwestern Kenya. Bulletin of the Museum of Comparative Zoology 143, 149296.Google Scholar
Cooke, H.B.S. and Wells, L.W. (1951). Fossil remains from Chelmer, near Bulawayo, Southern Rhodesia. South African Journal of Science 57, 205209.Google Scholar
Cooke, H.B.S. and Wilkinson, A.F. (1978). Suidae and Tayassuidae. In: Maglio, V.J. and Cooke, H.B.S. (Eds.,) Evolution of African Mammals. Cambridge, MA: Harvard University Press, pp. 435482.Google Scholar
Cooke, H.B.S., Malan, B.D. and Wells, L.H. (1945). 3. Fossil man in the Lebombo Mountains, South Africa: the ‘Border Cave’, Ingwavuma District, Zululand. Man, 45, 613.Google Scholar
Coombs, M.C. and Cote, S.M. (2010). Chalicotheriidae. In: Werdelin, L. and Sanders, W.J. (Eds.), Cenozoic Mammals of Africa. Berkeley: University of California Press, pp. 659667.Google Scholar
Copeland, S.R., Sponheimer, M., Spinage, C.A., et al. (2009). Stable isotope evidence for impala Aepyceros melampus diets at Akagera National Park, Rwanda. African Journal of Ecology 47(4), 490501.Google Scholar
Copeland, S.R., Sponheimer, M., Lee-Thorp, J.A., et al. (2010). Using strontium isotopes to study site accumulation processes. Journal of Taphonomy 8, 115127.Google Scholar
Copeland, S.R., Sponheimer, M., de Ruiter, D.J., et al. (2011). Strontium isotope evidence for landscape use by early hominins. Nature 474, 7679.Google Scholar
Coppa, A., Abbate, E., Bondioli, L., et al. (2012). Mulhuli-Amo, a new late Early Pleistocene paleoanthropological site in the northern Danakil Depression, Eritrea. Proceedings of the ESHE Meetings 1, 5859 (abstract).Google Scholar
Coppa, A., Bondioli, L., Candilio, F., et al. (2014). New 1 Ma old human cranial remains from Mulhuli-Amo, near Uadi Aalad, Danakil (Afar) depression of Eritrea. American Journal of Physical Anthropology 153(S58), 97.Google Scholar
Coppens, Y. (1966). An early hominid from Chad. Current Anthropology 7, 584585.Google Scholar
Coppens, Y. (1975). Evolution des hominidés et de leur environnement au cours du Plio-Pléistocène dans la basse vallée de l’Omo en Ethiopie. Comptes Rendu de l’Académie des Sciences Paris,Série D, Sciences Naturelles 281, 16931696.Google Scholar
Coppens, Y. (1977). Hominid remains from the Plio/Pleistocene formations of the Omo Basin, Ethiopia. Journal of Human Evolution 6, 169173.Google Scholar
Coppens, Y. (1978). Evolution of the hominids and their environment during the Plio-Pleistocene in the lower Omo Valley, Ethiopia. In: Bishop, W.W. (Ed.), Geological Background to Fossil Man. Edinburgh: Scottish Academic Press, pp. 499506.Google Scholar
Coppens, Y., Howell, F.C., Isaac, G.L. and Leakey, R.E. (1976). Earliest Man and Environments in the Lake Rudolf Basin: Stratigraphy, Paleoecology, and Evolution. Chicago: University of Chicago Press, p. 615.Google Scholar
Cordova, C. and Avery, G. (2017). African savanna elephants and their vegetation associations in the Cape Region, South Africa: opal phytoliths from dental calculus on prehistoric, historic and reserve elephants. Quaternary International 443, 189211.Google Scholar
Cordova, C.E. and Scott, L. (2011). The potential of Poaceae, Cyperaceae, and Restionaceae phytoliths to reflect past environmental conditions in South Africa. In: Runge, J. (Ed.), African Palaeoenviornments and Geomorphic Landscape Evolution. London: CRC Press, pp. 107133.Google Scholar
Corfield, T.F. (1973). Elephant mortality in Tsavo National Park, Kenya. African Journal of Ecology 11, 339368.Google Scholar
Cornelissen, E., Boven, A., Dabi, A., et al. (1990). The Kapthurin Formation revisited. African Archaeological Review 8, 2375.Google Scholar
Coryndon, S.C. (1966). Preliminary report on some fossils from the Chiwondo Beds of the Karonga District, Malawi. American Anthropologist 68, 5966.Google Scholar
Coryndon, S.C. (1970). Evolutionary trends in East African Hippopotamidae. Bulletin de Liaison de l’Association Sénégalaise pour l’Etude du Quaternaire de l’Ouest africain (ASEQUA), Dakar 25, 107116.Google Scholar
Coryndon, S.C. and Coppens, Y. (1973). Preliminary report on Hippopotamidae (Mammalia, Artiodactyla) from the Plio/Pleistocene of the lower Omo basin, Ethiopia. Fossil Vertebrates of Africa 3, 139157.Google Scholar
Cote, S.M. (2004). Origins of the African hominoids: an assessment of the palaeobiogeographical evidence. Comptes Rendus Palevol 3(4), 323340.Google Scholar
Cote, S. (2018). Savannah savvy. Nature Ecology & Evolution 2, 210211.Google Scholar
Coutros, P. (2016). All Climates are Local Climates: Palaeoenvironmental Perspectives from the Diallowali Site System. Society of Africanist Archaeologists, 23. Toulouse: Université Toulouse Jean Jaurès.Google Scholar
Couvreur, T.L.P., Chatrou, L.W., Sosef, M.S.M. and Richardson, J.E. (2008) Molecular phylogenetics reveal multiple tertiary vicariance origins of African rain forest trees. BMC Biology 6, 54.Google Scholar
Cowling, R.M., Cartwright, C.R., Parkington, J.E. and Allsopp, J.C. (1999). Fossil wood charcoal assemblages from Elands Bay Cave, South Africa: implications for Late Quaternary vegetation and climates in the winter‐rainfall fynbos biome. Journal of Biogeography 26(2), 367378.Google Scholar
Cowling, R.M., Procheş, Ş., Vlok, J.H.J. and van Staden, J. (2005). On the origin of southern African subtropical thicket vegetation. South African Journal of Botany 71(1), 123.Google Scholar
Cowling, S.A., Cox, P.M., Jones, C.D., et al. (2008). Simulated glacial and interglacial vegetation across Africa: implications for species phylogenies and trans-African migration of plants and animals. Global Change Biology 14, 827840.Google Scholar
Crader, D.C. (1974). The effects of scavengers on bone material from a large mammal: an experiment conducted among the Bisa of the Luangwa Valley, Zambia. In: Donnan, C.B. and Clewlow, C.W. (Eds.), Ethnoarchaeology: Archaeology Survey Monograph 4. Los Angeles: Institute of Archaeology, University of California, pp. 161173.Google Scholar
Crawford, T., McKee, J., Kuykendall, K., Latham, A. and Conroy, G.C. (2004). Recent paleoanthropological excavations of in situ deposits at Makapansgat, South Africa – a first report. Collegium Antropologicum 28(2), 4357.Google Scholar
Crevecoeur, I., Rougier, H., Grine, F. and Froment, A. (2009). Modern human cranial diversity in the late Pleistocene of Africa and Eurasia: evidence from Nazlet Khater, Pestera cu Oase, and Hofmeyr. American Journal of Physical Anthropology 140, 347358.Google Scholar
Crevecoeur, I., Skinner, M.M., Bailey, S.E., et al. (2014). First early hominin from central Africa (Ishango, Democratic Republic of Congo). PLoS ONE 9, e84652.Google Scholar
Cressier, P. (1980). Magnétostratigraphie du gisement pléistocène de Melka-Kunturé (Ethiopie). Datation des niveaux oldowayens et acheuléens. Thèse, University Louis Pasteur, Strasbourg, France.Google Scholar
Crompton, R.H., Pataky, T.C., Savage, R., et al. (2012). Human-like external function of the foot, and fully upright gait, confirmed in the 3.66 million year old Laetoli hominin footprints by topographic statistics, experimental footprint formation and computer simulation. Journal of the Royal Society Interface 9, 707719.Google Scholar
Crusafont-Pairó, M. (1979). Les Giraffidés des gisements du Bled Douarah (W. de Gafsa, Tunisie). Notes du Service géologique de Tunisie 44, 741.Google Scholar
Cruz-Uribe, K. (1983). The mammalian fauna from Redcliff Cave, Zimbabwe. South African Archaeological Bulletin 38, 716.Google Scholar
Cruz-Uribe, K. (1991). Distinguishing hyena from hominid bone accumulations. Journal of Field Archaeology 18, 467486.Google Scholar
Cruz-Uribe, K., Klein, R.G., Avery, G., et al. (2003). Excavation of buried Late Acheulean (Mid-Quaternary) land surfaces at Duinefontein 2, Western Cape Province, South Africa. Journal of Archaeological Science 30(5), 559575.Google Scholar
Csermely, D. (1996). Antipredator behavior in lemurs: evidence of an extinct eagle on Madagascar or something else? International Journal of Primatology 17: 349354.Google Scholar
Cumming, D.H.M. (1975). A Field Study of the Ecology and Behavior of Warthog. Salisbury, Rhodesia: The Trustees of the National Museums and Monuments of Rhodesia.Google Scholar
Curnoe, D. (2010). A review of early Homo in southern Africa focusing on cranial, mandibular and dental remains, with the description of a new species (Homo gautengensis sp. nov.). Homo 61, 151177.Google Scholar
Curnoe, D., Herries, A.I.R., Brink, J., et al. (2006). Discovery of Middle Pleistocene fossil and stone tool-bearing deposits at Groot Kloof, Ghaap Escarpment, Northern Cape Province. South African Journal of Science 102, 180184.Google Scholar
Curran, S. and Haile-Selassie, Y. (2016). Paleoecological reconstruction of hominin-bearing Middle Pliocene Localities at Woranso-Mille, Ethiopia. Journal of Human Evolution 96, 97112.Google Scholar
Cuthbert, M.O., Gleeson, T., Reynolds, S.C., et al. (2017). Modelling the role of groundwater hydro-refugia in East African hominin evolution and dispersal. Nature Communications 8, 15696.Google Scholar
Cutler, A.H., Behrensmeyer, A.K. and Chapman, R.E. (1999). Environmental information in a recent bone assemblage: roles of taphonomic processes and ecological change. Palaeogeography, Palaeoclimatology, Palaeoecology 149, 359372.Google Scholar
Daly, A.J., Baetens, J.M. and De Baets, B. (2018). Ecological diversity: measuring the unmeasurable. Mathematics, 6(7), 119.Google Scholar
Damuth, J. and Janis, C. (2011). On the relationship between hypsodonty and feeding ecology in ungulate mammals, and its utility in palaeoecology. Biological Reviews 86, 733758.Google Scholar
Damuth, J.D., DiMichele, W.A., Potts, R., Sues, H.D. and Wing, S.L. (1992). Terrestrial Ecosystems Through Time: Evolutionary Paleoecology of Terrestrial Plants and Animals (No. 2). Chicago: University of Chicago Press.Google Scholar
Darlington, J.P.E.C. (2005). Distinctive fossilized termite nests at Laetoli, Tanzania. Insectes Sociaux 52, 408409.Google Scholar
Darlington, J.P.E.C. (2011). Trace fossils interpreted in relation to the extant termite fauna at Laetoli, Tanzania. In: Harrison, T. (Ed.), Paleontology and Geology of Laetoli: Human Evolution in Context. Vol. 2: Fossil Hominins and the Associated Fauna. Dordrecht: Springer, pp. 555565.Google Scholar
Dart, R.A. (1925a). Australopithecus africanus: the man-ape of South Africa. Nature 115, 195199.Google Scholar
Dart, R.A. (1925b). A note on Makapansgat: a site of early human occupation. South African Journal of Science 22, 454.Google Scholar
Dart, R.A. (1926). Taungs and its significance. Natural History 26, 315327.Google Scholar
Dart, R.A. (1929). A note on the Taungs skull. South African Journal of Science 26, 648658.Google Scholar
Dart, R.A. (1948). The Makapansgat proto-human Australopithecus prometheus. American Journal of Physical Anthropology 6, 259281.Google Scholar
Dart, R.A. (1950). A note on the limestone caverns of Leba, near Humpata, Angola. The South African Archaeological Bulletin 5(20), 149151.Google Scholar
Dart, R.A. (1952). Faunal and climatic fluctuations in Makapansgat Valley: their relation to the geologic age and Promethean status of Australopithecus. In Leakey, L.S.B. and Cole, S. (Eds.), Proceedings of the 1st Pan African Congress on Prehistory, Nairobi, 1947. Oxford: Blackwell, pp. 96106.Google Scholar
Dart, R.A. (1955). The first australopithecine fragment from the Makapansgat pebble culture stratum. Nature 176(4473), 170171.Google Scholar
Dart, R.A. (1956). Cultural status of the South African man-apes. Smithsonian Report no. 4240, pp. 317338.Google Scholar
Dart, R.A. (1957). The Osteodontokeratic Culture of Australopithecus prometheus. Transvaal Museum Memoir No. 11. Pretoria: Transvaal Museum.Google Scholar
Dart, R.A. (1959). The first Australopithecus cranium from the pink breccia at Makapansgat. American Journal of Physical Anthropology 17, 7782.Google Scholar
Dart, R.A. (1960). The bone tool-manufacturing ability of Australopithecus prometheus. American Anthropologist 62, 134143.Google Scholar
Dart, R.A. (1962). The gradual appraisal of Australopithecus. In: Kurth, G. (Ed.), Evolution und Hominisation. Stuttgart: Gustav Fischer Verlag, pp. 141156.Google Scholar
Dart, R.A. and Craig, D. (1959). Adventures with the Missing Link. London: Hamish Hamilton.Google Scholar
Darwent, C.M. (2007). Lagomorphs (Mammalia) from the late Miocene deposits at Lemudong’o, southern Kenya. Kirtlandia 56, 112120.Google Scholar
Darwin, C. (1859). Origin of Species by Natural Selection.Google Scholar
Database for Hydrological Time Series of Inland Waters (DAHITI) – Victoria, Lake, retrieved 21 Nov 2017.Google Scholar
Daujeard, C., Geraads, D., Gallotti, R., Mohib, A. and Raynal, J.-P. (2012). Carcass acquisition and consumption by carnivores and hominins in Middle Pleistocene sites of Casablanca (Morocco). Journal of Taphonomy 10, 339363.Google Scholar
Daujeard, C., Geraads, D., Gallotti, R., et al. (2016). Pleistocene hominins as a resource for carnivores: a c.500,000-year-old human femur bearing tooth-marks in North Africa (Thomas Quarry I, Morocco). PLoS ONE 11(4), e0152284. doi:10.1371/journal.pone.0152284Google Scholar
Daujeard, C., Falguères, C., Shao, Q. (2020). Earliest African evidence of carcass processing and consumption in cave at 700ka, Casablanca, Morocco. Scientific Reports 10: 4761.Google Scholar
Dauphin, Y., Kowalski, C. and Denys, C. (1994). Assemblage data and bone and teeth modifications as an aid to paleoenvironmental interpretations of the open-air Pleistocene site of Tighenif (Algeria). Quaternary Research 42, 340349.Google Scholar
Davenport, T.R.B., Stanley, W.T., Sargis, E.J., et al. (2006). A new genus of African monkey, Rungwecebus: morphology, ecology, and molecular phylogenetics. Science 312, 13781381.Google Scholar
Davidson, A. (1983). The Omo River Project. Reconnaissance of Geology and Geochemistry of Parts of Ilubabor, Kefa, Gemu Gofa and Sidamo, Ethiopia. Ethipia: Ministry of Mines and Energy, Ethiopian Institute of Geological Surveys.Google Scholar
Davidson, A., Shiferaw, A., Leul, E.G., Davis, J.C. and Moore, J.M. (1973). Preliminary report on the geology and geochemistry of parts of Sidamo, Gemu Gofa, and Kefa Provinces, Ethiopia. Omo River Project report, no. 1. Imperial Ethiopian Government, Ministry of Mines.Google Scholar
Davidson, I. and Solomon, S. (1990). Was OH 7 the victim of a crocodile attack? In: Solomon, S., Davidson, I. and Watson, D. (Eds.), Problem Solving in Taphonomy: Archaeological and Palaeontological Studies from Europe, Africa and Oceania, Vol. 2. St. Lucia, Queensland: Tempus, pp. 197206.Google Scholar
Dawkins, R. (1986). The Blind Watchmaker. London: Longman.Google Scholar
Dawson, J.B. (2008). The Gregory Rift Valley and Neogene – Recent Volcanoes of Northern Tanzania. London: Geological Society of London.Google Scholar
Day, M.H. (1971). Postcranial remains of Homo erectus from bed IV, Olduvai Gorge, Tanzania. Nature 232, 383387.Google Scholar
Day, M.H. (1986). Guide to Fossil Man, 4th ed. Chicago: Chicago University Press.Google Scholar
Day, M.H. and Napier, J.R. (1964). Hominid fossils from Bed I, Olduvai Gorge, Tanganyika: fossil foot bones. Nature 201, 969970.Google Scholar
Day, M.H. and Wickens, E.H. (1980). Laetoli Pliocene hominid footprints and bipedalism. Nature 286, 385387.Google Scholar
Day, M., Leakey, R., Walker, A. and Wood, B. (1976). New hominids from East Turkana, Kenya. American Journal of Physical Anthropology 45, 369435.Google Scholar
Day, M.H., Leakey, M.D. and Magori, C. (1980). A new hominid fossil skull (L.H. 18) from the Ngaloba Beds, Laetoli, northern Tanzania. Nature 284, 5556.Google Scholar
d’Huart, J.-P. (1978). Ecologie de l’hylochere (Hylochoerus meinertzhageni Thomas) au Parc National des Virunga. Exploration du Parc National des Virunga, Deuxieme Serie, Fascicule 25. Brussels: Fondation pour Favoriser les Recherches scientifiques en Afrique.Google Scholar
de Bonis, L., Peigné, S., Likius, A., et al. (2007). The oldest African fox (Vulpes riffautae n. sp., Canidae, Carnivora) recovered in late Miocene deposits of the Djurab desert, Chad. Naturwissenschaften 94, 575.Google Scholar
de Bonis, L., Peigné, S., Mackaye, H.T., et al. (2008). The fossil vertebrate locality Kossom Bougoudi, Djurab desert, Chad: a window in the distribution of the carnivoran faunas at the Mio-Pliocene boundary in Africa. Comptes Rendus Palevol 7, 571581.Google Scholar
de Bonis, L., Peigné, S., Taisso Mackaye, H., et al. (2010). New sabre-toothed cats in the Late Miocene of Toros Menalla (Chad). Comptes Rendus Palevol 9, 221227.Google Scholar
De Bruijn, H. (1973) Analysis of the data bearing upon the correlation of the Messinian with the succession of land mammals. Messinian Events in the Mediterranean: Proceedings, Koninklijke Nederlandse Akademie van Wetenschappen, 76, 260262.Google Scholar
De Franceschi, D., Bamford, M., Pickford, M. and Senut, B. (2016). Fossil wood from the upper Miocene Mpesida Beds at Cheparain (Baringo District, Kenya): botanical affinities and palaeoenvironmental implications. Journal of African Earth Sciences 115, 271280.Google Scholar
De Graaff, G. (1957). A new chrysochlorid from Makapansgat. Palaeontologica Africana 5, 2127.Google Scholar
De Graaf, G. (1960). A preliminary investigation of the mammalian microfauna in Pleistocene deposits in the Transvaal System. Palaeontologia Africana 7, 59118.Google Scholar
de Heinzelin, J. (1962). Les formations du Western Rift et de la cuvette congolaise. Annales du Musée Royal de l’Afrique Centrale 40, 129243.Google Scholar
de Heinzelin, J. (1983a). Paleoenvironments. In: de Heinzelin, J. (Ed.), The Omo Group. Tervuren: Musée Royal de l’Afrique Centrale, pp. 209218.Google Scholar
de Heinzelin, J. (Ed.) (1983b). The Omo Group. Annales, Sciences Géologiques, Vol. 85. Tervuren: Musée Royal de l’Afrique Centrale.Google Scholar
de Heinzelin, J. and Haesaerts, P. (1983a) The Shungura Formation. In: de Heinzelin, J. (Ed.), The Omo Group. Tervuren: Musée Royal de l’Afrique Centrale, pp. 25127.Google Scholar
de Heinzelin, J. and Haesaerts, P. (1983b). The Usno Formation. In: de Heinzelin, J. (Ed.), The Omo Group. Tervuren: Musée Royal de l’Afrique Centrale, pp. 129139.Google Scholar
de Heinzelin, J. and Haesaerts, P. (1983c). The Mursi Formation. In: de Heinzelin, J. (Ed.), The Omo Group. Tervuren: Musée Royal de l’Afrique Centrale, pp. 141143.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. In: Coppens, Y., Howell, F.C., Isaac, G.L. and Leakey, R.E. (Eds.), Earliest Man and Environments in the Lake Rudolf Basin. Chicago: University of Chicago Press, pp. 2449.Google Scholar
de Heinzelin, J., Clark, J.D., White, T.D., et al. (1999). Environment and behavior of 2.5-million-year-old Bouri hominids. Science 284, 625629.Google Scholar
de Heinzelin, J., Clark, J.D. and Gilbert, W.H. (2000). The Acheulean and Plio-Pleistocene deposits of the Middle Awash valley, Ethiopia, Annales – Sciences Géologiques Vol. 104. Tervuren: Musée Royale de l’Afrique Central.Google Scholar
de la Torre, I. (2006). Estrategias tecnológicas en el Pleistoceno inferior de África oriental (Olduvai y Peninj, norte de Tanzania). Madrid: Servicio de Publicaciones Universidad Complutense de Madrid.Google Scholar
de la Torre, I. (2009). Technological strategies in the Lower Pleistocene at Peninj (West of Lake Natron, Tanzania). In: Schick, K. and Toth, N. (Eds.), The Cutting Edge: New Approaches to the Archaeology of Human Origins. Bloomington: Stone Age Institute Press, pp. 93113.Google Scholar
de la Torre, I. (2011). The Early Stone Age lithic assemblages of Gadeb (Ethiopia) and the developed Oldowan/early Acheulean in East Africa. Journal of Human Evolution 60, 768812.Google Scholar
de la Torre, I. and Benito-Calvo, A. (2013). Application of GIS methods to retrieve orientation patterns from imagery; a case study from Beds I and II, Olduvai Gorge (Tanzania). Journal of Archaeological Science 40, 24462457.Google Scholar
de la Torre, I. and Mora, R. (2004). El Olduvayense de la Sección Tipo de Peninj (Lago Natron, Tanzania), vol. 1. Barcelona: CEPAP.Google Scholar
de la Torre, I., Mora, R., Domínguez-Rodrigo, M., Luque, L. and Alcalá, L. (2003). The Oldowan industry of Peninj and its bearing on the reconstruction of the technological skills of Lower Pleistocene hominids. Journal of Human Evolution 44, 203224.Google Scholar
de la Torre, I., deBeaune, S., Davidson, I., et al. (2004). Omo revisited: evaluating the technological skills of Pliocene hominids. Current Anthropology 45(4), 439465.Google Scholar
de la Torre, I., Mora, R. and Martínez-Moreno, J. (2008). The early Acheulean in Peninj (Lake Natron, Tanzania). Journal of Anthropological Archaeology 27(2), 244264.Google Scholar
de la Torre, I., McHenry, L. and Njau, J. (2018). From the Oldowan to the Acheulean at Olduvai Gorge, Tanzania – an introduction to the special issue. Journal of Human Evolution 120, 16.Google Scholar
de Lapparent de Broin, F. (2000). African chelonians from the Jurassic to the present: phases of development and preliminary catalogue of the fossil record. Palaeontologia Africana 36, 4382.Google Scholar
de Lumley, M.-A. and Marchal, F. (2004). Les restes d’hominidés du site de FJ-1. In: de Lumley, H. and Beyene, Y. (Eds.), Les sites préhistoriques de la région de Fejej, Sud-Omo, Éthiopie, dans leur contexte stratigraphique et paléontologique. Paris: Éditions Recherche sur les Civilisations, pp. 203340.Google Scholar
de Ruiter, D.J. (2001). A methodological analysis of the relative abundance of hominids and other macromammals from the site of Swartkrans, South Africa. PhD thesis, Faculty of Science, University of the Witwatersrand, Johannesburg.Google Scholar
de Ruiter, D.J. (2003). Revised faunal lists for Members 1–3 of Swartkrans, South Africa. Annals of the Transvaal Museum 40, 2941.Google Scholar
de Ruiter, D.J. (2004). Relative abundance, skeletal part representation and accumulating agents of macromammals at Swartkrans. In: Brain, C. (Ed.), Swartkrans: A Cave’s Chronicle of Early Man, 2nd ed. Pretoria: Transvaal Museum Press, pp. 265278.Google Scholar
de Ruiter, D.J. and Berger, L.R. (2000). Leopards as taphonomic agents in dolomitic caves – implications for bone accumulations in the hominid-bearing deposits of South Africa. Journal of Archaeological Science 27(8), 665684.Google Scholar
de Ruiter, D.J., Sponheimer, M. and Lee-Thorp, J.A. (2008a). Indications of habitat association of Australopithecus robustus in the Bloubank Valley, South Africa. Journal of Human Evolution 55, 10151030.Google Scholar
de Ruiter, D., Brophy, J.K., Lewis, P.J., Churchill, S.E. and Berger, L.R. (2008b). Faunal assemblage composition and paleoenvironment of Plovers Lake, a Middle Stone Age locality in Gauteng Province, South Africa. Journal of Human Evolution 55, 11021117.Google Scholar
de Ruiter, D.J., Pickering, R., Steininger, C.M., et al. (2009). New Australopithecus robustus fossils and associated U–Pb dates from Cooper’s Cave (Gauteng, South Africa). Journal of Human Evolution 56, 497513.Google Scholar
de Ruiter, D.J., Copeland, S.R., Lee-Thorp, J.A. and Sponheimer, M. (2010a). Investigating the role of eagles as accumulating agents in the dolomitic cave infills of South Africa. Journal of Taphonomy 8, 129154.Google Scholar
de Ruiter, D.J., Brophy, J.K., Lewis, P.J., et al. (2010b). Preliminary investigation of Matjhabeng, a Pliocene fossil locality in the Free State of South Africa. Palaeontologica Africana, 45, 1122.Google Scholar
de Ruiter, D.J., Churchill, S.E. and Berger, L.R. (2013). Australopithecus sediba from Malapa, South Africa. In: Reed, K.E., Fleagle, J.G. and Leakey, R.E.F. (Eds.), The Paleobiology of Australopithecus. Dordrecht: Springer, pp. 147160.Google Scholar
d’Errico, F., Backwell, L., Villa, P., et al. (2012a). Early evidence of San material culture represented by organic artifacts from Border Cave, South Africa. Proceedings of the National Academy of Sciences 109(33), 1321413219.Google Scholar
d’Errico, F., Moreno, R.G. and Rifkin, R.F. (2012b). Technological, elemental and colorimetric analysis of an engraved ochre fragment from the Middle Stone Age levels of Klasies River Cave 1, South Africa. Journal of Archaeological Science 39(4), 942952.Google Scholar
Deacon, H.J. (1970). The Acheulian occupation at Amanzi Springs, Uitenhage District, Cape Province. Annals of the Cape Provincial Museums 8, 89189.Google Scholar
Deacon, H.J. (1979). Excavations at Boomplaas cave – a sequence through the upper Pleistocene and Holocene in South Africa. World Archaeology, 10(3), 241257.Google Scholar
Deacon, H.J. (1995). Two late Pleistocene–Holocene archaeological depositories from the southern Cape, South Africa. The South African Archaeological Bulletin 50, 121131.Google Scholar
Deacon, H.J. and Geleijnse, V.B. (1988). The stratigraphy and sedimentology of the main site sequence, Klasies River, South Africa. The South African Archaeological Bulletin, 43, 514.Google Scholar
Deacon, H.J., Deacon, J., Scholtz, A., et al. (1984). Correlation of palaeoenvironmental data from the Late Pleistocene and Holocene deposits at Boomplaas Cave, Southern Cape. In: Vogel, J.C. (Ed.), Late Cainozoic Palaeoclimates of the Southern Hemisphere. Rotterdam: Balkema, pp. 339352.Google Scholar
Deacon, J. (1984). The Later Stone Age of southernmost Africa. International Series 213. Oxford: British Archaeological Reports.Google Scholar
Deane, A.S. (2009). Early Miocene catarrhine dietary behaviour: the influence of the Red Queen effect on incisor shape and curvature. Journal of Human Evolution 56, 275285.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 fossils. In: Coppens, Y. (Ed.), L’Environnement des Hominidés au Plio-Pléistocène. Paris: Masson, pp. 175222.Google Scholar
Dechamps, R. and Maes, F. (1990). Woody plant communities and climate in the Pliocene of the Semliki Valley, Zaire. In: Boaz, N.T. (Ed.), Evolution of Environments and Hominidae in the African Western Rift Valley. Martinsville: Virginia Museum of Natural History, pp. 7194.Google Scholar
Dechamps, R., Senut, B. and Pickford, M. (1992). Fruits fossiles pliocènes et pléistocènes du Rift occidental ougandais. Signification paléoenvironnementale. Comptes rendus de l’Académie des sciences. Série 2, Mécanique, Physique, Chimie, Sciences de l’univers, Sciences de la Terre 314(3), 325331.Google Scholar
DeGusta, D. and Vrba, E. (2003). A method for inferring paleohabitats from the functional morphology of bovid astragali. Journal of Archaeological Science 30, 10091022.Google Scholar
DeGusta, D. and Vrba, E. (2005). Methods for inferring paleohabitats from the functional morphology of bovid phalanges. Journal of Archaeological Science 32, 10991113.Google Scholar
Dehghani, R., Wanntorp, L., Pagani, M., et al. (2008). Phylogeography of the white-tailed mongoose (Herpestidae, Carnivora, Mammalia) based on partial sequences of the mtDNA control region. Journal of Zoology 276, 385393.Google Scholar
Deino, A.L. (2011). 40Ar/39Ar dating of Laetoli, Tanzania. In: Harrison, T. (Ed.), Paleontology and Geology of Laetoli: Human Evolution in Context. Volume 1: Geology, Geochronology, Paleoecology and Paleoenvironment. Dordrecht: Springer, pp. 7797.Google Scholar
Deino, A.L. (2012). 40Ar/39Ar dating of Bed I, Olduvai Gorge, Tanzania, and the chronology of early Pleistocene climate change. Journal of Human Evolution 63(2), 251273.Google Scholar
Deino, A.L. and Ambrose, S.H. (2007). 40Ar/39Ar dating of the Lemudong’o late Miocene fossil assemblages, southern Kenya Rift. Kirtlandia 56, 6571.Google Scholar
Deino, A.L. and Hill, A. (2002). Ar-40/Ar-39 dating of Chemeron Formation strata encompassingthe site of hominid KNM-BC 1, Tugen Hills, Kenya. Journal of Human Evolution, 42(1–2), 141151.Google Scholar
Deino, A. and McBrearty, S. (2002). 40Ar/39Ar dating of the Kapthurin Formation, Baringo, Kenya. Journal of Human Evolution 42, 185210.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(B6), 84538470.Google Scholar
Deino, A.L., Tauxe, L., Hill, A. and Monaghan, M. (2002). 40Ar/39Ar geochronology and paleomagnetic stratigraphy of the Lukeino and lower Chemeron succession at Tabarin and Kapcheberek, Tugen Hills, Kenya. Journal of Human Evolution 42, 117140.Google Scholar
Deino, A.L., Kingston, J.D., Glen, J.M., Edgar, R.K. and Hill, A. (2006a). Precessional forcing of lacustrine sedimentation in the late Cenozoic Chemeron Basin, Central Kenya Rift, and calibration of the Gauss/Matuyama boundary. Earth and Planetary Science Letters 247(1–2), 4160.Google Scholar
Deino, A.L., Domínguez-Rodrigo, M. and Luque, L. (2006b). 40Ar/39Ar dating of the Pleistocene Peninj group, lake Natron, Tanzania. AGU Fall Meeting Abstracts, pp. V53C-1771.Google Scholar
Deino, A.L., Scott, G.R., Saylor, B., et al. (2010). 40Ar/39Ar dating, paleomagnetism, and tephrochemistry of Pliocene strata of the hominid-bearing Woranso-Mille area, west-central Afar Rift, Ethiopia. Journal of Human Evolution, 58, 111126.Google Scholar
Deino, A.L., Behrensmeyer, A.K., Brooks, A.S., et al. (2018). Chronology of the Acheulean to Middle Stone Age transition in Eastern Africa. Science 360, 9598.Google Scholar
Deino, A.L., Sier, M.J., Garello, D.I., et al. (2019). Chronostratigraphy of the Baringo–Tugen Hills–Barsemoi (HSPDP-BTB13-1A) core – 40Ar/39Ar dating, magnetostratigraphy, tephrostratigraphy, sequence stratigraphy and Bayesian age modeling. Palaeogeography, Palaeoclimatology, Palaeoecology 570, 109519.Google Scholar
Delagnes, A. and Roche, H. (2005). Late pliocene hominid knapping skills: the case of Lokalalei 2 C, West Turkana, Kenya. Journal of Human Evolution 48, 435472.Google Scholar
Delagnes, A., Boisserie, J.R., Beyene, Y., et al. (2011). Archaeological investigations in the Lower Omo Valley (Shungura Formation, Ethiopia): new data and perspectives. Journal of Human Evolution 61(2), 215222.Google Scholar
Delfino, M. (2020). Early Pliocene anuran fossils from Kanapoi, Kenya, and the first fossil record for the African burrowing frog Hemisus (Neobatrachia: Hemisotidae). Journal of Human Evolution 140, 102353.Google Scholar
Delfino, M., Segid, A., Yosief, D., et al. (2004). Fossil reptiles from the Pleistocene Homo-bearing locality of Buia (Eritrea, Northern Danakil Depression). Rivista Italiana di Stratigrafia e Paleontologia 110(Supplement), 5160.Google Scholar
Delson, E. (1984). Cercopithecid biochronology of the African Plio-Pleistocene: correlation among eastern and southern hominid-bearing localities. Courier Forschungsinstitut Senckenberg 69, 199218.Google Scholar
Delson, E. (1988). Chronology of South African australopiths site units. In: Grine, F.E. (Ed.), Evolutionary History of the Robust Australopithecines. New York: Aldine de Gruyter, pp. 317325.Google Scholar
Delson, E. and Hoffstetter, R. (1993). Theropithecus from Ternifine, Algeria. In: Jablonski, N.G. (Ed.), Theropithecus: The Rise and Fall of a Primate Genus. Cambridge: Cambridge University Press, pp. 191208.Google Scholar
Delson, E., Eck, G.G., Leakey, M.G. and Jablonski, N.G. (1993). A partial catalogue of fossil remains of Theropithecus. In: Jablonski, N.G. (Ed.), Theropithecus: The Rise and Fall of a Primate Genus. Cambridge: Cambridge University Press, pp. 499525.Google Scholar
Dembo, M., Matzke, N.J., Mooers, A.O. and Collard, M. (2015). Bayesian analysis of a morphological supermatrix sheds light on controversial fossil hominin relationships. Proceedings of the Royal Society B 282, 20150943.Google Scholar
deMenocal, P.B. (1995). Plio-Pleistocene African climate. Science 270, 5359.Google Scholar
deMenocal, P.B. (2004). African climate change and faunal evolution during the Pliocene–Pleistocene. Earth and Planetary Science Letters 220, 324.Google Scholar
deMenocal, P.B. (2011). Climate and human evolution. Science 331, 540542.Google Scholar
deMenocal, P.B. and Brown, F.H. (1999). Pliocene tephra correlations between East African hominid localities, the Gulf of Aden, and the Arabian Sea. In: Agustí, J., Rook, L. and Andrews, P. (Eds.), Evolution of Neogene Terrestrial Ecosystems in Europe. Cambridge: Cambridge University Press, pp. 2354.Google Scholar
deMenocal, P., Ortiz, J., Guilderson, T., et al. (2000). Abrupt onset and termination of the African Humid Period: rapid climate responses to gradual insolation forcing. Quaternary Science Reviews 19, 347361.Google Scholar
Deocampo, D.M. (2002). Sedimentary structures generated by Hippopotamus amphibius in a lake-margin wetland, Ngorongoro Crater, Tanzania. PALAIOS 17, 212217.Google Scholar
Denys, C. (1987a). Fossil rodents (other than Pedetidae) from Laetoli. In: Leakey, M.D. and Harris, J.M. (Eds.), Laetoli: A Pliocene Site in Northern Tanzania. Oxford: Clarendon Press, pp. 118170.Google Scholar
Denys, C. (1987b). Micromammals from the west Natron Pleistocene deposits (Tanzania): biostratigraphy and paleoecology. Sciences géologiques. Bulletin 40, 185201.Google Scholar
Denys, C. (1990a). Deux nouvelles espèces d’Aethomys (Rodentia, Muridae) à Langebaanweg (Pliocène, Afrique du Sud): implications phylogénétiques et paléoécologiques. Annales de Paléontologie 76, 4169.Google Scholar
Denys, C. (1990b). The oldest Acomys (Rodentia, Muridae) from the Lower Pliocene of South Africa and the problem of its murid affinities. Palaeontographica Abteilung A, Palaozoologie-stratigraphie 210, 7991.Google Scholar
Denys, C. (1992). Présence de Saccostomus (Rodentia, Mammalia) a Olduvai Bed I (Tanzanie, Pléistocčne inférieur). Implications phylétiques et paléobiogéographiques. Geobios, 25(1), 145154.Google Scholar
Denys, C. (1997). Rodent faunal lists in karstic and open-air sites of Africa: an attempt to evaluate predation and fossilisation biases on paleodiversity [Listas faunísticas de roedores en yacimientos kársticos y al aire libre de África: un intento para evaluar los sesgos de predación y fosilización sobre la paleodiversidad]. Journal of Iberian Geology [Cuadernos de geología ibérica] 23, 7394.Google Scholar
Denys, C. (1999). Of mice and men. Evolution in East and South Africa during Plio-Pleistocene times. In: Bromage, T.G. and Schrenk, F. (Eds.), African Biogeography, Climate Change and Human Evolution. New York: Oxford University Press, pp.226252.Google Scholar
Denys, C. (2011). Rodents. In: Harrison, T. (Ed.), Paleontology and Geology of Laetoli: Human Evolution in Context. Vol. 2: Fossil Hominins and the Associated Fauna. Dordrecht: Springer, pp. 1553.Google Scholar
Denys, C., Patou, M. and Djemmali, N. (1984). Tighennif (Ternifinne, Algérie). Premiers résultats concernant l’origine de l’accumulation du matériel osseux de ce gisement Pléistocène Comptes Rendu de l’Académie des Sciences Paris II 299, 481486.Google Scholar
Denys, C., Geraads, D., Hublin, J.-J. and Tong, H. (1987). Méthode d’étude taphonomique des microvertébrés. Application au site pléistocène de Tighenif (Ternifine, Algérie). Archeozoologia 1, 5382.Google Scholar
Denys, C., Williams, C., Dauphin, Y., Andrews, P. and Fernandez-Jalvo, Y. (1996). Diagenetical changes in Pleistocene small mammal bones from Olduvai Bed I. Palaeogeography, Palaeoclimatology, Palaeoecology 126, 121134.Google Scholar
Denys, C., Chitaukali, W., Mfune, J.K., Combrexelle, M. and Cacciani, F. (1999). Diversity of small mammals in owl pellet assemblages of Karonga district, northern Malawi. Acta Zoologica Cracoviensia, 42, 393396.Google Scholar
DeSilva, J.M., Steininger, C.M. and Patel, B.A. (2013). Cercopithecoid primate postcranial fossils from Cooper’s D, South Africa. Geobios 46, 381394.Google Scholar
DeSilva, J.M., Gill, C.M., Prang, T.C., Bredella, M.A. and Alemseged, Z. (2018). A nearly complete foot from Dikika, Ethiopia and its implications for the ontogeny and function of Australopithecus afarensis. Science Advances 4, eaar7723.Google Scholar
Desio, A. (1935). Studi geologici sulla Cirenaica, sul deserto Libico, sulla Tripolitania, et sulla Fezzan orientali. In: Missione scientifica della reale Accademia d’Italia a Cufra 1. Roma: Reale Accademia d’Italia.Google Scholar
D’Huart, J., de Jong, Y.A. and Butynski, T. (2016). Phacochoerus aethiopicus. The IUCN Red List of Threatened Species 2016: e. T41767A44140316.Google Scholar
Dibble, H.L., Aldeias, V., Alvarez-Fernandez, E., et al. (2012). New excavations at the site of Contrebandiers Cave, Morocco. PaleoAnthropology 2012, 145201.Google Scholar
Dice, L.R. (1945). Measures of the amount of ecologic association between species. Ecology 26, 297302.Google Scholar
Diekmann, B. and Kuhn, G. (2002). Sedimentary record of the mid-Pleistocene climate transition in the southeastern South Atlantic (ODP Site 1090). Palaeogeography, Palaeoclimatology, Palaeoecology 182(3–4), 241258.Google Scholar
Dietl, G.P. and Flessa, K.W. (2009). Conservation paleobiology: using the past to manage the future. The Paleontological Society Papers, Vol. 15, Paleontological Society.Google Scholar
Dietrich, W.O. (1941). Die säugetierpaläontologischen Ergebnisse der Kohl-Larsen’schen Expedition 1937–1939 im nördlichen Deutsch-Ostafrika. Zentralblatt für Mineralogie, Geologie und Paläontologie, Abt. B, Stuttgart, 217223.Google Scholar
Dietrich, W.O. (1942a). Ältestquartäre Säugetiere aus der südlichen Serengeti, Deutsch-Ostafrika. Palaeontographica 94A, 43133.Google Scholar
Dietrich, W.O. (1942b). Zur Entwicklungsmechanik des Gebisses der afrikanischen Nashörner. Zentralblatt für Mineralogie, Geologie und Paläontologie, Abt. B, Stuttgart, 297300.Google Scholar
Dietrich, W.O. (1945). Nashornreste aus dem Quartär Deutsch-Ostafrikas. Palaeontographica 96A, 4690.Google Scholar
Dietrich, W.O. (1950). Fossile Antilopen und Rinder Äquatorialafrikas (Material der Kohl-Larsen’schen Expeditionen). Palaeontographica 99A, 162.Google Scholar
Dietrich, W.O. (1951). Daten zu den fossilen Elefanten Afrikas und Ursprung der Gattung Loxodonta. Neues Jahrbuch für Mineralogie, Geologie und Paläontologie, Stuttgart, 93, 325378.Google Scholar
DiMaggio, E.N., Campisano, C.J., Arrowsmith, J.R., et al. (2008). Correlation and stratigraphy of the BKT-2 volcanic complex in west-central Afar, Ethiopia. In: Quade, J. and Wynn, J.G. (Eds.), The Geology of Early Humans in the Horn of Africa. Boulder: Geological Society of America, pp. 163178.Google Scholar
DiMaggio, E.N., Campisano, C.J., Rowan, J., et al. (2015a). Late Pliocene fossiliferous sedimentary record and the environmental context of early Homo from Afar, Ethiopia. Science 347, 13551359.Google Scholar
DiMaggio, E.N., Arrowsmith, J.R., Campisano, C.J., et al. (2015b). Tephrostratigraphy and depositional environment of young (<2.94 Ma) Hadar Formation deposits at Ledi-Geraru, Afar, Ethiopia. Journal of African Earth Sciences 112, 234250.Google Scholar
Di Vincenzo, F., Rodriguez, L., Carretero, J.M., et al. (2015). The massive fossil humerus from the Oldowan horizon of Gombore I, Melka Kunture (Ethiopia, >1.39 Ma). Quaternary Science Reviews 122, 207221.Google Scholar
Dirks, P.H. and Berger, L.R. (2013). Hominin-bearing caves and landscape dynamics in the Cradle of Humankind, South Africa. Journal of African Earth Sciences 78, 109131.Google Scholar
Dirks, P.H., Kibii, J.M., Kuhn, B.F., et al. (2010). Geological setting and age of Australopithecus sediba from southern Africa. Science, 328(5975), 205208.Google Scholar
Dirks, P.H., Berger, L.R., Roberts, E.M., et al. (2015). Geological and taphonomic context for the new hominin species Homo naledi from the Dinaledi Chamber, South Africa. Elife, 4, e09561.Google Scholar
Dirks, P.H., Roberts, E.M., Hilbert-Wolf, H., et al. (2017). The age of Homo naledi and associated sediments in the Rising Star Cave, South Africa. Elife, 6, e24231.Google Scholar
Disotell, T.R. (2006). ‘Chumanzee’ evolution: the urge to diverge and merge. Genome Biology 7(11), 1.Google Scholar
Disotell, T.R. (2013). Genetic perspectives on ape and human evolution. In: Begun, D.R. (Ed.), A Companion to Paleoanthropology. Chichester: Wiley-Blackwell, pp. 290305.Google Scholar
Ditchfield, P. and Harrison, T. (2011). Sedimentology, lithostratigraphy and depositional history of the Laetoli area. In: Harrison, T. (Ed.), Paleontology and Geology of Laetoli: Human Evolution in Context. Vol. 1: Geology, Geochronology, Paleoecology, and Paleoenvironment. Dordrecht: Springer, pp. 4776.Google Scholar
Ditchfield, P., Hicks, J., Plummer, T., Bishop, L.C. and Potts, R. (1999). Current research on the Late Pliocene and Pleistocene deposits north of Homa Mountain, southwestern Kenya. Journal of Human Evolution 36(2), 123150.Google Scholar
Ditchfield, P.W., Kingston, J.D., Plummer, T., et al. (2006). Tooth enamel oxygen isotope analysis to investigate seasonality at the late Pliocene Oldowan site Kanjera South, Kenya. PaleoAnthropology 2006:A16.Google Scholar
Ditchfield, P.W., Whitfield, E., Vincent, T., et al. (2019). Geochronology and physical context of Oldowan site formation at Kanjera South, Kenya. Geological Magazine 156(7), 11901200.Google Scholar
Dixey, F. (1927). The Tertiary and post-Tertiary lacustrine sediments of the Nyasan Rift-Valley. Quarterly Journal of the Geological Society 83(1–5), 432442.Google Scholar
Djemmali, N.-E. (1985). L’industrie lithique acheuléenne du gisement de Tighennif (Ternifine), Algérie. Thèse doctorat 3ème cycle, Université P. et M. Curie, Paris.Google Scholar
Dobson, M. (2004). Freshwater crabs in Africa. Freshwater Forum 21, 326.Google Scholar
Dodson, P. (1973). The significance of small bones in paleoecological interpretation. Contributions to Geology 12, 1519.Google Scholar
Dodson, P. and Wexlar, D. (1979). Taphonomic investigations of owl pellets. Paleobiology 5, 275284.Google Scholar
Dollion, A.Y., Cornette, R., Tolley, K.A., et al. (2015). Morphometric analysis of chameleon fossil fragments from the Early Pliocene of South Africa: a new piece of the chamaeleonid history. The Science of Nature 102(1–2), 2.Google Scholar
Doman, J.H. (2017). The Paleontology and Paleoecology of the Late Miocene Mpesida Beds and Lukeino Formation, Tugen Hills succession, Baringo, Kenya, Anthropology. Thesis, Department of Anthropology, Yale University, p. 202.Google Scholar
Doman, J. and Coutros, P. (2015). Environmental heterogeneity of a late Miocene east African landscape: introducing new mammalian fauna and integrating multiple paleoecological methods and modern forest ecology techniques in the Mpesida Beds, Baringo, Kenya. Journal of Vertebrate Paleontology, SVP Program and Abstracts Book (75).Google Scholar
Domínguez-Rodrigo, M. (1997). Meat-eating by early hominids at the FLK 22 Zinjanthropus site, Olduvai Gorge (Tanzania): an experimental approach using cut-mark data. Journal of Human Evolution 33, 669690.Google Scholar
Dominguez-Rodrigo, M.C. (2014). Is the ‘Savanna Hypothesis’ a dead concept for explaining the emergence of the earliest hominins? Current Anthropology 55(1), 5981.Google Scholar
Dominguez-Rodrigo, M. and Piqueras, A. (2003). The use of tooth pits to identify carnivore taxa in tooth-marked archaeofaunas and their relevance to reconstruct hominid carcass processing behaviours. Journal of Archaeological Science 30, 13851391.Google Scholar
Domínguez-Rodrigo, M., de la Torre, I., de Luque, L., et al. (2002). The ST site complex at Peninj, west Lake Natron, Tanzania: implications for early hominid behavioural models. Journal of Archaeological Science 29, 639665.Google Scholar
Domínguez-Rodrigo, M., Barba, R. and Egeland, C.P. (2007). Deconstructing Olduvai: A Taphonomic Study of the Bed I Sites. Dordrecht: Springer.Google Scholar
Domínguez-Rodrigo, M., Mabulla, A., Luque, L., et al. (2008). A new archaic Homo sapiens fossil from Lake Eyasi, Tanzania. Journal of Human Evolution 54, 899903.Google Scholar
Domínguez-Rodrigo, M., de Juana, S., Gala, A.B. and Rodríguez, M. (2009a). A new protocol to differentiate trampling marks from butchery cut marks. Journal of Archaeological Science 36, 26432654.Google Scholar
Domínguez-Rodrigo, M., Alcalá, L. and Luque, L. (Eds.) (2009b). Peninj: A Research Project on Human Origins 1995–2005. Oxford: Oxbow Books.Google Scholar
Domínguez-Rodrigo, M., Bunn, H.T., Mabulla, A., Baquedano, E. and Pickering, T.R. (2010a). Paleoecology and hominin behavior during Bed I at Olduvai Gorge (Tanzania). Quaternary Research 74, 301303.Google Scholar
Domínguez-Rodrigo, M., Pickering, T.R., Bunn, H.T. (2010b). Configurational approach to identifying the earliest hominin butchers. Proceedings of the National Academy of Sciences 107, 2092920934.Google Scholar
Domínguez-Rodrigo, M., Bunn, H.T., Pickering, T.R., et al. (2012). Autochthony and orientation patterns in Olduvai Bed I: a re-examination of the status of post-depositional biasing of archaeological assemblages from FLK North (FLKN). Journal of Archaeological Science 39, 21162127.Google Scholar
Dominguez-Rodrigo, M., Pickering, T. R., Baquedano, E., et al. (2013). First partial skeleton of a 1.34-million-year-old Paranthropus boisei from Bed II, Olduvai Gorge, Tanzania. PLoS ONE 8(12), e80347.Google Scholar
Domínguez-Rodrigo, M., Uribelarrea, D., Santonja, M., et al. (2014a). Autochthonous anisotropy of archaeological materials by the action of water: experimental and archaeological reassessment of the orientation patterns at the Olduvai sites. Journal of Archaeological Science 41, 4468.Google Scholar
Domínguez-Rodrigo, M., Diez-Martín, F., Mabulla, A., et al. (2014b). The evolution of hominin behavior during the Oldowan–Acheulean transition: recent evidence from Olduvai Gorge and Peninj (Tanzania). Quaternary International, 322, 16.Google Scholar
Domínguez-Rodrigo, M., Baquedano, E., Mabulla, A., Mercader, J. and Egeland, C.P. (2017). Paleoecological reconstructions of the Bed I and Bed II lacustrine basins of Olduvai Gorge (Tanzania) and insights into early human behavior. Palaeogeography, Palaeoclimatology, Palaeoecology 488, 18.Google Scholar
Dongmann, G., Nurnberg, H.W., Forstel, H. and Wagener, K. (1974). On the enrichment of H2 18O in the leaves of transpiring plants. Radiation and Environmental Biophysics 11, 4152.Google Scholar
Dorale, J.A., Edwards, R.L., Alexander, C.A. Jr, et al. (2004). Uranium-series dating of speleothems: current techniques, limits, & applications. In: Sasowsky, I.D. and Mylroie, J.E. (Eds)., Studies of Cave Sediments: Physical and Chemical Records of Paleoclimate. New York: Kluwer Academic/Plenum Publishers, pp. 177197.Google Scholar
Doran, T.L., Herries, A.I., Hopley, P.J., et al. (2015). Assessing the paleoenvironmental potential of Pliocene to Holocene tufa deposits along the Ghaap Plateau escarpment (South Africa) using stable isotopes. Quaternary Research 84(1), 133143.Google Scholar
Douady, C.J., Catzeflis, F., Raman, J., Springer, M.S. and Stanhope, M. J. (2003). The Sahara as a vicariant agent, and the role of Miocene climatic events, in the diversification of the mammalian order Macroscelidea (elephant shrews). Proceedings of the National Academy of Sciences, 100(14), 83258330.Google Scholar
Drake, R. and Curtis, G.H. (1987). K–Ar geochronology of the Laetoli fossil localities. In: Leakey, M.D. and Harris, J.M. (Eds.), Laetoli: A Pliocene Site in Northern Tanzania. Oxford: Clarendon Press, pp. 4852.Google Scholar
Drapeau, M.S.M., Bobe, R., Wynn, J.G., et al. (2014). The Omo Mursi Formation: a window into the East African Pliocene. Journal of Human Evolution, 75, 6479. doi: 10.1016/j.jhevol.2014.07.001.Google Scholar
Drapeau, M.S.M, Wynn, J.G., Geraads, D., et al. (2016). Where they were not: What can the Mursi Formation tell us about early hominins habitat preferences? Program of the 44th annual meeting of the Canadian Association of Physical Anthropology, 27. https://caba-acab.net/sites/default/files/basic-page/capa_2016_program_final.pdfGoogle Scholar
Dreyer, T.F. (1938). The archaeology of the Florisbad deposits. Argeologiese Navorsing van die Nasionale Museum, Bloemfontein 1, 183190.Google Scholar
Du, A. and Alemseged, Z. (2018). Diversity analysis of Plio-Pleistocene large mammal communities in the Omo-Turkana Basin, eastern Africa. Journal of Human Evolution 124, 2539.Google Scholar
Du, A., Robinson, J.R., Rowan, J., Lazagabaster, I.A. and Behrensmeyer, A.K. (2019). Stable carbon isotopes from paleosol carbonate and herbivore enamel document differing paleovegetation signals in the eastern African Plio-Pleistocene. Review of Palaeobotany and Palynology 261, 4152.Google Scholar
Du, A., Rowan, J., Wang, S.C., Wood, B.A. and Alemseged, Z. (2020). Statistical estimates of hominin origination and extinction dates: a case study examining the Australopithecus anamensis–afarensis lineage. Journal of Human Evolution 138, 102688.Google Scholar
Dubois, C., Quinif, Y., Baele, J.M., et al. (2014). The process of ghost-rock karstification and its role in the formation of cave systems. Earth-Science Reviews 131, 116148.Google Scholar
Du Toit, J.T. and Owen-Smith, N. (1989). Body size, population metabolism, and habitat specialization among large African herbivores. The American Naturalist, 133(5), 736740.Google Scholar
Dumouchel, L. (2017). Australopithecus anamensis paleoenvironments in the Omo-Turkana basin using stable isotope analysis of tooth enamel from mixed feeding taxa. PaleoAnthropology, A10.Google Scholar
Dumouchel, L. (2018). A multi-proxi analysis of Australopithecus anamensis paleoecology in the Omo-Turkana Basin. PhD dissertation, George Washington University, p. 171.Google Scholar
Dunbar, R.I.M. (1983). Theropithecines and hominids: contrasting solutions to the same ecological problem. Journal of Human Evolution 12, 647658.Google Scholar
Dunhill, A.M., Benton, M.J., Twitchett, R.J. and Newell, A.J. (2012). Completeness of the fossil record and validity of sampling proxies at outcrop level. Palaeontology 55(6), 11551175.Google Scholar
Dunhill, A.M., Hannisdal, B. and Benton, M. J. (2014). Disentangling rock record bias and common-cause from redundancy in the British fossil record. Nature Communications 5, 4818.Google Scholar
Dupont, L. and Leroy, S. (1995). Steps toward drier climatic conditions in Northwestern Africa during the Upper Pliocene. In: Vrba, E.S., Denton, G., Partridge, T. and Burckle, L. (Eds.), Paleoclimate and Evolution. New Haven: Yale University Press, pp. 289298.Google Scholar
Dupont-Nivet, G., Sier, M., Campisano, C.J., et al. (2008). Magnetostratigraphy of the eastern Hadar Basin (Ledi-Geraru research area, Ethiopia) and implications for hominin paleoenvironments. In: Quade, J. and Wynn, J.G. (Eds.), The Geology of Early Humans in the Horn of Africa. Boulder: The Geological Society of America, pp. 6785.Google Scholar
Duringer, P., Schuster, M., Genise, J.F., et al. (2006). The first fossil fungus gardens of Isoptera: oldest evidence of symbiotic termite fungiculture (Miocene, Chad basin). Naturwissenschaften 93, 610615.Google Scholar
Duringer, P., Schuster, M., Genise, J.F., et al. (2007). New termite trace fossils: galleries, nests and fungus combs from the Chad basin of Africa (Upper Miocene–Lower Pliocene). Palaeogeography, Palaeoclimatology, Palaeoecology 251, 323353.Google Scholar
Durkee, H. and Brown, F. H. (2014). Correlation of volcanic ash layers between the Early Pleistocene Acheulean sites of Isinya, Kariandusi, and Olorgesailie, Kenya. Journal of Archaeological Science 49, 510517.Google Scholar
Dusseldorp, G., Lombard, M. and Wurz, S. (2013). Pleistocene Homo and the updated Stone Age sequence of South Africa. South African Journal of Science 109(5/6), Art.#0042.Google Scholar
Duvernoy, G.L. (1851). Note sur une espèce de buffle fossile (Bubalus (Arni) antiquus), découverte en Algérie. Comptes Rendus de l’Académie des Sciences 33, 595597.Google Scholar
Dynesius, M. and Jansson, R. (2000). Evolutionary consequences of changes in species’ geographical distributions driven by Milankovitch climate oscillations. Proceedings of the National Academy of Sciences of the United States of America 97(16), 91159120.Google Scholar
Eberz, G.W., Williams, F.M. and Williams, M.A.J. (1988). Plio-Pleistocene volcanism and sedimentary facies changes at Gadeb prehistoric site, Ethiopia. Geologische Rundschau 77, 513527.Google Scholar
Ebinger, C.J., 1989. Tectonic development of the western branch of the East African rift system. Geological Society of America Bulletin 101, 885903.Google Scholar
Ebinger, C. and Scholz, C.A. (2012). Continental rift basins: the East African perspective. In: Busby, C. and Azor, A. (Eds.), Tectonics of Sedimentary Basins: Recent Advances. Oxford: Wiley-Blackwell, pp. 183208.Google Scholar
Ebinger, C.J., Deino, A.L., Drake, R.E. and Tesha, A.L. (1989). Chronology of volcanism and rift basin propagation: Rungwe volcanic province, East Africa. Journal of Geophysical Research 94, 1578515803.Google Scholar
Ebinger, C.J., Deino, A.L., Tesha, A.L. and Ring, U. (1993). Regional tectonic controls on rift basin geometry: evolution of an East African lake basin. Journal of Geophysical Research B98, 1782117836.Google Scholar
Ebinger, C.J., Yemane, T., Harding, D.J., et al. (2000). Rift deflection, migration, and propagation: linkage of the Ethiopian and Eastern rifts, Africa. Geological Society of America Bulletin 112, 163176.Google Scholar
Echassoux, A., Moullé, P.-É., Desclaux, E. and Alemseged, Z. (2004). Les faunes Plio-Pléistocène du site de Fejej FJ1. In: de Lumley, H. and Beyene, Y. (Eds.), Les sites préhistoriques de la région de Fejej, Sud-Omo, Éthiopie, dans leur contexte stratigraphique et paléontologique. Paris: Éditions Recherche sur les Civilisations, pp. 203340.Google Scholar
Eck, G.G. (1993). Theropithecus darti from the Hadar Formation, Ethiopia. In: Jablonski, N.G. (Ed.), Theropithecus: The Rise and Fall of a Primate Genus. Cambridge: Cambridge University Press, pp. 1583.Google Scholar
Eck, G.G. (2007). The effects of collection strategy and effort on faunal recovery. A case study of the American and French collections from the Shungura Formation, Ethiopia. In: Behrensmeyer, A.K., Bobe, R. and Alemseged, Z. (Eds.), Hominin Environments in the East African Pliocene: An Assessment of the Faunal Evidence. Dordrecht: Springer, pp. 183215.Google Scholar
Ecker, M. (2016). Two million years of environmental change: a case study from Wonderwerk Cave, Northern Cape, South Africa. DPhil dissertation, University of Oxford.Google Scholar
Ecker, M. and Lee-Thorp, J.A. (2018). The dietary ecology of the extinct springbok Antidorcas bondi. Quaternary international 495, 136143.Google Scholar
Ecker, M., Botha-Brink, J., Lee-Thorp, J., Piuz, A. and Horwitz, L. (2015a). Ostrich eggshell as a source of palaeoenvironmental information in the arid interior of South Africa: a case study from Wonderwerk Cave. In Runge, J. (Ed.), Changing Climates, Ecosystems and Environments within Arid Southern Africa and Adjoining Regions. Palaeoecology of Africa, 33. Boca Raton: CRC Press, pp. 95115.Google Scholar
Ecker, M., Brink, J., Scott, L., et al. (2015b). New isotopic insights into palaeoecology and palaeoenvironment over 2 million years from Wonderwerk Cave, South Africa. Poster Session 2, Poster 75, 5th Annual meeting of ESHE, London, 10–12 September. Abstract p. 84.Google Scholar
Ecker, M., Brink, J., Horwitz, L.K., Scott, L. and Lee-Thorp, J.A. (2018a). A 12,000 year record of changes in herbivore niche separation and palaeoclimate (Wonderwerk Cave, South Africa). Quaternary Science Reviews 180, 132144.Google Scholar
Ecker, M., Brink, J.S., Rossouw, L., et al. (2018b). The palaeoecological context of the Oldowan–Acheulean in southern Africa. Nature Ecology & Evolution 2(7), 10801086.Google Scholar
Edwards, E.J., Osborne, C.P., Strömberg, C.A., et al. (2010). The origins of C4 grasslands: integrating evolutionary and ecosystem science. Science 328, 587591.Google Scholar
Efremov, J.A. (1940). Taphonomy: new branch of paleontology. Pan-American Geologist 74, 8193.Google Scholar
Egeland, C.P. (2007). Zooarchaeology and taphonomy of the DK site. In: Domínguez-Rodrigo, M., Barba, R. and Egeland, C. (Eds.), Deconstructing Olduvai: A Taphonomic Study of the Bed I Sites. Dordrecht: Springer, pp. 253268.Google Scholar
Ehleringer, J.R., Cerling, T.E. and Helliker, B.R. (1997). C4 photosynthesis, atmospheric CO2, and climate. Oecologia 112(3), 285299.Google Scholar
Eisenhart, W.L. (1974). The fossil Cercopithecoids of Makapansgat and Sterkfontein. BA thesis, Harvard College.Google Scholar
Eisenmann, V. (1976a). Le protostylide: Valeur systematique et signification phylétique chez les espèces actuelles et fossiles du genre Equus (Perissodactyla, Mammalia). Zeitschrift für Säugetierkunde 41, 349365.Google Scholar
Eisenmann, V. (1976b). Equidae from the Shungura Formation. In: Coppens, Y., Howell, F.C., Isaac, G.L., and Leakey, R.E., R.E. (Eds.), Earliest Man and Environments in the Lake Rudolf Basin. Chicago: University of Chicago Press, pp. 225233.Google Scholar
Eisenmann, V. (1979a). Etude des cornets des dents incisives inférieures des Equus (Mammalia, Perissodactyla) actuels et fossiles. Palaeontographia Italica 71, 5575.Google Scholar
Eisenmann, V. (1979b). Les métapodes d’ Equus sensu lato (Mammalia, Perissodactyla). Geobios 12, 863886.Google Scholar
Eisenmann, V. (1980). Les chevaux (Equus sensu lato) fossiles et actuels: crânes et dents jugales supérieures. Cahiers de paléontologie. Paris: Centre national de la recherche scientifique.Google Scholar
Eisenmann, V. (1981). Etude des dents jugales inférieures des Equus (Mammalia, Perissodactyla) actuels et fossiles. Palaeovertebrata 10, 127226.Google Scholar
Eisenmann, V. (1984). Sur quelques caractères adaptatifs du squelette d’Equus (Mammalia, Perissodactyla) et leurs implications paléoécologiques. Bulletin du Museum national d’histiore naturelle, C, 4ème séres 6, 185195.Google Scholar
Eisenmann, V. (1985). Indications paléoécologiques fournies par les Equus (Mammalia, Perissodactyla) plio-pléistocènes d’Afrique. In: L’Environnement des Hominidés au Plio-Pléistocène. Paris: Masson, pp. 5779.Google Scholar
Eisenmann, V. (1994). Equidae of the Albertine Rift Valley, Uganda. In: Senut, B. and Pickford, M. (Eds.), Geology and Palaeobiology of the Albertine Rift Valley, Uganda–Zaire, Vol. II, Palaeobiology. Publication Occasionelle. Orléans: Centre International pour la Formation et les Echanges Géologiques – CIFEG, pp. 289307.Google Scholar
Eisenmann, V. (1998). Folivores et tondeurs d’herbe: forme de la symphyse mandibulaire des Equidés et des Tapiridés (Perissodactyla, Mammalia). Geobios 31, 113123.Google Scholar
Eisenmann, V. and Geraads, D. (2007). The hipparion from the late Pliocene of Ahl al Oughlam, Morocco, and a revision of the relationships of Pliocene and Pleistocene African hipparions. Palaeontologica Africana 42, 5198.Google Scholar
Eizirik, E., Murphy, W.J., Springer, M.S. and O’Brien, S J. (2004). Molecular phylogeny and dating of early primate divergences. In: Ross, C.F. and Kay, R.F. (Eds.), Anthropoid Origins. Developments in Primatology: Progress and Prospects. Boston, MA: Springer, pp. 4564.Google Scholar
El-Saadawi, W., Kamal-El-Din, M., Wheeler, E.A., et al. (2014). Early Miocene woods of Egypt. IAWA Journal 35, 3550.Google Scholar
El-Shawaihdi, M.H., Mozley, P.S., Boaz, N.T., et al. (2016). Stratigraphy of the Neogene Sahabi units in the Sirt Basin, northeast Libya. Journal of African Earth Sciences 118, 87106.Google Scholar
Eldredge, N. and Gould, S.J. (1972). Punctuated equilibria: an alternative to phyletic gradualism. In: Schopf, T.J.M. (Ed.), Models in Paleobiology. San Francisco: Freeman, Cooper and Company, pp. 82115.Google Scholar
Ellis, R.P. (1987). A review of comparative leaf blade anatomy in the systematics of the Poaceae: the past twenty five years. In: Soderstrom, T.R., Hilu, K.W., Campbell, C.S., and Barkworth, M.E. (Eds.), Grass Systematics and Evolution. Washington, DC: Smithsonian Institution Press, pp. 310.Google Scholar
Ellison, A.M. (2010). Partitioning diversity 1. Ecology 91(7), 19621963.Google Scholar
Elton, S. (2001). Locomotor and habitat classification of cercopithecoid postcranial material from Sterkfontein Member 4, Bolt’s Farm, and Swartkrans Member 1 and 2, South Africa. Palaeontologica Africana 37, 115126.Google Scholar
Elton, S. (2006). Forty years on and still going strong: the use of hominin–cercopithecid comparisons in palaeoanthropology. Journal of the Royal Anthropological Institute 12, 1938.Google Scholar
Engel, J., Woodhead, J., Hellstrom, J., et al. (2019). Corrections for initial isotopic disequilibrium in the speleothem U–Pb dating method. Quaternary Geochronology 54, 101009.Google Scholar
Ennouchi, E. (1962). Un Neanderthalien: l’homme du Jebel Irhoud (Maroc). L’Anthropologie 66, 279299.Google Scholar
Ennouchi, E. (1968). Le deuxième crâne de l’homme d’Irhoud. Annales de Paléontologie (Vertébrés) 54, 117128.Google Scholar
Ennouchi, E. (1969). Découverte d’un pithécanthropien au Maroc. Comptes Rendus de l’Académie des Sciences D 269, 763765.Google Scholar
Ennouchi, E. (1970). Un nouvel archanthropien au Maroc. Annales de Paléontologie 56, 95107.Google Scholar
Ennouchi, E. (1972). Nouvelle découverte d’un archanthropien au Maroc. Comptes Rendus de l’Académie des Sciences Paris D 274, 30883090.Google Scholar
Ennouchi, E. (1975). New discovery of an Archanthropian in Morocco. Journal of Human Evolution 4, 441443.Google Scholar
Ennouchi, E. (1976). Un deuxième archanthropien à la carrière Thomas III (Maroc). Bulletin du Museum national d’Histouire Naturelle., sciences des Terre, séries 3 56, 273296.Google Scholar
Erlanger, E.D., Granger, D.E., Gibbon, R.J. (2012). Rock uplift rates in South Africa from isochron burial dating of fluvial and marine terraces. Geology 40, 10191022.Google Scholar
Eriksson, P.G., Schweitzer, J.K., Bosch, P.J.A., et al. (1993). The Transvaal sequence: an overview. Journal of African Earth Science 16, 2551.Google Scholar
Eronen, J.T., Ataabadi, M.M., Micheels, A., et al. (2009). Distribution history and climatic controls of the Late Miocene Pikermian chronofauna. Proceedings of the National Academy of Sciences 106(29), 1186711871.Google Scholar
Eronen, J.T., Puolamäki, P., Liu, L., et al. (2010a). Precipitation and large herbivorous mammals I: estimates from present-day communities. Evolutionary Ecology Research 12, 217233.Google Scholar
Eronen, J.T., Puolamäki, P., Liu, L., et al. (2010b). Precipitation and large herbivorous mammals II: application to fossil data. Evolutionary Ecology Research 12, 235248.Google Scholar
Esau, K. (1960). Anatomy of Seed Plants. New York: John Wiley and Sons.Google Scholar
Esterhuysen, A.B., Sanders, V.M. and Smith, J.M. (2008). Human skeletal and mummified remains from the AD1854 siege of Mugombane, Limpopo, South Africa. Journal of Archaeological Sciences 36, 10381049.Google Scholar
Estes, R.D. (2012). The Behavior Guide to African Mammals: Including Hoofed Mammals, Carnivores, Primates. Berkeley: University of California Press.Google Scholar
Eugster, H.P. (1981). Lake Magadi, Kenya, and its precursors. In: Nissenbaum, A. (Ed.), Hypersaline Brines and Evaporitic Environments. Amsterdam: Elsevier, pp. 195232.Google Scholar
Evans, E.M.N., van Couvering, J.H. and Andrews, P. (1981). Palaeoecology of Miocene sites in Western Kenya. Journal of Human Evolution 10, 3548.Google Scholar
Everett, M. (2010). The paleoecology of the Pleistocene Upper Busidima Formation, Gona, Afar Depression, Ethiopia. PhD thesis, University of Indiana.Google Scholar
Ewer, R.F. (1956). The dating of the Australopithecinae: faunal evidence. South African Archaeological Bulletin 11, 4145.Google Scholar
Ewer, R.F. (1957a). Faunal evidence on the dating of the Australopithecinae. In: Clark, J.D. and Cole, S. (Eds.), Proceedings of the Third Pan-African Congress on Prehistory, 1955. London: Livingstone, pp. 135142.Google Scholar
Ewer, R.F. (1957b). Some fossil carnivores from Makapansgat valley. Palaeontologica Africana 4, 5767.Google Scholar
Ewer, R.F. (1958). The fossil Suidae of Makapansgat. Proceedings of the Zoological Society of London 1303, 329372.Google Scholar
Ewer, R.F. (1967). The fossil hyaenids of Africa – a reappraisal. In Bishop, W.W. and Clard, J.D. (Eds.), Background to Evolution in Africa. Chicago: University of Chicago Press, pp. 109123.Google Scholar
Faith, J.T. (2011). Ungulate community richness, grazer extinctions, and human subsistence behavior in southern Africa’s Cape Floral Region. Palaeogeography, Palaeoclimatology, Palaeoecology 306, 219227.Google Scholar
Faith, J.T. (2012a). Conservation implications of fossil roan antelope (Hippotragus equinus) in southern Africa’s Cape Floristic Region. In: Louys, J. (Ed.), Paleontology in Ecology and Conservation. Heidelberg: Springer, pp. 239251.Google Scholar
Faith, J.T. (2012b). Palaeozoological insights into management options for a threatened mammal: southern Africa’s Cape mountain zebra (Equus zebra zebra). Diversity and Distributions 18, 438447.Google Scholar
Faith, J.T. (2013). Taphonomic and paleoecological change in the large mammal sequence from Boomplaas Cave, Western Cape, South Africa. Journal of Human Evolution 65, 715730.Google Scholar
Faith, J.T. (2014). Late Pleistocene and Holocene mammal extinctions on continental Africa. Earth-Science Reviews 128, 105121.Google Scholar
Faith, J.T. and Behrensmeyer, A.K. (2013). Climate change and faunal turnover: testing the mechanics of the turnover–pulse hypothesis with South African fossil data. Paleobiology 39, 609627.Google Scholar
Faith, J.T. and O’Connell, J. F. (2011). Revisiting the late Pleistocene mammal extinction record at Tight Entrance Cave, southwestern Australia. Quaternary Research 76(3), 397400.Google Scholar
Faith, J.T. and Thompson, J.C. (2013). Fossil evidence for seasonal calving and migration of extinct blue antelope (Hippotragus leucophaeus) in southern Africa. Journal of Biogeography 40(11), 21082118.Google Scholar
Faith, J.T., Choiniere, J.N., Tryon, C.A., Peppe, D.J. and Fox, D.L. (2011). Taxonomic status and paleoecology of Rusingoryx atopocranion (Mammalia, Artiodactyla), an extinct Pleistocene bovid from Rusinga Island, Kenya. Quaternary Research 75, 697707.Google Scholar
Faith, J.T., Potts, R., Plummer, T.W., et al. (2012). New perspectives on middle Pleistocene change in the large mammal faunas of East Africa: Damaliscus hypsodon sp. nov. (Mammalia, Artiodactyla) from Lainyamok, Kenya. Palaeogeography, Palaeoclimatology, Palaeoecology 361–362, 8493.Google Scholar
Faith, J.T., Tryon, C.A., Peppe, D.J. and Fox, D.L. (2013). The fossil history of Grévy’s zebra (Equus grevyi) in equatorial East Africa. Journal of Biogeography 40, 359369.Google Scholar
Faith, J.T., Tryon, C.A., Peppe, D.J., et al. (2015). Paleoenvironmental context of the Middle Stone Age record from Karungu, Lake Victoria Basin, Kenya, and its implications for human and fauna dispersals in East Africa. Journal of Human Evolution 83, 2845.Google Scholar
Faith, J.T., Tryon, C.A. and Peppe, D.J. (2016). Environmental change, ungulate biogeography, and their implications for early human dispersals in equatorial East Africa. In: Jones, S.C. and Stewart, B.A. (Eds.), Africa from MIS 6–2: Population Dynamics and Paleoenvironments. Dordrecht: Springer, pp. 233245.Google Scholar
Faith, J.T., Rowan, J., Du, A. and Koch, P.L. (2018). Plio-Pleistocene decline of African megaherbivores: no evidence for ancient hominin impacts. Science 362, 938.Google Scholar
Faith, J.T., Du, A. and Rowan, J. (2019a). Addressing the effects of sampling on ecometric-based paleoenvironmental reconstructions. Palaeogeography, Palaeoclimatology, Palaeoecology 528, 175185.Google Scholar
Faith, J.T., Chase, B.M. and Avery, D.M. (2019b). Late Quaternary micromammals and the precipitation history of the southern Cape, South Africa. Quaternary Research 91(2), 848860.Google Scholar
Fara, E., Likius, A., Mackaye, H.T., Vignaud, P. and Brunet, M. (2005). Pliocene large-mammal assemblages from northern Chad: sampling and ecological structure. Naturwissenschaften 92, 537541.Google Scholar
Fauquette, S., Suc, J.-P., Bertini, A., et al. (2006). How much did climate force the Messinian salinity crisis? Quantified climatic conditions from pollen records in the Mediterranean region. Palaeogeography, Palaeoclimatology, Palaeoecology 238, 281301.Google Scholar
Feakins, S.J. (2013). Pollen-corrected leaf wax D/H reconstructions of northeast African hydrological changes during the late Miocene. Palaeogeography, Palaeoclimatology, Palaeoecology 374, 6271.Google Scholar
Feakins, S.J. and deMenocal, P.B. (2010). Global and African regional climate during the Cenozoic. In: Werdelin, L. and Sanders, W.J. (Eds.), Cenozoic Mammals of Africa. Berkeley: University of California Press, pp. 4555.Google Scholar
Feakins, S.J., deMenocal, P.B. and Eglinton, T.J. (2005). Biomarker records of late Neogene changes in northeast African vegetation. Geology 33, 977980.Google Scholar
Feakins, S.J., Levin, N.E., Liddy, H.M., et al. (2013). Northeast African vegetation change over 12 m.y. Geology 41(3), 295298.Google Scholar
Feathers, J.K. (2002). Luminescence dating in less than ideal conditions: case studies from Klasies River main site and Duinefontein, South Africa. Journal of Archaeological Science 29(2), 177194.Google Scholar
Feathers, J.K. and Bush, D.A. (2000). Luminescence dating of middle stone age deposits at Die Kelders. Journal of Human Evolution 38(1), 91119.Google Scholar
Feddi, N., Fauquette, S. and Suc, J.-P. (2011). Histoire plio-pléistocène des écosystèmes végétaux de Méditerranée sud-occidentale: apport de l’analyse pollinique de deux sondages en mer d’Alboran. Geobios 44, 5769.Google Scholar
Fedorov, A., Dekens, P., McCarthy, M., et al. (2006) The Pliocene paradox (mechanisms for a permanent El Niño). Science 312(5779), 14851489.Google Scholar
Fedorov, A.V., Brierley, C.M., Lawrence, K.T., et al. (2013). Patterns and mechanisms of early Pliocene warmth. Nature 496, 4349.Google Scholar
Feibel, C.S. (1988). Paleoenvironments of the Koobi Fora Formation, northern Kenya. Thesis. University of Utah, p. 330.Google Scholar
Feibel, C.S. (1994). Freshwater stingrays from the Plio-Pleistocene of the Turkana Basin, Kenya and Ethiopia. Lethaia 26, 359366.Google Scholar
Feibel, C.S. (1999a). Basin evolution, sedimentary dynamics, and hominid habitats in East Africa: an ecosystem approach. In: Bromage, T.G. and Schrenk, F. (Eds.), African Biogeography, Climate Change, and Human Evolution. Oxford: Oxford University Press, pp. 276281.Google Scholar
Feibel, C.S. (1999b). Tephrostratigraphy and geological context in paleoanthropology. Evolutionary Anthropology 8, 87100.Google Scholar
Feibel, C.S. (2003). Stratigraphy and depositional setting of the Pliocene Kanapoi Formation, lower Kerio valley, Kenya. In: Harris, J.M. and Leakey, M.G. (Eds.), Geology and Vertebrate Paleontology of the Early Pliocene site of Kanapoi, Northern Kenya. Los Angeles: Natural History Museum of Los Angeles County, pp. 920.Google Scholar
Feibel, C. (2011). A geological history of the Turkana Basin. Evolutionary Anthropology: Issues, News, and Reviews 20(6), 206216.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(4), 595622.Google Scholar
Feibel, C.S., Harris, J.M. and Brown, F.H. (1991). Palaeoenvironmental context for the late Neogene of the Turkana Basin. In: Harris, J.M. (Ed.), Koobi Fora Research Project, Volume 3: The Fossil Ungulates: Geology, Fossil Artiodactyls, and Paleoenvironments. Oxford: Clarendon Press, pp. 321370.Google Scholar
Feibel, C.S., Agnew, N., Latimer, B., et al. (1996). The Laetoli hominid footprints – a preliminary report on the conservation and scientific restudy. Evolutionary Anthropology 4, 149154.Google Scholar
Feibel, C.S., Lepre, C.J. and Quinn, R.L. (2009). Stratigraphy, correlation, and age estimates for fossils from Area 123, Koobi Fora. Journal of Human Evolution 57(2), 112122.Google Scholar
Fernandez, P., Bouzouggar, A., Collina-Girard, J. and Coulon, M. (2015). The last occurrence of Megaceroides algericus Lyddekker, 1890 (Mammalia, Cervidae) during the middle Holocene in the cave of Bizmoune (Morocco, Essaouira region). Quaternary International 374, 154167.Google Scholar
Fernández-Jalvo, Y. and Andrews, P. (2003). Experimental effects of water abrasion on bone fragments. Journal of Taphonomy 1, 147163.Google Scholar
Fernández-Jalvo, Y. and Andrews, P. (2016). Atlas of Taphonomic Identifications. Dordrecht: SpringerScience and Business.Google Scholar
Fernández-Jalvo, Y. and Avery, D.M. (2015). Pleistocene micromammals and their predators at Wonderwerk Cave, South Africa. African Archaeological Review 32, 751791.Google Scholar
Fernandez-Jalvo, Y., Denys, C., Andrews, P., et al. (1998). Taphonomy and palaeoecology of Olduvai Bed-I (Pleistocene, Tanzania). Journal of Human Evolution 34, 137172.Google Scholar
Fernández-Jalvo, Y., Tormoa, L., Andrews, P. and Marin-Monfort, M.D. (2018). Taphonomy of burnt bones from Wonderwerk Cave (South Africa). Quaternary International 495, 1929.Google Scholar
Ferraro, J.V., Plummer, T.W., Pobiner, B.L., et al. (2013). Earliest archaeological evidence of persistent hominin carnivory. PLoS ONE 8(4), e62174.Google Scholar
Ferretti, M.P., Torre, D. and Rook, L. (2001). The Stegotetrabelodon remains from Cessaniti (Calabria, Southern Italy) and their bearing on Late Miocene biogeography of the genus. In: Cavarretta, G. et al. (Eds.), Proceedings of the First International Congress of La Terra degli Elefanti, The World of Elephants. Rome: Consiglio Nazionale delle Ricerche, pp. 633636.Google Scholar
Ferretti, M.P., Ficcarelli, G., Libsekal, Y., Tecle, T.M. and Rook, L. (2003a). Fossil elephants from Buia (Northern Afar Depression, Eritrea) with remarks on the systematics of Elephas recki (Proboscidea, Elephantidae). Journal of Vertebrate Paleontology 23, 244257.Google Scholar
Ferretti, M.P., Rook, L. and Torre, D. (2003b). Stegotetrabelodon (Proboscidea, Elephantidae) from the Late Miocene of Southern Italy. Journal of Vertebrate Paleontology 23, 659666.Google Scholar
Fichtel, C., Perry, S. and Gros-Louis, J. (2005). Alarm calls of white-faced capuchin monkeys: an acoustic analysis. Animal Behaviour 70, 165176.Google Scholar
Field, D.J. (2020). Preliminary paleoecological insights from the Pliocene avifauna of Kanapoi, Kenya: Implications for the ecology of Australopithecus anamensis. Journal of Human Evolution 140, 102384.Google Scholar
Fiore, I. and Tagliacozzo, A. (2004). Taphonomic analysis of the bone remains from the Oldowan site of Garba IV. In: Chavaillon, J. and Piperno, M. (Eds.), Studies on the Early Paleolithic site of Melka Kunture, Ethiopia. Florence: Istituto Italiano di Preistoria e Protostoria, pp. 639682.Google Scholar
Fiorillo, A.R. (1988). Taphonomy of Hazard Homestead Quarry (Ogallala Group), Hitchcock County, Nebraska. Rocky Mountain Geology 26, 5797.Google Scholar
Fleagle, J.G., Rasmussen, D.T., Yirga, S., Bown, T.M. and Grine, F.E. (1991). New hominid fossils from Fejej, Southern Ethiopia. Journal of Human Evolution 21(2), 145152.Google Scholar
Flecker, R., Krijgsman, W., Capella, W., et al. (2015). Evolution of the Late Miocene Mediterranean–Atlantic gateways and their impact on regional and global environmental change. Earth-Science Reviews 150, 365392.Google Scholar
Flowers, R.M. and Schoene, B. (2010). (U–Th)/He thermochronometry constraints on unroofing of the eastern Kaapvaal craton and significance for uplift of the southern African Plateau. Geology 38, 827830.Google Scholar
Flügel, T.J., Eckardt, F.D. and Cotterill, F.P.D. (2015). The present day drainage patterns of the Congo River system and their Neogene evolution. In: de Wit, M.J., Guillocheau, F. and de Wit, M.C.J. (Eds.), Geology and Resource Potential of the Congo Basin. Heidelberg: Springer, pp. 315337.Google Scholar
Flynn, L.J. and Maurice, S. (1984). A muroid rodent of Asian affinity from the Miocene of Kenya. Journal of Vertebrate Paleontology 3, 160165.Google Scholar
Foley, R. (1993). African terrestrial primates: the comparative evolutionary biology of Theropithecus and the Hominidae. In: Jablonski, N.G. (Ed.), Theropithecus: The Rise and Fall of a Primate Genus. Cambridge: Cambridge University Press, pp. 245270.Google Scholar
Foley, R.A. (1994). Speciation, extinction and climatic change in hominid evolution. Journal of Human Evolution 26, 275289.Google Scholar
Foley, R. (2005). Species diversity in human evolution: challenges and opportunities. Transactions of the Royal Society of South Africa 60, 6772.Google Scholar
Folinsbee, K.E. and Reisz, R.R. (2013). New craniodental fossils of Papionin monkeys from Cooper’s D, South Africa. American Journal of Physical Anthropology 151, 613629.Google Scholar
Forbes, S.A. (1907). On the local distribution of certain Illinois fishes: an essay in statistical ecology. Bulletin of the Illinois State Laboratory of Natural History 7, 273303.Google Scholar
Forrest, F.L. (2017). Zooarchaeological and palaeoenvironmental reconstruction of newly excavated Middle Pleistocene deposits from Elandsfontein, South Africa. Unpublished doctoral dissertation, City University of New York.Google Scholar
Forstén, A. (1972). Hipparion primigenium from Southern Tunisia. Notes du Service géologique de Tunisie 35, 728.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
Fortelius, M., Andrews, P., Bernor, R., Viranta, S. and Werdelin, L. (1996). Preliminary analysis of taxonomic diversity, turnover and provinciality in a subsample of large land mammals from the later Miocene of western Eurasia. Acta Zoologica Cracoviensia, 39, 167178.Google Scholar
Fortelius, M., Eronen, J.T., Jernvall, J., et al. (2002). Fossil mammals resolve regional patterns of Eurasian climate change over 20 million years. Evolutionary Ecology Research 4, 10051016.Google Scholar
Fortelius, M., Žliobaitė, I., Kaya, F., et al. (2016). An ecometric analysis of the fossil mammal record of the Turkana Basin. Philosophical Transactions of the Royal Society of London B: Biological Sciences 371, 20150232.Google Scholar
Foster, A., Ebinger, C., Mbede, E., Rex, D., 1997. Tectonic development of the northern Tanzanian sector of the East African Rift System. Journal of the Geological Society of London 154, 689700.Google Scholar
Fourie, N.N., Lee-Thorp, J.A. and Ackermann, R.R. (2002). Biogeochemical and craniometric investigation of dietary ecology, niche separation, and taxonomy of Plio-Pleistocene Cercopithecoids from the Makapansgat Limeworks. American Journal of Physical Anthropology 135, 121135.Google Scholar
Fourtau, R. (1920). Contribution à l’étude des vertébrés miocènes de l’Egypte. Cairo: Ministry of Finance, Survey Dept, 121 pp.Google Scholar
Fourvel, J.-B. (2018). Civettictis braini nov. sp. (Mammalia: Carnivora), a new viverrid from the hominin-bearing site of Kromdraai (Gauteng, South Africa). Comptes Rendus Palevol 17(6), 366377.Google Scholar
Fourvel, J.-B., Brink, J., O’Regan, H., Beaudet, A., Pavia, M. (2016). Some preliminary interpretations of the oldest faunal assemblage from Kromdraai. In: Braga, J. and Thackeray, J.F. (Eds.), Kromdraai, a Birthplace of Paranthropus in the Cradle of Humankind. Stellenbosch: African Sun Media Metro, pp. 71104.Google Scholar
Franz-Odendaal, T.A., Lee-Thorp, J.A. and Chinsamy, A. (2002). New evidence for the lack of C-4 grassland expansions during the early Pliocene at Langebaanweg, South Africa. Paleobiology 28(3), 378388.Google Scholar
Fredieu, J., Levin, N., Melillo, S., Semaw, S. and Simpson, S. W. (2007). Dietary reconstruction of Early Pliocene bovids from the Gona Project Area, Ethiopia. Journal of Human Evolution S44, 110.Google Scholar
Freedman, L. (1957). The fossil Cercopithecoidea of South Africa. Annals of the Transvaal Museum 23, 121262.Google Scholar
Freedman, L. (1960). Some new cercopithecoid specimens from Makapansgat, South Africa. Palaeontologica Africana 7, 745.Google Scholar
Freedman, L. (1976). South African fossil Cercopithecoidea: a reassessment including a description of new material from Makapansgat, Sterkfonein, and Taung. Journal of Human Evolution 5, 297315.Google Scholar
Friis, I., Demissew, S. and van Breugel, P. (2011). Atlas of the Potential Vegetation of Ethiopia. Addis Ababa: Addis Ababa University Press and Shama Books.Google Scholar
Frost, S.R. (2001). New early Pliocene Cercopithecidae (Mammalia: Primates) from Aramis, Middle Awash Valley, Ethiopia. American Museum Novitates 3350, 136.Google Scholar
Frost, S.R. (2007). African Pliocene and Pleistocene cercopithecid evolution and global climate change. In: Bobe, R., Alemseged, Z. and Behrensmeyer, A.K. (Eds.), Hominin Environments in the East African Pliocene: An Assessment of the Faunal Evidence. Dordrecht: Springer, pp. 5176.Google Scholar
Frost, S.R. (2014). Fossil Cercopithecidae of the Konso Formation. In: Suwa, G., Beyene, Y. and Asfaw, B. (Eds.), The Konso-Gardula Research Project Volume 1. Paleonotological Collections: Background and Fossil Aves, Cercopithecidae, and Suidae. Tokyo: The University Museum, The University of Tokyo, Bulletin 47, pp. 4172.Google Scholar
Frost, S.R. and Alemseged, Z. (2007). Middle Pleistocene fossil Cercopithecidae from Asbole, Afar Region, Ethiopia. Journal of Human Evolution 53, 227259.Google Scholar
Frost, S.R. and Delson, E. (2002). Fossil Cercopithecidae from the Hadar Formation and surrounding areas of the Afar Depression, Ethiopia. Journal of Human Evolution 43, 687748.Google Scholar
Frost, S.R. and Kullmer, O. (2008). Cercopithecidae from the Pliocene Chiwondo Beds, Malawi-Rift. Geobios 41(6), 743749.Google Scholar
Frost, S.R., Plummer, T., Bishop, L.C., et al. (2003). Partial cranium of Cercopithecoides kimeui Leakey, 1982 from Rawi Gully, Southwestern Kenya. American Journal of Physical Anthropology 122(3), 191199.Google Scholar
Frost, S.R., Haile-Selassie, Y. and Hlusko, L. (2009). Cercopithecidae. In: Haile-Selassie, Y. and WoldeGabriel, G. (Eds.), Ardipithecus kadabba: Late Miocene Evidence from the Middle Awash, Ethiopia. Berkeley: University of California Press, pp. 135158.Google Scholar
Frost, S.R., Schwartz, H.L., Giemsch, L., et al. (2012). Refined age estimates and paleoanthropological investigations of the Manyara Beds, Tanzania. Journal of Anthropological Sciences 90, 112.Google Scholar
Frost, S.R., Jablonski, N.G. and Haile-Selassie, Y. (2014). Early Pliocene Cercopithecidae from Woranso-Mille (Central Afar, Ethiopia) and the origins of the Theropithecus oswaldi lineage. Journal of Human Evolution 76, 3953.Google Scholar
Frost, S.R., Gilbert, C.C., Pugh, K.D., Guthrie, E.H. and Delson, E. (2015). The hand of Cercopithecoides williamsi (Mammalia, Primates): earliest evidence for thumb reduction among colobine monkeys. PLoS ONE 10, e0125030.Google Scholar
Frost, S.R., Ward, C.V., Manthi, F.K. and Plavcan, J.M. (2020). Cercopithecid fossils from Kanapoi, West Turkana, Kenya (2007–2015). Journal of Human Evolution 140, 102642.Google Scholar
Fuchs, V.E. (1934). The geological work of the Cambridge Expedition to the East African Lakes, 1930–31. Geological Magazine 71, 145166.Google Scholar
Fuchs, V.E. (1950). Pleistocene events in the Baringo basin, Kenya Colony. Geological Magazine, 87(3), 149174.Google Scholar
Furstenburg, D. (2002). Springbok Antidorcas marsupialis (Zimmerman, 1780). Wild & Jag/Game & Hunt 8, 1214.Google Scholar
Gagnon, M. and Chew, A. (2000). Dietary preferences in extant African Bovidae. Journal of Mammology 81, 490511.Google Scholar
Gallotti, R. (2013). An older origin for the Acheulean at Melka Kunture (Upper Awash, Ethiopia): techno-economic behaviours at Garba IVD. Journal of Human Evolution 65, 594620.Google Scholar
Gallotti, R. and Mussi, M. (2015). The unknown Oldowan: ~1.7-million-year-old standardized obsidian small tools from Garba IV, Melka Kunture, Ethiopia. PLoS ONE 10, e0145101.Google Scholar
Gallotti, R. and Mussi, M. (Eds.) (2018). Before, during, and after the early Acheulean at Melka Kunture (Upper Awash, Ethiopia): a technoeconomic comparative analysis. In: The Emergence of the Acheulean in East Africa and Beyond: Contributions in honor of Jean Chavaillon. Cham: Springer International Publishing, pp. 5392.Google Scholar
Gallotti, R. and Piperno, M. (2003). Recent activities of the Italian Archaeological Mission at Melka Kunture: the Open Air Museum Project and the GIS application to the study of the Oldowan sites. In: Moreno, J.M., Torcal, R.M. and de la Torre Sainz, I. (Eds.), Oldowan: Rather More Than Smashing Tools, First Hominid Technology Workshop. Treballs d’Arqueologia, 9, Bellaterra. Barcelona: Universitat Autònoma de Barcelona, pp. 3775.Google Scholar
Gallotti, R., Collina, C., Raynal, J.-P., et al. (2010). The Early Middle Pleistocene site of Gombore II (Melka Kunture, Upper Awash, Ethiopia) and the issue of Acheulean bifacial shaping strategies. African Archaeological Review 27, 291322.Google Scholar
Gallotti, R., Raynal, J.-P., Geraads, D. and Mussi, M. (2014). Garba XIII (Melka Kunture, Upper Awash, Ethiopia): a new Acheulean site of the late Lower Pleistocene. Quaternary International 343, 1727.Google Scholar
Gallotti, R., Mohib, A., Fernandes, P., Graoui, M., Lefèvre, D. (2020). Dedicated core-on-anvil production of bladelet-like flakes in the Acheulean at Thomas Quarry I—L1 (Casablanca,Morocco). Scientific Reports 10, 9225.Google Scholar
García-Alix, A., Minwer-Barakat, R., Martín Suarez, E., et al. (2016). Updating the Europe–Africa small mammal exchange during the late Messinian. Journal of Biogeography 43, 13361348.Google Scholar
Garcia-Castellanos, D. and Villaseñor, A. (2011). Messinian salinity crisis regulated by competing tectonics and erosion at the Gibraltar arc. Nature 480, 359.Google Scholar
Gargani, J. and Rigollet, C.C.L. (2007). Mediterranean Sea level variations during the Messinian salinity crisis. Geophysical Research Letters 34(10).Google Scholar
Gautier, A. (1976). Animal remains from archaeological sites of terminal Paleolithic to Old Kingdom age in the Fayum. In: Wendorf, F. and Schild, R. (Eds.), Prehistory of the Nile Valley. New York: Academic Press, pp. 369381.Google Scholar
Gautier, A. (1993). The Middle Palaeolithic archaeofaunas from Bir Tarfawi (Western Desert, Egypt). In:Wendorf, F., Schild, R. and Close, A.E. (Eds.), Egypt during the Last Interglacial. New York: Plenum Press, pp. 121143.Google Scholar
Gautier, A. and Van Neer, W. (1989). Animal remains from the late Palaeolithic sequence at Wadi Kubbaniya. In:Wendorf, F., Schild, R. and Close, A.E. (Eds.), The Prehistory of Wadi Kubbanniya. Vol 2. Stratigraphy, Paleoeconomy and Environment. Dallas: SMU Press, pp. 119161.Google Scholar
Gautier, A., Ballmann, P. and Van Neer, W. (1980). Molluscs, fish, birds and mammals from the Late Paleolithic sites in Wadi Kubbaniya. In: Wendorf, F., Schild, R. and Close, A.E. (Eds.), Loaves and Fishes: The Prehistory of Wadi Kubbaniya. New Delhi: Pauls Press, pp. 281295.Google Scholar
Gawthorpe, R.L. and Leeder, M.R. (2000). Tectono-sedimentary evolution of active extensional basins. Basin Research, 12, 195218.Google Scholar
Ge, D., Zhang, Z., Xia, L., et al. (2012). Did the expansion of C 4 plants drive extinction and massive range contraction of micromammals? Inferences from food preference and historical biogeography of pikas. Palaeogeography, Palaeoclimatology, Palaeoecology 326, 160171.Google Scholar
Genise, J.F. and Harrison, T. (2018). Walking on ashes: insect trace fossils from Laetoli indicate poor grass cover associated with early hominin environments. Palaeontology 61(4), 597624.Google Scholar
Gentry, A.W. (1967). Pelorovis oldowayensis Reck, an extinct bovid from East Africa. Bulletin of the British Museum (Natural History) Geology 14, 245299.Google Scholar
Gentry, A.W. (1970). Revised classification for Makapania broomi Wells and Cooke Bovidae, Mammalia. Palaeontologica Africana 13, 6367.Google Scholar
Gentry, A.W. (1981). Notes on Bovidae (Mammalia) from the Hadar Formation, and from Amado and Geraru, Ethiopia. Kirtlandia 33, 130.Google Scholar
Gentry, A.W. (1985). The bovidae of the Omo group deposits, Ethiopia (French and American collections). In: Coppens, Y. and Howell, F.C. (Eds.), Les faunes plio-pléistocènes de la basse vallée de l’Omo (Éthiopie) Tome 1 Périssodactyles, artiodactyles (bovidae). Paris: Éditions du CNRS, pp. 119214.Google Scholar
Gentry, A.W. (1987). Pliocene Bovidae from Laetoli. In: Leakey, M.D. and Harris, J.M (Eds.), Laetoli: A Pliocene Site in Northern Tanzania. Oxford: Clarendon Press, pp. 378408.Google Scholar
Gentry, A.W. (1990). The Semliki fossil bovids. In: Boaz, N.T. (Ed.), Evolution of Environments and Hominidae in the African Western Rift Valley. Martinsville: Virginia Museum of Natural History, pp. 225234.Google Scholar
Gentry, A.W. (1996). A fossil Budorcas (Mammalia, Bovidae) from Africa. In: Stewart, K.M. and Seymour, K.L. (Eds.), Palaeoecology and Palaeoenvironments of Late Cenozoic Mammals. Toronto: University of Toronto Press, pp. 571587.Google Scholar
Gentry, A.W. (1997). Fossil ruminants (Mammalia) from the Manonga Valley, Tanzania. In: Harrison, T. (Ed.), Neogene Paleontology of the Manonga Valley, Tanzania. New York: Plenum Press, pp. 107135.Google Scholar
Gentry, A.W. (2006). A new bovine (Bovidae, Artiodactyla) from the Hadar Formation, Ethiopia. Transactions of the Royal Society of South Africa 61, 4150.Google Scholar
Gentry, A.W. (2010). Bovidae. In: Werdelin, S. and Sanders, W.J. (Eds.), Cenozoic Mammals of Africa. Berkeley: University of California Press, pp. 741796.Google Scholar
Gentry, A.W. (2011). Bovidae. In: Harrison, T. (Ed.), Paleontology and Geology of Laetoli: Human Evolution in Context. Vol. 2: Fossil Hominins and the Associated Fauna. Dordrecht: Springer, pp. 363465.Google Scholar
Gentry, A.W. and Gentry, A. (1978a). Fossil Bovidae (Mammalia) of Olduvai Gorge, Tanzania. Part I. Bulletin of the British Museum (Natural History) Geology 29, 289446.Google Scholar
Gentry, A.W. and Gentry, A. (1978b). Fossil Bovidae (Mammalia) of Olduvai Gorge, Tanzania, Part II. Bulletin of the British Museum (Natural History) Geology 30, 183.Google Scholar
George, M. (1950). A chalicothere from the Limeworks quarry of the Makapan Valley, Potgietersrust District. South African Journal of Science 46, 241242.Google Scholar
Geraads, D. (1979). La faune des gisements de Melka-Kunturé (Ethiopie): Artiodactyles, Primates. Abbay 10, 2149.Google Scholar
Geraads, D. (1980a). La faune des sites à Homo erectus des carrières Thomas (Casablanca, Maroc). Quaternaria 22, 6594.Google Scholar
Geraads, D. (1980b). Un nouveau Félidé (Fissipeda, Mamalia) du Pléistocène moyen du Maroc: Lynx thomasi n.sp. Geobios 13, 441444.Google Scholar
Geraads, D. (1981). Bovidae et Giraffidae (Artiodactyla, Mammalia) du Pléistocène de Ternifine (Algérie). Bulletin du Museum national d’ Histoire naturelle, C, 4ème séries 3, 4786.Google Scholar
Geraads, D. (1985). La faune des gisements de Melka Kunturé (Ethiopie). In: L’environnement des Hominidés au Plio-Pléistocène. Fondation Singer-Polignac. Paris: Masson, pp. 165174.Google Scholar
Geraads, D. (1986). Ruminants pléistocènes d’Oubeidiyeh. Mémoires et Travaux du Centre Recherche Français de Jérusalem 5, 93105.Google Scholar
Geraads, D. (1987a). La faune des dépôts pléistocènes de l’Ouest du lac Natron (Tanzanie); interprétation biostratigraphique. Sciences Géologiques, Bulletins et mémoires 40, 167184.Google Scholar
Geraads, D. (1987b). Dating the northern African cercopithecid fossil record. Human Evolution 2(1), 1927.Google Scholar
Geraads, D. (1989) Vertébrés fossiles du miocène supérieur du Djebel Krechem el Artsouma (Tunisie centrale). Comparaisons biostratigraphiques. Geobios 22(6), 777801.Google Scholar
Geraads, D. (1993a). Middle Pleistocene Crocidura (Mammalia, Insectivora) from Oulad Hamida I, Morocco, and their phylogenetic relationships. Proceedings, Koninklijike Nederlandse Akademie van Wetenschappen 96, 281294.Google Scholar
Geraads, D. (1993b). Kolpochoerus phacochoeroides (Thomas, 1884) (Suidae, Mammalia), du Pliocène supérieur de Ahl al Oughlam (Casablanca, Maroc). Geobios 26, 731743.Google Scholar
Geraads, D. (1994a). Rongeurs et Lagomorphes du Pléistocène moyen de la “Grotte des Rhinocéros”, carrière Oulad Hamida 1, à Casablanca, Maroc. Neues Jahrbuch für Geologie und Paläontologie – Abhandlungen 191, 147172.Google Scholar
Geraads, D. (1994b). Evolution of Bovid diversity in the Plio-Pleistocene of Africa. Historical Biology 7, 221237.Google Scholar
Geraads, D. (1994c). Girafes fossiles d’Ouganda. In: Senut, B. and Pickford, M. (Eds.), Geology and Palaeobiology of the Albertine Rift Valley, Uganda–Zaire. Orléans: CIFEG – Centre International pour la Formation et les Echanges Géologiques, pp. 375381.Google Scholar
Geraads, D. (1995). Rongeurs et insectivores du Pliocène final de Ahl al Oughlam, Casablanca, Maroc. Geobios 28, 99115.Google Scholar
Geraads, D. (1996). Le Sivatherium (Giraffidae, Mammalia) du Pliocène final d’Ahl al Oughlam (Casablanca, Maroc) et l’évolution du genre en Afrique. Paläontologische Zeitschrift 70, 623629.Google Scholar
Geraads, D. (1997). Carnivores du Pliocène terminal de Ahl al Oughlam (Casablanca, Maroc). Geobios 30, 127164.Google Scholar
Geraads, D. (1998). Rongeurs du Mio-Pliocène de Lissasfa (Casablanca, Maroc). Geobios 31, 229245.Google Scholar
Geraads, D. (2001). Rongeurs du Miocene supérieur de Chorora (Ethiopie): Dendromuridae, Muridae et conclusions. Palaeovertebrata 30, 89109.Google Scholar
Geraads, D. (2002). Plio-Pleistocene Mammalian biostratigraphy of Atlantic Morocco. Quaternaire 13, 4353.Google Scholar
Geraads, D. (2004a). First record of Dinofelis (Felidae, Mammalia) from North Africa. Neues Jahrbuch fur Geologie und Palaontologie Monatshefte 2004, 308320.Google Scholar
Geraads, D. (2004b). New skulls of Kolpochoerus phacochoeroides (Suidae: Mammalia) from the late Pliocene of Ahl al Oughlam, Morocco. Palaeontologica Africana 40, 6983.Google Scholar
Geraads, D. (2005). Pliocene Rhinocerotidae (Mammalia) from Hadar and Dikika (Lower Awash, Ethiopia), and a revision of the origin of modern African rhinos. Journal of Vertebrate Paleontology 25, 451461.Google Scholar
Geraads, D. (2006). The late Pliocene locality of Ahl al Oughlam, Morocco: vertebrate fauna and interpretation. Transactions of the Royal Society of South Africa 61, 97101.Google Scholar
Geraads, D. (2008). Plio-Pleistocene Carnivora of Northwestern Africa: a short review. Comptes Rendus Palevol 7, 591599.Google Scholar
Geraads, D. (2010a). Biogeographic relationships of Pliocene and Pleistocene North-western African mammals. Quaternary International 212, 159168.Google Scholar
Geraads, D. (2010b). Biochronologie mammalienne du Quaternaire du Maroc atlantique, dans son cadre régional. L’Anthropologie 114, 324340.Google Scholar
Geraads, D. (2010c). Tragulidae. In: Werdelin, L. and Sanders, W.J. (Eds.), Cenozoic Mammals of Africa. Berkeley: University of California Press, pp. 723729.Google Scholar
Geraads, D. (2010d). Rhinocerotidae. In: Werdelin, L. and Sanders, W.J. (Eds.), Cenozoic Mammals of Africa. Berkeley: University of California Press, pp. 675689.Google Scholar
Geraads, D. (2011). A revision of the fossil Canidae (Mammalia) of North-western Africa. Palaeontology 54, 429446.Google Scholar
Geraads, D. (2012). The faunal context of human evolution in the late Middle/Late Pleistocene of North-western Africa. In: Hublin, J.-J. and McPherron, S. (Eds.), Modern Origins: A North African Perspective. Dordrecht: Springer, pp. 4960.Google Scholar
Geraads, D. (2014). How old is the cheetah skull shape? The case of Acinonyx pardinensis (Mammalia, Felidae). Geobios 47, 3944.Google Scholar
Geraads, D. (2016a). Pleistocene Carnivora (Mammalia) from Tighennif (Ternifine), Algeria. Geobios 49, 445445.Google Scholar
Geraads, D. (2016b). Rodentia et Lagomorpha. In: Raynal, J.-P. and Mohib, A. (Eds.), Préhistoire de Casablanca. 1 ‒ La Grotte des Rhinocéros (fouilles 1991 et 1996). Villes et Sites Archéologiques du Maroc, volume 6. Rabat: Ministère de la Culture, INSAP, pp. 95104.Google Scholar
Geraads, D. (2016c). Insectivores. In: Raynal, J.-P. and Mohib, A. (Eds.), Préhistoire de Casablanca. 1 ‒ La Grotte des Rhinocéros (fouilles 1991 et 1996). Villes et Sites Archéologiques du Maroc, volume 6. Rabat: Ministère de la Culture, INSAP, pp. 105110.Google Scholar
Geraads, D. (2020). Perissodactyla (Rhinocerotidae and Equidae) from Kanapoi. Journal of Human Evolution 140, 102373.Google Scholar
Geraads, D. (2016c). Cercopithecidae. In: Raynal, J.-P. and Mohib, A. (Eds.), Préhistoire de Casablanca. 1 ‒ La Grotte des Rhinocéros (fouilles 1991 et 1996). Villes et Sites Archéologiques du Maroc, volume 6. Rabat: Ministère de la Culture, INSAP, pp. 9193.Google Scholar
Geraads, D. and Amani, F. (1997). La faune du gisement à Homo erectus de l’Aïn Maarouf près de El Hajeb (Maroc). L’Anthropologie 101, 522530.Google Scholar
Geraads, D. and Amani, F. (1998). Bovidae (Mammalia) du Pliocène final d’Ahl al Oughlam, Casablanca, Maroc. Paläontologische Zeitschrift 72, 191205.Google Scholar
Geraads, D. (2016). Bovidae. In: Raynal, J.-P. and Mohib, A. (Eds.), Préhistoire de Casablanca. 1 ‒ La Grotte des Rhinocéros (fouilles 1991 et 1996). Villes et Sites Archéologiques du Maroc, volume 6. Rabat: Ministère de la Culture, INSAP, pp. 135139.Google Scholar
Geraads, D. (2016a). Carnivora. In: Raynal, J.-P. and Mohib, A. (Eds.), Préhistoire de Casablanca. 1 ‒ La Grotte des Rhinocéros (fouilles 1991 et 1996). Villes et Sites Archéologiques du Maroc, volume 6. Rabat: Ministère de la Culture, INSAP, pp. 111119.Google Scholar
Geraads, D. (2016b). Rhinocerotidae. In: Raynal, J.-P. and Mohib, A. (Eds.), Préhistoire de Casablanca. 1 ‒ La Grotte des Rhinocéros (fouilles 1991 et 1996). Villes et Sites Archéologiques du Maroc, volume 6. Rabat: Ministère de la Culture, INSAP, pp. 121127.Google Scholar
Geraads, D. (2016c). Equidae. In: Raynal, J.-P. and Mohib, A. (Eds.), Préhistoire de Casablanca. 1 ‒ La Grotte des Rhinocéros (fouilles 1991 et 1996). Villes et Sites Archéologiques du Maroc, volume 6. Rabat: Ministère de la Culture, INSAP, pp. 129131.Google Scholar
Geraads, D. (2016d). Hippopotamidae, Suidae et Camelidae. In: Raynal, J.-P. and Mohib, A. (Eds.), Préhistoire de Casablanca. 1 ‒ La Grotte des Rhinocéros (fouilles 1991 et 1996). Villes et Sites Archéologiques du Maroc, volume 6. Rabat: Ministère de la Culture, INSAP, pp. 133134.Google Scholar
Geraads, D. and Bobe, R. (2020a). Ruminants (Giraffidae and Bovidae) from Kanapoi. Journal of Human Evolution 140, 102383.Google Scholar
Geraads, D. and Bobe, R. (2020b). Suidae from Kanapoi. Journal of Human Evolution 140, 102337.Google Scholar
Geraads, D. and Metz-Muller, F. (1999). Proboscidea (Mammalia) du Pliocène final d’Ahl al Oughlam (Casablanca, Maroc). Neues Jahrbuch fur Geologie und Palaontologie Monatshefte 1999, 5264.Google Scholar
Geraads, D. and Thomas, H. (1994). Bovidés du Plio-Pléistocène d’Ouganda. In: Senut, B. and Pickford, M. (Eds.), Geology and Palaeobiology of the Albertine Rift Valley, Uganda–Zaire. Orléans: CIFEG – Centre International pour la Formation et les Echanges Géologiques, pp. 383407.Google Scholar
Geraads, D., Zouhri, S. (2021). Eoazara xerrii n.gen. n.sp. (Mammalia, Rhinocerotidae, Elasmotheriinae) from the upper Miocene of Skoura, Ouarzazate, Morocco. Acta Palaeontologica Polonica. 66, 753765.Google Scholar
Geraads, D., Beriro, P. and Roche, H. (1980). La faune et l’industrie des sites à Homo erectus des carrières Thomas (Casablanca, Maroc). Précisions sur l’âge de ces Hominidés. Comptes Rendus de l’Académie des Sciences Paris D 291, 195198.Google Scholar
Geraads, D., Hublin, J.-J., Jaeger, J.-J., et al. (1986). The Pleistocene hominid site of Ternifine, Algeria: new results on the environment, age and human industries. Quaternary Research 25, 380386.Google Scholar
Geraads, D., Amani, F. and Hublin, J.-J. (1992). Le gisement pléistocène moyen de l’Aïn Maarouf près de El Hajeb, Maroc: présence d’un Hominidé. Comptes Rendus de l’Académie des Sciences Paris II 314, 319323.Google Scholar
Geraads, D., Amani, F., Raynal, J-P. and Sbihi-Alaoui, F.Z. (1998). La faune de mammifères du Pliocène terminal d’Ahl al Oughlam, Casablanca, Maroc. Comptes Rendus de l’Académie des Sciences Series IIA – Earth and Planetary Science 326, 671676.Google Scholar
Geraads, D., Brunet, M., Mackaye, H.T. and Vignaud, P. (2001). Pliocene Bovidae (Mammalia) from the Koro Toro australopithecine sites, Chad. Journal of Vertebrate Paleontology 21, 335346.Google Scholar
Geraads, D., Eisenmann, V. and Petter, G. (2004a). The large mammal fauna of the Oldowayan sites of Melka-Kunturé, Ethiopia. In: Chavaillon, J. and Piperno, M. (Eds.), Studies on the Early Paleolithic site of Melka Kunture, Ethiopia. Florence: Istituto Italiano di Preistoria e Protostoria, pp. 169192.Google Scholar
Geraads, D., Raynal, J.-P. and Eisenmann, V. (2004b). The earliest human occupation of North Africa: a reply to Sahnouni et al. (2002). Journal of Human Evolution 46, 751761.Google Scholar
Geraads, D., Alemseged, Z., Reed, D., Wynn, J. and Roman, D.C. (2004c). The Pleistocene fauna (other than Primates) from Asbole, lower Awash Valley, Ethiopia, and its environmental and biochronological implications. Géobios 37, 697718.Google Scholar
Geraads, D., Blondel, C., Likius, A., et al. (2008). New Hippotragini (Bovidae, Mammalia) from the Late Miocene of Toros-Menalla (Chad). Journal of Vertebrate Paleontology 28, 231242.Google Scholar
Geraads, D., Melillo, S. and Haile-Selassie, Y. (2009a). Middle Pliocene Bovidae from hominid-bearing sites in the Woranso-Mille area, Afar region, Ethiopia. Palaeontologia Africana 44, 5970.Google Scholar
Geraads, D., Blondel, C., Mackaye, H.T., et al. (2009b). Bovidae (Mammalia) from the lower Pliocene of Chad. Journal of Vertebrate Paleontology 29, 923933.Google Scholar
Geraads, D., Raynal, J.-P. and Sbihi-Alaoui, F.-Z. (2010a). Mammalian faunas from the Pliocene and Pleistocene of Casablanca (Morocco). Historical Biology 22, 275285.Google Scholar
Geraads, D., Alemseged, Z., Bobe, R. and Reed, D. (2010b). Nyctereutes lockwoodi, n. sp., a new canid (Carnivora: Mammalia) from the middle Pliocene of Dikika, Lower Awash, Ethiopia. Journal of Vertebrate Paleontology 30, 981987.Google Scholar
Geraads, D., Alemseged, Z., Bobe, R. and Reed, D. (2011). Enhydriodon dikikae, sp. nov. (Carnivora: Mammalia), a gigantic otter from the Pliocene of Dikika, Lower Awash, Ethiopia. Journal of Vertebrate Paleontology 31, 447453.Google Scholar
Geraads, D., Bobe, R. and Reed, K. (2012a). Pliocene Bovidae (Mammalia) from the Hadar Formation of Hadar and Ledi-Geraru, Lower Awash, Ethiopia. Journal of Vertebrate Paleontology 32, 180197.Google Scholar
Geraads, D., El Boughabi, S. and Zouhri, S. (2012b). Skouraia helicoides n.gen., n.sp., a new caprin bovid (Mammalia) from the late Miocene of Morocco. Palaeontologia Africana 47, 1924.Google Scholar
Geraads, D., Bobe, R. and Reed, K. (2013a). Pliocene Giraffidae (Mammalia) from the Hadar Formation of Hadar and Ledi-Geraru, Lower Awash, Ethiopia. Journal of Vertebrate Paleontology 33, 470481.Google Scholar
Geraads, D., Amani, F., Ben-Ncer, A., et al. (2013b). The rodents from the late Middle Pleistocene hominid-bearing site of J’bel Irhoud, Morocco, and their paleoclimatic significance. Quaternary Research 80, 552561.Google Scholar
Geraads, D., Bobe, R. and Manthi, F.K. (2013c). New ruminants (Mammalia) from the Pliocene of Kanapoi, Kenya, and a revision of previous collections, with a note on the Suidae. Journal of African Earth Sciences 85: 5361.Google Scholar
Geraads, D., Drapeau, M.S.M., Bobe, R. and Fleagle, J.G. (2015a). Vulpes mathisoni, sp. nov., a new fox from the Pliocene Mursi Formation of southern Ethiopia and its contribution to the origin of African foxes. Journal of Vertebrate Paleontology 34, e943765.Google Scholar
Geraads, D., Alemseged, Z., Bobe, R. and Reed, D. (2015b). Pliocene Carnivora (Mammalia) from the Hadar Formation at Dikika, lower Awash Valley, Ethiopia. Journal of African Earth Sciences 107, 2835.Google Scholar
Geraads, D., Zouhri, S. and Markov, G.N. (2019). The first Tetralophodon (Mammalia, Proboscidea) cranium from Africa. Journal of Vertebrate Paleontology 39(3), e1632321.Google Scholar
Geraads, D., Barr, W.A., Reed, D., Laurin, M. and Alemseged, Z. (2021). New remains of Camelus grattardi (Mammalia, Camelidae) from the Plio-Pleistocene of Ethiopia and the phylogeny of the genus. Journal of Mammalian Evolution 28, 359370.Google Scholar
Gereta, E. and Wolanski, E. (1998). Wildlife–water quality interactions in the Serengeti National Park, Tanzania. African Journal of Ecology 36, 114.Google Scholar
Gerlach, J. (2001). Tortoise phylogeny and the ‘Geochelone’ problem. Phelsuma 9, 124.Google Scholar
Gervais, P. (1869). Zoologie et Paléontologie générales. Paris: Arthus Bertrand.Google Scholar
Gèze, R. (1980). Les Hippopotamidae (Mammalia, Artiodactyla) du Plio-Pléistocène de l’Ethiopie (Afrique Orientale). PhD thesis, University Paris VI.Google Scholar
Gèze, R. (1985). Répartition paléoécologique et relations phylogénétiques des Hippopotamidae (Mammalia, Artiodactyla) du néogène d’Afrique Orientale. In: L’environnement des Hominidés au Plio-Pléistocène. Fondation Singer-Polignac. Paris: Masson, pp.81100.Google Scholar
Ghinassi, M., Libsekal, Y., Papini, M. and Rook, L. (2009). Palaeoenvironments of the Buia Homo site: high-resolution facies analysis and non-marine sequence stratigraphy in the Alat formation (Pleistocene Dandiero Basin, Danakil depression, Eritrea). Palaeogeography, Palaeoclimatology, Palaeoecology 280(3–4), 415431.Google Scholar
Ghinassi, M., Billi, P., Libsekal, Y., Papini, M. and Rook, L. (2013). Inferring fluvial morphodynamics and overbank flow control from 3D outcrop sections of a Pleistocene point bar, Dandiero Basin, Eritrea. Journal of Sedimentary Research 83, 10651083.Google Scholar
Ghinassi, M., Oms, O., Papini, M., et al. (2015). An integrated study of the Homo-bearing Aalat stratigraphic section (Eritrea): an expanded continental record at the Early–Middle Pleistocene transition. Journal of African Earth Sciences 112, 163185.Google Scholar
Gibbard, P.L. and Head, M.J. (2009). IUGS ratification of the Quaternary System/Period and the Pleistocene Series/Epoch with a base at 2.58 Ma. Quaternaire. Revue de l’Association française pour l’étude du Quaternaire 20(4), 411412.Google Scholar
Gibbon, R.J., Granger, D.E., Kuman, K. and Partridge, T.C. (2009). Early Acheulean technology in the Rietputs Formation, South Africa, dated with cosmogenic nuclides. Journal of Human Evolution 56(2), 152160.Google Scholar
Gibbon, R.J., Pickering, T.R., Sutton, M.B., et al. (2014). Cosmogenic nuclide burial dating of hominin-bearing Pleistocene cave deposits at Swartkrans, South Africa. Quaternary Geochronology 24, 1015.Google Scholar
Gibert, L., Scott, G.R., Scholz, D., et al. (2016). Chronology for the Cueva Victoria fossil site (SE Spain): evidence for early Pleistocene Afro-Iberian dispersals. Journal of Human Evolution 90, 183197.Google Scholar
Gifford-Gonzalez, D. (2003). The fauna from Ele Bor: evidence for the persistence of foragers into the later Holocene of arid north Kenya. African Archaeological Review 20, 81119.Google Scholar
Gilbert, C.C. (2007a). Craniomandibular morphology supporting the diphyletic origin of mangabeys and a new genus of the Cercocebus/Mandrillus clade, Procercocebus. Journal of Human Evolution 53, 69102.Google Scholar
Gilbert, C.C. (2007b). Identification and description of the first Theropithecus (Primates: Cercopithecidae) material from Bolt’s Farm, South Africa. Annals of the Transvaal Museum 44(1), 110.Google Scholar
Gilbert, C.C. (2013). Cladistic analysis of extant and fossil African papionins using craniodental data. Journal of Human Evolution 64(5), 399433.Google Scholar
Gilbert, C.C., McGraw, W.S. and Delson, E. (2009). Brief communication: Plio-Pleistocene eagle predation on fossil cercopithecids from the Humpata Plateau, southern Angola. American Journal of Physical Anthropology 139, 421429.Google Scholar
Gilbert, C., Goble, E. and Hill, A. (2010). Miocene Cercopithecoidea from the Tugen Hills, Kenya. Journal of Human Evolution 59, 465483.Google Scholar
Gilbert, C.C., Goble, E.D., Kingston, J.D. and Hill, A. (2011). Partial skeleton of Theropithecus brumpti (Primates, Cercopithecidae) from the Chemeron Formation of the Tugen Hills, Kenya. Journal of Human Evolution 61, 347362.Google Scholar
Gilbert, C.C., Steininger, C.M., Kibii, J.M. and Berger, L.R. (2015). Papio cranium from the hominin-bearing site of Malapa: implications for the evolution of modern baboon cranial morphology and South African Plio-Pleistocene biochronology. PLoS ONE 10(8), e0133361.Google Scholar
Gilbert, C.C., Frost, S.R. and Delson, E. (2016). Reassessment of Olduvai Bed I cercopithecoids: a new biochronological and biogeographical link to the South African fossil record. Journal of Human Evolution 92, 5059.Google Scholar
Gilbert, W.H. (2008). Suidae. In: Gilbert, W. and Asfaw, B. (Eds.) Homo erectus: Pleistocene Evidence from the Middle Awash, Ethiopia. Berkeley: University of California Press, pp. 4595.Google Scholar
Gilbert, W.H. and Asfaw, B. (Eds.) (2008). Homo erectus. Pleistocene Evidence from the Middle Awash, Ethiopia. Berkeley: University of California Press.Google Scholar
Ginsburg, L. (1977a). L’Hyracoïde (Mammifère subongulé) du Miocène de Beni Mellal (Maroc). Géologie Méditerranéenne 4, 241254.Google Scholar
Ginsburg, L. (1977b). Listriodon juba, suidé nouveau du Miocène de Beni Mellal (Maroc). Géologie Méditerranéenne 4, 221224.Google Scholar
Ginsburg, L. (1977c). Les carnivores du Miocène de Beni Mellal. Géologie Méditerranéenne 4, 225240.Google Scholar
Giresse, P. (2005). Mesozoic–Cenozoic history of the Congo Basin. Journal of African Earth Sciences 43, 301315.Google Scholar
Glazko, G.V. and Nei, M. (2003). Estimation of divergence times for major lineages of primate species. Molecular Biology and Evolution 20(3), 424434.Google Scholar
Gmira, S., De Lapparent, F., Geraads, D., et al. (2013). Les Tortues d’Ahl Al Oughlam, et de ses environs au Pliocène supérieur du Maroc. Geodiversitas 35, 691733.Google Scholar
Goble, E.D. (2011). Paleontology of the Chemeron Formation, Tugen Hills, Kenya with emphasis on faunal shifts and precessional climatic forcing. Doctoral dissertation, Yale University, New Haven.Google Scholar
Goble, E.D., Hill, A. and Kingston, J. (2008). Digital elevation models as heuristic tools. Paper presented at the The Paleoanthropology Society Meetings, Vancouver, CA.Google Scholar
Goldberg, P., Berna, F. and Chazan, M. (2015). Deposition and diagenesis in the Earlier Stone Age of Wonderwerk Cave, Excavation 1, South Africa. African Archaeological Review 32, 613643.Google Scholar
Gommery, D., Thackeray, J.F., Sénégas, F., Potze, S. and Kgasi, L. (2008a). The earliest primate (Parapapio sp.) from the Cradle of Humankind World Heritage site (Waypoint 160, Bolt’s Farm, South Africa). South African Journal of Science 104(9–10), 405408.Google Scholar
Gommery, D., Thackeray, J.F., Potze, S. and Braga, J. (2008b). The first recorded occurrence of honey badger of the genus Mellivora (Carnivora: Mustelidae) at Kromdraai B, South Africa. Annals of the Transvaal Museum 45(1), 145148.Google Scholar
Gommery, D., Senegas, F., Badenhorst, S., et al. (2012a). Preliminary results concerning the discovery of new fossiliferous sites at Bolt’s Farm (Cradle of Humankind, South Africa). Annals of the Ditsong National Museum of Natural History 2(1), 3345.Google Scholar
Gommery, D., Senegas, F., Badenhorst, S., Potze, S. and Kgasi, L. (2012b). Minnaar’s Cave: a Plio-Pleistocene site in the Cradle of Humankind, South Africa: its history, location, and fauna. Annals of the Ditsong National Museum of Natural History 2(1), 1931.Google Scholar
Gommery, D., Kgasi, L., Sénégas, F., et al. (2019). Waypoint 160, Bolt’s Farm Cave System: first in situ primate remains. Annals of the Ditsong National Museum of Natural History 8(8), S15.Google Scholar
Goodman, S.M. (1994). The enigma of antipredator behavior in lemurs: evidence of a large extinct eagle on Madagascar. International Journal of Primatology 15, 129134.Google Scholar
Goodwin, A.J.H. (1928). The archaeology of the Vaal River gravels. Transactions of the Royal Society of South Africa 16, 77102.Google Scholar
Goodwin, D.H., Flessa, K.W., Tellez-Duarte, M.A., et al. (2004). Detecting time-averaging and spatial mixing using oxygen isotope variation: a case study. Palaeogeography, Palaeoclimatology, Palaeoecology 205(1–2), 121.Google Scholar
Gordon, C.C. and Buikstra, J.E. (1981). Soil pH, bone preservation, and sampling bias at mortuary sites. American Antiquity 46, 566571.Google Scholar
Gordon, I. and Prins, H.H.T. (Eds.) (2008). The Ecology of Browsing and Grazing. Ecological studies vol. 195. Berlin: Springer.Google Scholar
Gordon, K.D. (1982). A study of microwear on chimpanzee molars: implications for dental microwear analysis. American Journal of Physical Anthropology 59, 195215.Google Scholar
Gordon, K.D. (1988). A review of methodology and quantification in dental microwear analysis. Scanning Microscopy 2, 11391147.Google Scholar
Goren-Inbar, N., Alperson-Afil, N., Sharon, G. and Herzlinger, G. (2018). The Acheulian Site of Gesher Benot Ya‘aqov Volume IV: The Lithic Assemblages. New York: Springer.Google Scholar
Gose, W.A. (2000). Palaeomagnetic studies of burned rocks. Journal of Archaeological Science 27, 409421.Google Scholar
Goudie, A.S. (2005). The drainage of Africa since the Cretaceous. Geomorphology 67, 437456.Google Scholar
Gouws, G. and Stewart, B.A. (2001). Potamonautid river crabs (Decapoda, Brachyura, Potamonautidae) of KwaZulu-Natal, South Africa. Water SA 27, 8598.Google Scholar
Govender, R. (2015). Preliminary phylogenetics and biogeographic history of the Pliocene seal, Homiphoca capensis from Langebaanweg, South Africa. Transactions of the Royal Society of South Africa 70(1), 2539.Google Scholar
Govender, R. and Chinsamy, A. (2013). Early Pliocene (5 Ma) shark–cetacean trophic interaction from Langebaanweg, western coast of South Africa. Palaios 28(5), 270277.Google Scholar
Govender, R., Chinsamy, A. and Ackermann, R.R. (2012). Anatomical and landmark morphometric analysis of fossil phocid seal remains from Langebaanweg, West Coast of South Africa. Transactions of the Royal Society of South Africa 67(3), 135149.Google Scholar
Gowlett, J.A.J. (1996). Rule systems in the artefacts of Homo erectus and early Homo sapiens: constrained or chosen? In: Mellars, P. and Gibson, K. (Eds.), Modelling the Early Human Mind. Cambridge: McDonald Institute, pp. 191215.Google Scholar
Gowlett, J.A.J. and Crompton, R.H. (1994). Kariandusi: Acheulean morphology and the question of allometry. African Archaeological Review 12, 342.Google Scholar
Gowlett, J.A.J., Harris, J.W.K., Walton, D. and Wood, B.A. (1981). Early archaeological sites, hominid remains and traces of fire from Chesowanja, Kenya. Nature 294(5837), 125129.Google Scholar
Gowlett, J.A.J., Brink, J.S., Herries, A.I.R., et al. (2015). At the heart of the African Acheulean: the physical, social and cognitive landscapes of Kilombe. In: Coward, F., Hosfield, R. and WenbanSmith, F. (Eds.), Settlement, Society and Cognition in Human Evolution: Landscapes in Mind. Cambridge: Cambridge University Press, pp. 7593.Google Scholar
Graham, A. and Bell, R. (1969). Factors influencing the countability of animals. East African Argriculture and Forestry Journal 34, 3843.Google Scholar
Granger, D.E., Gibbon, R.J., Kuman, K., et al. (2015). Newcosmogenic burial ages for Sterkfontein Member 2 Australopithecus and Member 5 Oldowan. Nature 522(7554), 8588.Google Scholar
Granjon, L. and Dempster, E.R. (2013). Genus Gerbilliscus gerbils. In:Happold, D. (Ed.), Mammals of Africa, Volume III: Rodents, Hares and Rabbits. London: Bloomsbury, pp. 268270.Google Scholar
Grant, K.M., Rohling, E.J., Westerhold, T., et al. (2017). 3 million year index for North African humidity/aridity and the implication of potential pan-African humid periods. Quaternary Science Reviews 171, 100118.Google Scholar
Gray, B.T. (1980). Environmental reconstruction of the Hadar Formation (Afar, Ethiopia). PhD dissertation, Case Western Reserve University.Google Scholar
Grayson, D.K. (2005). A brief history of Great Basin pikas. Journal of Biogeography 32, 21032111.Google Scholar
Green, D.J. and Alemseged, Z. (2012). Australopithecus afarensis scapular ontogeny, function, and the role of climbing in human evolution. Science 338, 514517.Google Scholar
Green, D.J., Spiewak, T.A., Seitelman, B. and Gunz, P. (2016). Scapular shape of extant hominoids and the African ape/modern human last common ancestor. Journal of Human Evolution 94, 112.Google Scholar
Greenacre, M.J. (1993). Correspondence Analysis in Practice. London: Academic Press.Google Scholar
Greenacre, M.J. (2007). Correspondence Analysis in Practice. 2nd ed. London: Chapman and Hall.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(3), 984997.Google Scholar
Greenwood, M. (1955). Fossil Hystricoidea from the Makapan Valley, Transvaal. Palaeontologica Africana 3, 7785.Google Scholar
Greenwood, M. (1958). Fossil Hystricoidea from the Makapan Valley, Transvaal. Annals and Magazine of Natural History 13, 365.Google Scholar
Gregory, J.W. (1921). The Rift Valleys and Geology of East Africa. London: London Seeley Service.Google Scholar
Griffin, D.L. (1999). The late Miocene climate of northeastern Africa: unravelling the signals in the sedimentary succession. Journal of the Geological Society, London 156, 817826.Google Scholar
Griffin, D.L. (2002). Aridity and humidity: two aspects of the late Miocene climate of North Africa and the Mediterranean. Palaeogeography, Palaeoclimatology, Palaeoecology 182(1–2), 6591.Google Scholar
Griffin, D.L. (2006). The late Neogene Sahabi rivers of the Sahara and their climatic and environmental implications for the Chad Basin. Journal of the Geological Society 163(6), 905921.Google Scholar
Griffiths, J.F. (1976). Climate and the Environment. Boulder: Westview Press.Google Scholar
Grine, F.E. (1981). Trophic differences between ‘gracile’ and ‘robust’ australopithecines: a scanning electron microscope analysis of occlusal events. South African Journal of Science 77, 203230.Google Scholar
Grine, F.E. (1982). A new juvenile hominid (Mammalia; Primates) from Member 3, Kromdraai Formation, Transvaal, South Africa. Annals of the Transvaal Museum 33, 165239.Google Scholar
Grine, F.E. (1986). Dental evidence for dietary differences in Australopithecus and Paranthropus: a quantitative analysis of permanent molar microwear. Journal of Human Evolution 15, 783822.Google Scholar
Grine, F.E. (Ed.) (1988). New craniodental fossils of Paranthropus from the Swartkrans Formation and their significance in “robust” australopithecine evolution. In: Evolutionary History of the “Robust” Australopithecines. New York: Aldine de Gruyter, pp. 223243.Google Scholar
Grine, F.E. (1989). New hominid fossils from the Swartkrans Formation (1979–1986 excavations): craniodental specimens. American Journal of Physical Anthropology 79, 409449.Google Scholar
Grine, F.E. (2005). Early Homo at Swartkrans, South Africa: a review of the evidence and an evaluation of recently proposed morphs. South African Journal of Science 101, 4352.Google Scholar
Grine, F.E. (2012). Observations on Middle Stone Age human teeth from Klasies River Main Site, South Africa. Journal of Human Evolution 63(5), 750758.Google Scholar
Grine, F.E., Klein, R.G. and Volman, T.P. (1991). Dating, archaeology and human fossils from the Middle Stone Age levels of Die Kelders, South Africa. Journal of Human Evolution, 21(5), 363395.Google Scholar
Grine, F.E., Jungers, W.L., Tobias, P.V. and Pearson, O.M. (1995). Fossil Homo femur from Berg Aukas, northern Namibia. American Journal of Physical Anthropology 97(2), 151185.Google Scholar
Grine, F.E., Jungers, W.L. and Schultz, J. (1996). Phenetic affinities among early Homo crania from East and South Africa. Journal of Human Evolution 30, 189225.Google Scholar
Grine, F.E., Leakey, M.G., Gathago, P.N., et al. (2019). Complete permanent mandibular dentition of early Homo from the upper Burgi Member of the Koobi Fora Formation, Ileret, Kenya. Journal of Human Evolution 131, 152175.Google Scholar
Grove, M. (2012). Amplitudes of orbitally induced climatic cycles and patterns of hominin speciation. Journal of Archaeological Science 39, 30853094.Google Scholar
Grün, R. and Beaumont, P. (2001). Border Cave revisited: a revised ESR chronology. Journal of Human Evolution 40(6), 467482.Google Scholar
Grün, R., Beaumont, P.B. and Stringer, C.B. (1990). ESR dating evidence for early modern humans at Border Cave in South Africa. Nature 344(6266), 537.Google Scholar
Grün, R., Brink, J.S., Spooner, N.A., et al. (1996). Direct dating of Florisbad hominid. Nature 382(6591), 500.Google Scholar
Guérin, C. (1976). Les restes de Rhinocéros du gisement Miocène de Beni Mellal, Maroc. Géologie Méditerranéenne 3, 105108.Google Scholar
Guest, N.J. (1953). The geology and petrology of the Engaruka Oldoinyo Lengai-Lake Natron Area of Northern Tanganyka Territory. PhD thesis, Sheffield.Google Scholar
Guillocheau, F., Chelalou, R., Linol, B., et al. (2015). Cenozoic landscape evolution in and around the Congo Basin: constraints from sediments and planation surfaces. In: de Wit, M.J., Guillocheau, F. and de Wit, M.C.J. (Eds.), Geology and Resource Potential of the Congo Basin. Berlin: Springer, pp. 271313.Google Scholar
Guillemot, C. (1997). Recherches paléoanthropologiques dans la basse vallée de l’Omo. Bulletin de la maison des études éthiopiennes 11, 6978.Google Scholar
Gundling, T. and Hill, A. (2000). Geological context of fossil Cercopithecoidea from eastern Africa. In: Whitehead, P.F. and Jolly, C.J. (Eds.), Old World Monkeys Cambridge; New York: Cambridge University Press, pp. 180213.Google Scholar
Gunnell, G.G., Eiting, T.P. and Geraads, D. (2011). New late Pliocene bats (Chiroptera) from Ahl al Oughlam, Morocco. Neues Jahrbuch für Geologie und Paläontologie - Abhandlungen 260, 5571.Google Scholar
Gunz, P., Bookstein, F.L., Mitteroecker, P., et al. (2009). Early modern human diversity suggests subdivided population structure and a complex out-of-Africa scenario. Proceedings of the National Academy of Sciences of the USA 106, 60946098.Google Scholar
Gunz, P., Neubauer, S., Falk, D., et al. (2020). Australopithecus afarensis endocasts suggest ape-like brain organization and prolonged brain growth. Science Advances 6, eaaz4729.Google Scholar
Gutherz, X., Lesur, J., Cauliez, J., et al. (2015). New insights on the first Neolithic societies in the Horn of Africa: the site of Wakrita, Djibouti. Journal of Field Archaeology 40, 5568.Google Scholar
Habermann, J.M., Stanistreet, I.G., Stollhofen, H., et al. (2016). In situ ~2.0 Ma trees discovered as fossil rooted stumps, lowermost Bed I, Olduvai Gorge, Tanzania. Journal of Human Evolution 90, 7487.Google Scholar
Habermann, J.M., Alberti, M., Aldeias, V., et al. (2019). Gorongosa by the sea: first Miocene fossil sites from the Urema Rift, central Mozambique, and their coastal paleoenvironmental and paleoecological contexts. Palaeogeography, Palaeoclimatology, Palaeoecology 514, 723738.Google Scholar
Hadjouis, D. (1985). Les Bovidés (Artiodactyla, Mammalia) du gisement Atérien des Phacochères (Alger, Algérie). Interprétations paléontologiques et phylogénétiques. Comptes Rendu de l’Académie des Sciences Paris 301, 12511254.Google Scholar
Hadjouis, D. (1990). Megaceroides algericus (Lydekker, 1890), du gisement des Phacochères (Alger, Algérie). Etude critique de la position systématique de Megaceroides. Quaternaire 1, 247258.Google Scholar
Hadjouis, D. (2002). Un nouveau Bovini dans le faune du Pléistocène supérieur d’Algérie. L’Anthropologie 106, 377386.Google Scholar
Hadjouis, D. (2010). The paleontology of North Africa vertebrates through Camille Arambourg’s research: a report on vertebrates’ faunae of the North Africa Neogene. Historical Biology 22, 200214.Google Scholar
Haesaerts, P., Stoops, G. and Van Vliet-Lanoë, B. (1983). Data on sediments and fossil soils. In: de Heinzelin, J. (Ed.), The Omo Group. Tervuren: Musée Royal de l’Afrique Centrale, pp. 149185.Google Scholar
Hagemann, S.I. (2010). Paleoecology and taphonomy of a hominid-bearing site: Locality 261–1, Allia Bay, Kenya. Thesis, Rutgers University, p. 129.Google Scholar
Hahsler, M. (2017). arulesViz: Interactive Visualization of Association Rules with R. The R Journal 9, 163175.Google Scholar
Hahsler, M. and Karpienko, R. (2017). Visualizing association rules in hierarchical groups. Journal of Business Economics 87, 317335.Google Scholar
Hahsler, M., Grün, B. and Hornik, K. (2005). arules – a computational environment for mining association rules and frequent item sets. Journal of Statistical Software 14(15), 125.Google Scholar
Hahsler, M., Chelluboina, S., Hornik, K. and Buchta, C. (2011). The arules R-Package Ecosystem: analyzing interesting patterns from large transaction data sets. Journal of Machine Learning Research 12, 20212025.Google Scholar
Haile-Selassie, Y. (2001). Late Miocene hominids from the Middle Awash, Ethiopia. Nature 412, 178181.Google Scholar
Haile-Selassie, Y. (2010). Phylogeny of early Australopithecus: new fossil evidence from the Woranso-Mille (central Afar, Ethiopia). Philosophical Transactions of the Royal Society B 365, 33233331.Google Scholar
Haile-Selassie, Y. and Asfaw, B. (2000). A newly discovered Early Pliocene hominid bearing site in the Mulu Basin. Journal of Physical Anthropology, 111(Supp. 30). 170.Google Scholar
Haile-Selassie, Y. and Howell, F.C. (2009). Carnivora. In: Haile-Selassie, Y. and WoldeGabriel, G. (Eds.), Ardipithecus kadabba: Late Miocene Evidence from the Middle Awash, Ethiopia. Berkeley: University of California Press, pp. 237276.Google Scholar
Haile-Selassie, Y. and Simpson, S.W. (2013). A new species of Kolpochoerus (Mammalia: Suidae) from the Pliocene of Central Afar, Ethiopia: its taxonomy and phylogenetic relationships. Journal of Mammalian Evolution 20, 115127.Google Scholar
Haile-Selassie, Y. and WoldeGabriel, G. (Eds). (2009a) Ardipithecus kadabba: Late Miocene Evidence from the Middle Awash, Ethiopia. Berkeley: University of California Press.Google Scholar
Haile-Selassie, Y. and WoldeGabriel, G. (2009b). Introduction. In: Ardipithecus kadabba: Late Miocene Evidence from the Middle Awash, Ethiopia. Berkeley: University of California Press, pp. 125.Google Scholar
Haile-Selassie, Y., Suwa, G. and White, T.D. (2004a). Late Miocene teeth from Middle Awash, Ethiopia, and early hominid dental evolution. Science 303(5663), 15031505.Google Scholar
Haile-Selassie, Y., WoldeGabriel, G., White, T.D., et al. (2004b). Mio-Pliocene mammals from the Middle Awash, Ethiopia. Geobios 37(4), 536552.Google Scholar
Haile-Selassie, Y., Deino, A., Saylor, B., Umer, M. and Latimer, B. (2007). Preliminary geology and paleontology of new hominid-bearing Pliocene localities in the central Afar region of Ethiopia. Anthropological Sciences 115, 215222.Google Scholar
Haile-Selassie, Y., White, T., Bernor, R.L., Rook, L. and Vrba, E.S. (2009a) Biochronology, faunal turnover, and evolution. In: Haile-Selassie, Y. and WoldeGabriel, G. (Eds.), Ardipithecus kadabba: Late Miocene Evidence from the Middle Awash, Ethiopia. Berkeley: University of California Press, pp. 565583.Google Scholar
Haile-Selassie, Y., Suwa, G. and White, T. (2009b). Hominidae. InHaile-Selassie, Y. and WoldeGabriel, G. (Eds.), Ardipithecus kadabba: Late Miocene Evidence from the Middle Awash, Ethiopia. Berkeley: University of California Press, pp. 159236.Google Scholar
Haile-Selassie, Y., Vrba, E. and Bibi, F. (2009c). Bovidae. In: Haile-Selassie, Y. and WoldeGabriel, G. (Eds.), Ardipithecus kadabba: Late Miocene Evidence from the Middle Awash, Ethiopia. Berkeley: University of California Press, pp. 277330.Google Scholar
Haile-Selassie, Y., Latimer, B.M., Alene, M., et al. (2010a). An early Australopithecus afarensis postcranium from Woranso-Mille, Ethiopia. Proceedings of the National Academy of Sciences, USA 107, 1212112126.Google Scholar
Haile-Selassie, Y., Saylor, B.Z., Deino, A., Alene, M., and Latimer, B.M. (2010b). New hominid fossils from Woranso-Mille (Central Afar, Ethiopia) and taxonomy of early Australopithecus. American Journal of Physical Anthropology 141, 406417.Google Scholar
Haile-Selassie, Y., Saylor, B.Z., Deino, A., et al. (2012). A new hominin foot from Ethiopia shows multiple Pliocene bipedal adaptations. Nature 483, 565569.Google Scholar
Haile-Selassie, Y., Gibert, L, Melillo, S.M., et al. (2015). New species from Ethiopia further expands Middle Pliocene hominin diversity. Nature 521, 483488.Google Scholar
Haile-Selassie, Y., Mellilo, S. and Su, D.F. (2016a). The Pliocene hominin conundrum: do more fossils mean less clarity? Proceedings of the National Academy of Sciences, USA 113, 63646371.Google Scholar
Haile-Selassie, Y., Melillo, S.M., Ryan, T.M., et al. (2016b). Dentognathic remains of Australopithecus afarensis from Nefuraytu (Woranso-Mille, Ethiopia): Comparative description, geology, and paleoecological context. Journal of Human Evolution 100, 3553.Google Scholar
Haile-Selassie, Y., Melillo, S.M., Vazzana, A., Benazzi, S. and Ryan, T.M. (2019). A 3.8-million-year-old hominin cranium from Woranso-Mille, Ethiopia. Nature 573, 214219.Google Scholar
Haileab, B. and Brown, F.H. (1994). Tephra correlations between the Gadeb prehistoric site and the Turkana Basin. Journal of Human Evolution 26, 167173.Google Scholar
Haileab, B. and Feibel, C.S. (1993). Tephra from Fejej, Ethiopia. Journal of Human Evolution 25, 515517.Google Scholar
Haileab, B., Brown, F.H., McDougall, I. and Gathogo, P.N. (2004). Gombe Group basalts and initiation of Pliocene deposition in the Turkana depression, northern Kenya and southern Ethiopia. Geological Magazine 141, 4153.Google Scholar
Hailemichael, M. (2000). The Pliocene environment of Hadar, Ethiopia: a comparative isotopic study of paleosol carbonates and lacustrine mollusk shells of the Hadar Formation. PhD dissertation, Case Western Reserve University.Google Scholar
Hailemichael, M., Aronson, J.L., Savin, S., Tevesz, M.J.S. and Carter, J.G. (2002) δ18O in mollusk shells from Pliocene Lake Hadar and modern Ethiopian lakes: implications for history of the Ethiopian monsoon. Palaeogeography, Palaeoclimatology, Palaeoecology 186, 8199.Google Scholar
Haines, R.W. and Lye, K.A. (1983). The Sedges and Rushes of East Africa. Nairobi: East African Natural History Society.Google Scholar
Halkett, D., Hart, T., Yates, R., et al. (2003). First excavation of intact Middle Stone Age layers at Ysterfontein, Western Cape Province, South Africa: implications for Middle Stone Age ecology. Journal of Archaeological Science 30, 955971.Google Scholar
Hallett-Desguez, E. (2012). Analyse Archéozoologique de la Macrofaune. In: El Hajraoui, M.A., Nespoulet, R., Debénath, A. and Dibble, H.L. (Eds.), Préhistoire de la Région de Rabat-Témara. Villes et Sites Archéologiques du Maroc. Rabat: Institut National des Sciences de l’Archéologie et du Patrimoine, pp. 239248.Google Scholar
Hamdine, W., Thévenot, M., Michaux, J. (1998). Histoire récente de l’ours brun au Maghreb. Comptes Rendus de l’Académie des Sciences Paris 321, 565570.Google Scholar
Hamilton, W.R. (1973a). North African lower Miocene rhinoceroses. Bulletin of the British Museum (Natural History) Geology 24, 349395.Google Scholar
Hamilton, W.R. (1973b). The Lower Miocene ruminants of Gebel Zelten, Libya. Bulletin of the British Museum (Natural History) Geology 21, 75150.Google Scholar
Hammer, Ø. and Harper, D.A.T. (2006). Paleontological Data Anlalysis. Oxford: Blackwell Publishing.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(1), 9 pp.Google Scholar
Hammer, Ø., Harper, D.A. and Ryan, P.D. (2020). PAST – Palaeontological Statistics, version 4.0. Paleontological Museum, University of Oslo, Oslo.Google Scholar
Handa, N., Nakatsukasa, M., Kunimatsu, Y., Tsubamoto, T. and Nakaya, H. (2015). New specimens of Chilotheridium (Perissodactyla, Rhinocerotidae) from the upper Miocene Namurungule and Nakali Formations, northern Kenya. Paleontological Research 19, 181194.Google Scholar
Handa, N., Nakatsukasa, M., Kunimatsu, Y. and Nakaya, H. (2017). A new Elasmotheriini (Perissodactyla, Rhinocerotidae) from the upper Miocene of Samburu Hills and Nakali, northern Kenya. Geobios 50, 197209.Google Scholar
Handa, N., Nakatsukasa, M., Kunimatsu, Y. and Nakaya, H. (2019). Additional specimens of Diceros (Perissodactyla, Rhinocerotidae) from the upper Miocene Nakali Formation in Nakali, central Kenya. Historical Biology 31, 262273.Google Scholar
Hanon, R., Péan, S. and Prat, S. (2018). Reassessment of anthropic modifications on the early Pleistocene hominin specimen Stw53 (Sterkfontein, South Africa). Bulletins et Mémoires de la Société d’Anthropologie de Paris 30(1–2), 4958.Google Scholar
Happold, D. and Lock, J.M. (2013). The biotic zones of Africa. In: Kingdon, J., Happold, D.C., and Butynski, T.M. (Eds.), Mammals of Africa. London: Bloomsbury, pp. 5774.Google Scholar
Haq, B.U., Hardenbol, J. and Vail, P.R. (1987). Chronology of fluctuating sea levels since the Triassic. Science 235(4793), 11561167.Google Scholar
Haradon, C.M. (2010). The ecological context of the Acheulean to Middle Stone Age transition in Africa. PhD thesis, George Washington University.Google Scholar
Harcourt-Smith, W.E.H. (2015). Origin of bipedal locomotion. In: Henke, W. and Tattersall, I. (Eds.), Handbook of Paleoanthropology, 2nd ed. Berlin: Springer, pp. 19191959.Google Scholar
Harcourt-Smith, W.E.H. and Aiello, L.C. (2004). Fossils, feet and the evolution of bipedal locomotion. Journal of Anatomy 204, 403416.Google Scholar
Hardenbol, J., Thierry, J., Farley, M.B., et al. (1998). Mesozoic and Cenozoic sequence chronostratigraphic framework of European basins. In: Graciansky, C.-P., Hardenbol, J., Jacquin, T. and Vail, P.R. (Eds.), Mesozoic and Cenozoic Sequence Stratigraphy of European Basins. SEPM, Special Publications, 60, 313.Google Scholar
Harding, R.M. and McVean, G. (2004). A structured ancestral population for the evolution of modern humans. Current Opinion in Genetics and Development 14, 667674.Google Scholar
Hare, V. and Sealy, J. (2013). Middle Pleistocene dynamics of southern Africa’s winter rainfall zone from δ13 C and δ18O values of Hoedjiespunt faunal enamel. Palaeogeography, Palaeoclimatology, Palaeoecology 374, 7280.Google Scholar
Harmand, S. (2009). Raw material and economic behaviours at Oldowan and Acheulean in the West Turkana region, Kenya. In: Adams, B. and Blades, B. (Eds.), Lithic Materials and Paleolithic Societies. Oxford: Blackwell,pp. 314.Google Scholar
Harmand, S., Lewis, J.E., Feibel, C.S., et al. (2015). 3.3-million-year-old stone tools from Lomekwi 3, West Turkana, Kenya. Nature 521, 310315.Google Scholar
Harris, J.M. (1976a). Pleistocene Giraffidae (Mammalia, Artiodactyla) from East Rudolf, Kenya. In: Savage, R.J.G. and Coryndon, S.C. (Eds.), Fossil Vertebrates of Africa, vol. 4. London: Academic Press, pp. 283332.Google Scholar
Harris, J.M. (1976b). Cranial and dental remains of Deinotherium bozasi (Mammalia: Proboscidea) from East Rudolf, Kenya. Journal of Zoology 178, 5775.Google Scholar
Harris, J.M. (1983a). Koobi Fora Research Project, Volume 2: the fossil ungulates: Proboscidea, Perissodactyla, and Suidae. In: Leakey, R.E. and Isaac, G.L. (Eds.), Koobi Fora: Researches into Geology, Palaeontology, and Human Origins. Oxford: Clarendon Press, p. 321.Google Scholar
Harris, J.M. (Ed.) (1983b). Family Suidae. In: Koobi Fora Research Project. Vol. 2: The Fossil Ungulates: Proboscidea, Perissodactyla, and Suidae. Oxford: Clarendon Press, pp. 215302.Google Scholar
Harris, J.M. (1985). Age and paleoecology of the Upper Laetolil Beds, Laetoli, Tanzania. In: Delson, E. (Ed.), Ancestors: The Hard Evidence. New York: Alan R. Liss, pp. 7681.Google Scholar
Harris, J.M. (1987). Summary. In: Leakey, M.D. and Harris, J.M. (Eds.), Laetoli: A Pliocene Site in Northern Tanzania. Oxford: Clarendon Press, pp. 524531.Google Scholar
Harris, J.M. (Ed.) (1991a). Family Bovidae. In: Koobi Fora Research Project, Vol. 3: The Fossil Ungulates: Geology, Fossil Artiodactyls, and Palaeoenvironments. Oxford: Clarendon Press, pp. 139320.Google Scholar
Harris, J.M. (Ed.) (1991b). Family Hippopotamidae. In: Koobi Fora Research Project, Vol. 3: The Fossil Ungulates: Geology, Fossil Artiodactyls, and Palaeoenvironments. Oxford: Clarendon Press, pp.3185.Google Scholar
Harris, J.M. (Ed.) (1991c). Koobi Fora Research Project, Vol. 3: The Fossil Ungulates: Geology, Fossil Artiodactyls, and Palaeoenvironments. Oxford: Clarendon Press.Google Scholar
Harris, J.M. (Ed.) (1991d). Family Giraffidae. In: Koobi Fora Research Project, Vol. 3: The Fossil Ungulates: Geology, Fossil Artiodactyls, and Palaeoenvironments. Oxford: Clarendon Press, pp. 93138.Google Scholar
Harris, J.M. (2003). Bovidae from the Lothagam succession. In: Leakey, M.G. and Harris, J.M. (Eds.), Lothagam: The Dawn of Humanity in Eastern Africa. New York: Columbia University Press, pp. 531579.Google Scholar
Harris, J.M. and Cerling, T.E. (2002). Dietary adaptations of extant and Neogene African suids. Journal of Zoology 256, 4554.Google Scholar
Harris, J.M. and Leakey, M.G. (2003a). Geology and Vertebrate Paleontology of the Early Pliocene Site of Kanapoi, Northern Kenya, Contributions in Science. Los Angeles: Natural History Museum of Los Angeles County, p. 132.Google Scholar
Harris, J.M. and Leakey, M.G. (2003b). Lothagam Suidae. In: Leakey, M.G., and Harris, J.M. (Eds.), Lothagam: The Dawn of Humanity in Eastern Africa. New York: Columbia University Press, pp. 485519.Google Scholar
Harris, J.M. and White, T.D. (1979). Evolution of the Plio-Pleistocene African Suidae. Transactions of the American Philosophical Society 69(2), 1128.Google Scholar
Harris, J.M., Brown, F.H., Leakey, M.G., Walker, A.C. and Leakey, R.E. (1988a). Pliocene and Pleistocene Hominid-Bearing Sites from West of Lake Turkana, Kenya. Science 239(4835), 2733.Google Scholar
Harris, J.M., Brown, F.H. and Leakey, M.G. (1988b). Stratigraphy and Paleontology of Pliocene and Pleistocene Localities West of Lake Turkana, Kenya. Los Angeles: Natural History Museum of Los Angeles County.Google Scholar
Harris, J.M., Leakey, M.G. and Cerling, T.E. (2003). Early Pliocene tetrapod remains from Kanapoi, Lake Turkana Basin, Kenya. In: Harris, J.M., and Leakey, M.G. (Eds.), Geology and Vertebrate Paleontology of the Early Pliocene Site of Kanapoi, Northern Kenya Contributions in Science. Los Angeles: Natural History Museum of Los Angeles County, pp. 39113.Google Scholar
Harris, J.M., Leakey, M.G. and Brown, F.H. (2006). A brief history of research at Koobi Fora, northern Kenya. Ethnohistory 53, 3569.Google Scholar
Harris, J.M., Cerling, T.E., Leakey, M.G. and Passey, B.H. (2008). Stable isotope ecology of fossil hippopotamids from the Lake Turkana Basin of East Africa. Journal of Zoology 275(3), 323331.Google Scholar
Harris, J.M., Geraads, D. and Solounias, N. (2010a). Camelidae. In: Werdelin, L. and Sanders, W.J. (Eds.), Cenozoic Mammals of Africa. Berkeley, University of California Press, pp. 815820.Google Scholar
Harris, J.M., Solounias, N. and Geraads, D. (2010b). Giraffoidea. In: Werdelin, L. and Sanders, W.J. (Eds.), Cenozoic Mammals of Africa. Berkeley, University of California Press, pp. 797811.Google Scholar
Harris, J.W.K. and Harris, K. (1981). A note on the archaeology at Laetoli. Nyame Akuma 18, 1821.Google Scholar
Harris, J.W.K. and Isaac, G.L. (1976). The Karari Industry: Early Pleistocene archaeological evidence from the terrain east of Lake Turkana, Kenya. Nature 262, 102107.Google Scholar
Harris, J.W.K., Williamson, P.G., Verniers, J., et al. (1987). Late Pliocene hominid occupation in Central Africa: the setting, context, and character of the Senga 5A site, Zaire. Journal of Human Evolution 16, 701728.Google Scholar
Harris, J.W.K., Williamson, P.G., Morris, P.J., et al. (1990). Archaeology of the Lusso Beds. In: Boaz, N.T. (Ed.), Evolution of Environments and Hominidae in the African Western Rift Valley. Martinsville: Virginia Museum of Natural History, pp. 237272.Google Scholar
Harrison, J.L. (1962). The distribution of feeding habits among animals in a tropical rain forest. Journal of Animal Ecology 31, 5364.Google Scholar
Harrison, T. (1997a). Neogene Paleontology of the Manonga Valley, Tanzania: A Window into the Evolutionary History of East Africa. New York: Springer.Google Scholar
Harrison, T. (Ed.) (1997b). Paleoecology and taphonomy of fossil localities in the Manonga Valley, Tanzania. In: Neogene Paleontology of the Manonga Valley, Tanzania: A Window into the Evolutionary History of East Africa. New York: Springer, pp. 79105.Google Scholar
Harrison, T. (2002). First recorded hominins from the Ndolanya Beds, Laetoli, Tanzania. American Journal of Physical Anthropolopy 32(Suppl.), 83.Google Scholar
Harrison, T. (2005). Fossil bird eggs from Laetoli, Tanzania: their taxonomic and paleoecological implications. Journal of African Earth Sciences 41, 289302.Google Scholar
Harrison, T. (2010a). Apes among the tangled branches of human origins. Science 327(5965), 532534.Google Scholar
Harrison, T. (2010b). Dendropithecoidea, Proconsuloidea, and Hominoidea. In: Werdelin, L. and Sanders, W.J. (Eds.), Cenozoic Mammals of Africa. Berkeley: University of California Press, pp. 429469.Google Scholar
Harrison, T. (2011a). Paleontology and Geology of Laetoli: Human Evolution in Context, Vol. 1: Geology, Geochronology, Paleoecology and Paleoenvironment. Dordrecht: Springer.Google Scholar
Harrison, T. (2011b). Paleontology and Geology of Laetoli: Human Evolution in Context, Vol. 2: Fossil Hominins and the Associated Fauna. Dordrecht: Springer.Google Scholar
Harrison, T. (Ed.) (2011c). Hominins from the Upper Laetolil and Upper Ndolanya Beds, Laetoli. In: Paleontology and Geology of Laetoli: Human Evolution in Context, Vol. 2: Fossil Hominins and the Associated Fauna. Dordrecht: Springer, pp. 141188.Google Scholar
Harrison, T. (Ed.) (2011d). Introduction: the Laetoli hominins and associated fauna. In: Paleontology and Geology of Laetoli: Human Evolution in Context, Vol. 2: Fossil Hominins and the Associated Fauna. Dordrecht: Springer, pp. 114.Google Scholar
Harrison, T. (Ed.) (2011e). Laetoli revisited: Renewed paleontological and geological investigations at localities on the Eyasi Plateau in northern Tanzania. In: Paleontology and Geology of Laetoli: Human Evolution in Context, Vol. 2: Fossil Hominins and the Associated Fauna. Dordrecht: Springer, pp. 115.Google Scholar
Harrison, T. (Ed.) (2011f). Cercopithecids (Cercopithecidae, Primates). In: Paleontology and Geology of Laetoli: Human Evolution in Context, Vol. 2: Fossil Hominins and the Associated Fauna. Dordrecht: Springer, pp. 83139.Google Scholar
Harrison, T. (Ed.) (2011g). Galagidae (Lorisoidea, Primates). In: Paleontology and Geology of Laetoli: Human Evolution in Context, Vol. 2: Fossil Hominins and the Associated Fauna. Dordrecht: Springer, pp. 7581.Google Scholar
Harrison, T. (Ed.) (2011h). Tortoises (Chelonii, Testudinidae). In: Paleontology and Geology of Laetoli: Human Evolution in Context, Vol. 2: Fossil Hominins and the Associated Fauna. Dordrecht: Springer, pp. 479503.Google Scholar
Harrison, T. (Ed.) (2011i). Coprolites: Taphonomic and paleoecological implications. In: Paleontology and Geology of Laetoli: Human Evolution in Context, Vol. 1: Geology, Geochronology, Paleoecology and Paleoenvironment. Dordrecht: Springer, pp. 279292.Google Scholar
Harrison, T. (2017). The paleoecology of the Upper Ndolanya Beds, Laetoli, Tanzania, and its implications for hominin evolution. In: Marom, A. and Hovers, E. (Eds.), Human Paleontology and Prehistory: Contributions in Honor of Yoel Rak. Dordrecht: Springer, pp. 3144.Google Scholar
Harrison, T. and Baker, E. (1997). Paleontology and biochronology of fossil localities in the Manonga Valley, Tanzania. In: Harrison, T. (Ed.), Neogene Paleontology of the Manonga Valley, Tanzania. New York: Plenum, pp. 361393.Google Scholar
Harrison, T. and Kweka, A. (2011). Paleontological localities on the Eyasi Plateau, including Laetoli. In: Harrison, T. (Ed.), Paleontology and Geology of Laetoli, Tanzania: Human Evolution in Context. Volume 1: Geology, Geochronology, Paleoecology and Paleoenvironment. Dordrecht: Springer, pp. 1745.Google Scholar
Harrison, T. and Mbago, M.L. (1997). Introduction: paleontological and geological research in the Manonga Valley, Tanzania. In: Harrison, T. (Ed.), Neogene Paleontology of the Manonga Valley, Tanzania: A Window into the Evolutionary History of East Africa. New York: Springer, pp. 132.Google Scholar
Harrison, T. and Msuya, C.P. (2005). Fossil struthionid eggshells from Laetoli, Tanzania: their taxonomic and biostratigraphic significance. Journal of African Earth Science 41, 303315.Google Scholar
Harrison, T. and Rein, T.R. (2016). The hands of fossil non-hominoid anthropoids. In: Kivell, T.L., Lemelin, P., Richmond, B.G. and Schmitt, D. (Eds.), The Evolution of the Primate Hand: Anatomical, Developmental, Functional, and Paleontological Evidence. New York: Springer, pp. 455483.Google Scholar
Hartstone-Rose, A., de Ruiter, D.J., Berger, L.R. and Churchill, S.E. (2007). A sabre-tooth felid from Cooper’s Cave (Gauteng, South Africa) and its implications for Megantereon (Felidae: Machairodontinae) taxonomy. Palaeontologica Africana 42, 99108.Google Scholar
Hartstone-Rose, A., Werdelin, L., de Ruiter, D.J., Berger, L.R. and Churchill, S.E. (2010). The Plio-Pleistocene ancestor of wild dogs, Lycaon sekowei n. sp. Journal of Paleontology 84, 290308.Google Scholar
Harvey, P.H. and Pagel, M.D. (1991). The Comparative Method in Evolutionary Biology. Oxford: Oxford University Press.Google Scholar
Hasegawa, M., Thorne, J.L. and Kishino, H. (2003). Time scale of eutherian evolution estimated without assuming a constant rate of molecular evolution. Genes & Genetic Systems 78(4), 267283.Google Scholar
Hatala, K.G., Roach, N.T., Ostrofsky, K.R., et al. (2016). Footprints reveal direct evidence of group behavior and locomotion in Homo erectus. Scientific Reports 6, 28766.Google Scholar
Hatala, K.G., Roach, N.T., Ostrofsky, K.R., et al. (2017). Hominin track assemblages from Okote Member deposits near Ileret, Kenya, and their implications for understanding fossil hominin paleobiology at 1.5 Ma. Journal of Human Evolution 112, 93104.Google Scholar
Haug, G.H. and Tiedemann, R. (1998). Effect of the formation of the Isthmus of Panama on Atlantic Ocean thermohaline circulation. Nature 393(6686), 673676.Google Scholar
Harvati, K., Stringer, C., Grün, R., et al. (2013). The Later Stone Age calvaria from Iwo Eleru, Nigeria: morphology and chronology. PLos ONE 6, e24024.Google Scholar
Hay, R.L. (1976). Geology of the Olduvai Gorge: A Study of Sedimentation in a Semiarid Basin. Berkeley: University of California Press.Google Scholar
Hay, R.L. (1978). Melilitite–carbonatite tuffs in the Laetolil Beds of Tanzania. Contributions in Mineralogy and Petrology 17, 255274.Google Scholar
Hay, R.L. (1981). Palaeoenvironment of the Laetolil Beds, northern Tanzania. In: Rapp, G. and Vondra, C.F. (Eds.), Hominid Sites: Their Geological Settings. American Association for the Advancement of Science, Selected Symposium 63. Boulder: Westview Press, pp. 724.Google Scholar
Hay, R.L. (1986). Role of tephra in the preservation of fossils in Cenozoic deposits of East Africa. In: Frostick, L.E., Renaut, R.W., Reid, I. and Tiercelin, J.J. (Eds.), Sedimentation in the African Rifts. Geological Society Special Publication No. 25. Oxford: Blackwell Scientific Publications, pp. 339344.Google Scholar
Hay, R.L. (1987). Geology of the Laetoli area. In: Leakey, M.D. and Harris, J.M. (Eds.), Laetoli: A Pliocene Site in Northern Tanzania. Oxford: Oxford University Press, pp. 2347.Google Scholar
Hay, R.L. (1990). Olduvai Gorge: a case history in the interpretation of hominid paleoenvironments in East Africa. Geological Society of America Special Paper 242, 2337.Google Scholar
Hay, R.L. (1996). Stratigraphy and lake-margin paleoenvironments of lowermost Bed II in Olduvai Gorge. Kaupia 6, 223230.Google Scholar
Hay, R.L. and Kyser, T.K. (2001). Chemical sedimentology and paleoenvironmental history of Lake Olduvai, a Pliocene lake in northern Tanzania. Geological Society of America Bulletin 113, 15051521.Google Scholar
Hay, R.L. and Leakey, M.D. (1982). The fossil footprints of Laetoli. Scientific American 246, 5057.Google Scholar
Hay, R.L. and Reeder, R.J. (1978). Calcretes of Olduvai Gorge and the Ndolanya Beds of northern Tanzania. Sedimentology 25, 649673.Google Scholar
Hays, J.D., Imbrie, J. and Shackleton, N.J. (1976). Variations in Earth’s orbit: pacemaker of ice ages. Science 194, 11211132.Google Scholar
Hayward, M.W. (2006). Prey preferences of the spotted hyaena (Crocuta crocuta) and degree of dietary overlap with the lion (Panthera leo). Journal of Zoology 270(4), 606614.Google Scholar
Head, J.J. and Bell, C.J. (2007). Snakes from Lemudong’o, Kenya Rift Valley. Kirtlandia 56, 177179.Google Scholar
Head, J.J. and Müller, J. (2020). Squamate reptiles from Kanapoi: faunal evidence for hominin paleoenvironments. Journal of Human Evolution 140, 102451.Google Scholar
Head, M.J. and Gibbard, P.L. (Eds.) (2005). Early–Middle Pleistocene transitions: an overview and recommendation for the defining boundary. In: EarlyMiddle Pleistocene Transitions: The Land–Ocean Evidence. London: Geological Society of London, Special Publication 247, pp. 118.Google Scholar
Heaton, J.L. (2006). Taxonomy of the Sterkfontein fossil Cercopithecinae: the Papionini of Members 2 and 4 (Gauteng, South Africa). Unpublished PhD dissertation, Indiana University.Google Scholar
Hedges, R.E.M. (2002). Bone diagenesis: an overview of processes. Archaeometry 44, 319328.Google Scholar
Heinrich, S., Zonneveld, K.A., Bickert, T. and Willems, H. (2011). The Benguela upwelling related to the Miocene cooling events and the development of the Antarctic Circumpolar Current: evidence from calcareous dinoflagellate cysts. Paleoceanography 26(3).Google Scholar
Heintz, E. (1973). Un nouveau bovidé du Miocène de Beni Mellal, Maroc: Benicerus theobaldi n.g. n.sp. (Bovidae, Artiodactyla, Mammalia). Annales Scientifiques de l’Université de Besançon, Géologie 3, 245248.Google Scholar
Heintz, E. (1976). Les Giraffidae (Artiodactyla, Mammalia) du Miocène de Beni Mellal, Maroc. Géologie Méditerranéenne 3, 91104.Google Scholar
Helgren, D.M. and Brooks, A.S. (1983). Geoarchaeology at Gi, a Middle Stone Age and Later Stone Age site in the northern Kalahari. Journal of Archaeological Science 10, 181197.Google Scholar
Helm, C., Cawthra, H., Cowling, R., et al. (2018). Palaeoecology of giraffe tracks in Late Pleistocene aeolianites on the Cape south coast. South African Journal of Science 114(1–2), 18.Google Scholar
Helm, C.W., Cawthra, H.C., Cowling, R.M., et al. (2019a). Pleistocene vertebrate tracksites on the Cape south coast of South Africa and their potential palaeoecological implications. Quaternary Science Reviews 235, 105857.Google Scholar
Helm, C.W., Cawthra, H.C., de Vynck, J.C., et al. (2019b). The Pleistocene fauna of the Cape south coast revealed through ichnology at two localities. South African Journal of Science 115(1–2), 19.Google Scholar
Henderson, Z., Scott, L., Rossouw, L. and Jacob, Z. (2006). Dating, paleoenvironments, and archaeology: a progress report on the Sunnyside 1 Site, Clarens, South Africa. Archeological Papers of the American Anthropological Association 16, 139149.Google Scholar
Hendey, Q.B. (1968). The Melkbos site: an Upper Pleistocene fossil occurrence in the south-western Cape Province. Annals of the South African Museum 52, 89119.Google Scholar
Hendey, Q.B. (1970). The age of the fossiliferous deposits at Langebaanweg, Cape Province. Annals of the South African Museum 56(3), 119131.Google Scholar
Hendey, Q.B. (1974). The late Cainozoic carnivora of the south-western Cape Province. Annals of the South African Museum 63(1), 112116.Google Scholar
Hendey, Q.B. (1977). Fossil bear from South Africa. South African Journal of Science 73(4), 112.Google Scholar
Hendey, Q.B. (1981a). Geological succession at Langebaanweg, Cape Province, and global events of the late Tertiary. South African Journal of Science 77(1), 3338.Google Scholar
Hendey, Q.B. (1981b). Palaeoecology of the late tertiary fossil occurrences in” E” Quarry, Langebaanweg, South Africa and a reinterpretation of their geological context. Annals of the South African Museum 84, 1104.Google Scholar
Hendey, Q.B. and Cooke, H.B.S. (1985). Kolpochoerus paiceae (Mammalia, Suidae) from Skurwerug, near Saldanha, South Africa, and its palaeoenvironmental implications. Annales of the South African Museum 97, 956.Google Scholar
Hendey, Q.B. and Hendey, H. (1968). New Quaternary fossil sites near Swartklip, Cape Province. Annals of the South African Museum 52, 4373.Google Scholar
Hennig, E. (1948). Quartärfaunen und Urgeschichte Ostafrikas. Naturwissenschaftliche Rundschau 1, 212217.Google Scholar
Henry, A.G., Ungar, P.S., Passey, B.H., et al. (2012). The diet of Australopithecus sediba. Nature 487(7405), 90.Google Scholar
Henshilwood, C.S. (1997). Identifying the collector: evidence for human processing of the Cape Dune Mole-Rat, Bathyergus suillus, from Blombos Cave, southern Cape, South Africa. Journal of Archaeological Science 24(7), 659662.Google Scholar
Henshilwood, C.S., Sealy, J.C., Yates, R., et al. (2001). Blombos Cave, Southern Cape, South Africa: preliminary report on the 1992–1999 excavations of the Middle Stone Age levels. Journal of Archaeological Science 28, 421448.Google Scholar
Henshilwood, C.S., Errico, F., Yates, R., et al. (2002). Emergence of modern human behavior: Middle Stone Age engravings from South Africa. Science 295(5558), 12781280.Google Scholar
Henshilwood, C.S., d’Errico, F., Van Niekerk, K.L., et al. (2011). A 100,000-year-old ochre-processing workshop at Blombos Cave, South Africa. Science 334(6053), 219222.Google Scholar
Hernandez Fernandez, M. and Vrba, E. (2006). Plio-Pleistcoene climatic change in the Turkana basin (East Africa): evidence from large mammal faunas. Journal of Human Evolution 50, 595626.Google Scholar
Hernesniemi, E., Giaourtsakis, I.X., Evans, A.R. and Fortelius, M. (2011). Rhinoceroses. In: Harrison, T. (Ed.), Paleontology and Geology of Laetoli: Human Evolution in Context. Vol. 2: Fossil Hominins and the Associated Fauna. Dordrecht: Springer, pp. 275294.Google Scholar
Herries, A.I.R. (2003). Magnetostratigraphic seriation of the South African hominin paleocaves. PhD dissertation, University of Liverpool.Google Scholar
Herries, A.I. and Adams, J.W. (2013). Clarifying the context, dating and age range of the Gondolin hominins and Paranthropus in South Africa. Journal of Human Evolution 65(5), 676681.Google Scholar
Herries, A.I.R. and Latham, A.G. (2009). Archaeomagnetic studies at the Cave of Hearths. The cave of hearths: Makapan middle Pleistocene research project. University of Southampton Series in Archaeology 1, 5964.Google Scholar
Herries, A.I.R. and Shaw, J. (2011). Palaeomagnetic analysis of the Sterkfontein palaeocave deposits; age implications for the hominin fossils and stone tool industries. Journal of Human Evolution 60, 523539.Google Scholar
Herries, A.I., Reed, K.E., Kuykendall, K.L. and Latham, A.G. (2006a). Speleology and magnetobiostratigraphic chronology of the Buffalo Cave fossil site, Makapansgat, South Africa. Quaternary Research 66(2), 233245.Google Scholar
Herries, A.I., Adams, J.W., Kuykendall, K.L. and Shaw, J. (2006b). Speleology and magnetobiostratigraphic chronology of the GD 2 locality of the Gondolin hominin-bearing paleocave deposits, North West Province, South Africa. Journal of Human Evolution 51(6), 617631.Google Scholar
Herries, A.I.R., Curnoe, D. and Adams, J.W. (2009). A multi-disciplinary seriation of early Homo and Paranthropus bearing palaeocaves in southern Africa. Quaternary International 202(1–2), 1428.Google Scholar
Herries, A.I., Hopley, P.J., Adams, J.W., Curnoe, D. and Maslin, M.A. (2010). Letter to the editor: Geochronology and palaeoenvironments of Southern African hominin‐bearing localities – a reply to Wrangham et al., 2009.“Shallow‐water habitats as sources of fallback foods for hominins”. American Journal of Physical Anthropology 143(4), 640646.Google Scholar
Herries, A.I.R., Pickering, R., Adams, J.W., et al. (2013). A multi-disciplinary perspective on the age of Australopithecus in southern Africa. In Reed, K.E., Fleagle, J.G. and Leakey, R.F. (Eds.), The Paleobiology of Australopithecus. Dordrecht: Springer, pp. 2140.Google Scholar
Herries, A.I., Kappen, P., Kegley, A.D., et al. (2014). Palaeomagnetic and synchrotron analysis of > 1.95 Ma fossil-bearing palaeokarst at Haasgat, South Africa. South African Journal of Science 110(3–4), 112.Google Scholar
Herries, A.I., Murszewski, A., Pickering, R., et al. (2018). Geoarchaeological and 3D visualisation approaches for contextualising in-situ fossil bearing palaeokarst in South Africa: a case study from the ~2.61 Ma Drimolen Makondo. Quaternary international 483, 90110.Google Scholar
Herries, A.I.R., Adams, J.W., Joannes-Boyau, R., et al. (2019). Integrating palaeocaves into palaeolandscapes: an analysis of cave levels and karstification history across the Gauteng Malmani dolomite, South Africa. Quaternary Science Reviews 220, 310334.Google Scholar
Herries, A.I., Martin, J.M., Leece, A.B., et al. (2020). Contemporaneity of Australopithecus, Paranthropus, and early Homo erectus in South Africa. Science 368(6486).Google Scholar
Higgs, E.S. (1967). Environment and chronology: the evidence from mammalian fauna. In: McBurney, C.B.M. (Ed.), The Haua Fteah (Cyrenaica) and the Stone Age of the South-East Mediterranean. Cambridge: Cambridge University Press, pp. 149164.Google Scholar
Hilgen, F., Kuiper, K., Krijgsman, W., Snel, E. and van der Laan, E. (2007). Astronomical tuning as the basis for high resolution chronostratigraphy: the intricate history of the Messinian Salinity Crisis. Stratigraphy 4(2–3), 231238.Google Scholar
Hilgen, F.J., Lourens, L.J., and Van Dam, J.A. (2012). The Neogene Period. In: Gradstein, F.M., Ogg, J.G., Schmitz, M.D. and Ogg, G.M. (Eds.), The Geologic Time Scale 2012. Oxford: Elsevier, pp. 923978.Google Scholar
Hill, A. (1975). Taphonomy of contemporary and Late Cenozoic East African vertebrates. PhD dissertation, University of London.Google Scholar
Hill, A. (1979a). Disarticulation and scattering of mammal skeletons. Paleobiology 5, 261274.Google Scholar
Hill, A. (1979b). Butchery and natural disarticulation: an investigatory technique. American Antiquity 44, 739744.Google Scholar
Hill, A. (1980). Postmortem damage to the remains of contemporary East African mammals. In: Behrensmeyer, A.K. and Hill, A.P. (Eds.), Fossils in the Making: Vertebrate Taphonomy and Paleoecology. Chicago: The University of Chicago Press, pp. 131152.Google Scholar
Hill, A. (1985). Early Hominid from Baringo, Kenya. Nature 315(6016), 222224.Google Scholar
Hill, A. (1989). Bone modification by modern spotted hyenas. In: Bonnichsen, R. and Sorg, M.H. (Eds.), Bone Modification. Orono: Center for the Study of the First American, Institute for Quaternary Studies, University of Maine, pp. 169178.Google Scholar
Hill, A. (1995) Faunal and environmental change in the Neogene of east Africa: evidence from the Tugen Hills sequence, Baringo District, Kenya. In: Vrba, E.S., Denton, G.H., Partridge, T.C. and Burckle, L.H. (Eds.), Paleoclimate and Evolution, with Emphasis on Human Origins. New Haven: Yale University Press, pp. 178193.Google Scholar
Hill, A. (1999). The Baringo Basin, Kenya: from Bill Bishop to BPRP. In: Andrews, P. and Banham, P. (Eds.), Late Cenozoic Environments and Hominid Evolution: a Tribute to Bill Bishop. London: Geological Society of London, pp. 8597.Google Scholar
Hill, A. (2002). Paleoanthropological research in the Tugen Hills, Kenya – introduction. Journal of Human Evolution, 42(1–2), 110. doi:10.1006/jhev.2001.0520Google Scholar
Hill, A. (2007). Preface. In: Bobe, R., Alemseged, Z. and Behrensmeyer, A.K. (Eds.), Hominin Environments in the East African Pliocene: An Assessment of the Faunal Evidence. Dordrecht: Springer, pp. xviixx.Google Scholar
Hill, A. and Behrensmeyer, A.K. (1984). Disarticulation patterns of some modern East African mammals. Paleobiology 10, 366376.Google Scholar
Hill, A. and Ward, S. (1988). Origin of the Hominidae: the record of African large Hominoid evolution between 14 my and 4 my. Yearbook of Physical Anthropology 31, 4983.Google Scholar
Hill, A., Drake, R., Tauxe, L., et al. (1985). Neogene palaeontology and geochronology of the Baringo Basin, Kenya. Journal of Human Evolution 14, 759773.Google Scholar
Hill, A., Curtis, G. and Drake, R. (1986). Sedimentary stratigraphy of the Tugen Hills, Baringo, Kenya. In: Frostick, L.E., Renaut, R.W., Rein, I. and Tiercelin, J.J. (Eds.), Sedimentation in the African Rifts. Geological Society Special Publication No. 25. London: Geological Society of London, pp. 285295.Google Scholar
Hill, A., Ward, S., Deino, A., Curtis, G., and Drake, R. (1992). Earliest Homo. Nature 355(6362), 719722.Google Scholar
Hill, A., Leakey, M., Kingston, J.D. and Ward, S.C. (2002). New cercopithecoids and a hominoid from 12.5 Ma in the Tugen Hills succession, Kenya. Journal of Human Evolution 42, 7593.Google Scholar
Hill, M.O. (1973). Diversity and evenness: a unifying notation and its consequences. Ecology 54(2), 427432.Google Scholar
Hivernel-Guerre, F. (1976). Les industries du Late Stone Age dans la région de Melka-Kunturé. In: Berhanou, A., Chavaillon, J. and Sutton, R. (Eds.), Proceedings of the 7th Congress of Prehistory and Quaternary Studies, 1971. Addis-Abeba: Ministry of Culture, pp. 9398.Google Scholar
Hlubik, S., Berna, F., Feibel, C., Braun, D. and Harris, J.W.K. (2017). Researching the nature of fire at 1.5 Mya on the site of FxJj20 AB, Koobi Fora, Kenya, using high-resolution spatial analysis and FTIR spectrometry. Current Anthropology 58, S243S257.Google Scholar
Hlubik, S., Cutts, R., Braun, D.R., et al. (2019). Hominin fire use in the Okote Member at Koobi Fora, Kenya: new evidence for the old debate. Journal of Human Evolution 133, 214229.Google Scholar
Hlusko, L.J. (2007a). Earliest evidence for Atherurus and Xenohystrix (Hystricidae, Rodentia) in Africa, from the late Miocene site of Lemudong’o, Kenya. Kirtlandia 56, 8691.Google Scholar
Hlusko, L.J. (2007b). A new late Miocene species of Paracolobus and other Cercopithecoidea (Mammalia: Primates) fossils from Lemudong’o, Kenya. Kirtlandia 56, e85.Google Scholar
Hobolth, A., Christensen, O.F., Mailund, T. and Schierup, M.H. (2007). Genomic relationships and speciation times of human, chimpanzee, and gorilla inferred from a coalescent hidden Markov model. PLoS Genetics 3(2), e7.Google Scholar
Hodell, D.A., Elmstrom, K.M. and Kennett, J.P. (1986). Latest Miocene benthic δ18O changes, global ice volume, sea level and the ‘Messinian salinity crisis. Nature 320, 411414.Google Scholar
Hodell, D.A., Curtis, J.H., Sierro, F.J. and Raymo, M.E. (2001). Correlation of late Miocene to early Pliocene sequences between the Mediterranean and North Atlantic. Paleoceanography 16(2), 164178.Google Scholar
Hodgkins, J., Marean, C.W., Venter, J.A., et al. (2020). An isotopic test of the seasonal migration hypothesis for large grazing ungulates inhabiting the Palaeo-Agulhas Plain. Quaternary Science Reviews, 235, 106221.Google Scholar
Hoetzel, S., Dupont, L.M. and Wefer, G. (2015). Miocene–Pliocene vegetation change in south-western Africa (ODP Site 1081, offshore Namibia). Palaeogeography, Palaeoclimatology, Palaeoecology 423, 102108.Google Scholar
Holbourn, A.E., Kuhnt, W., Clemens, S.C., et al. (2018). Late Miocene climate cooling and intensification of southeast Asian winter monsoon. Nature Communications 9, 1584.Google Scholar
Holland, S.M. (2003). Confidence limits on fossil ranges that account for facies changes. Paleobiology 29(4), 468479.Google Scholar
Holland, S.M. (2016). The non-uniformity of fossil preservation. Philosophical Transactions: Biological Sciences 371(1699), 111.Google Scholar
Holmgren, K., Lee-Thorp, J.A., Cooper, G.R.I., et al. (2003). Persistent millennial-scale climatic variability over the past 25,000 years in Southern Africa. Quaternary Science Reviews 22, 23112326.Google Scholar
Holt, S., Horwitz, L.K., Hoffman, J. and Codron, D. (2019). Structural density of the leopard tortoise (Stigmochelys pardalis) shell and its implications for taphonomic research. Journal of Archaeological Science: Reports 26, 101819.Google Scholar
Hönisch, B., Hemming, N.G., Archer, D., Siddal, M. and McManus, J.F. (2009). Atmospheric carbon dioxide concentration across the mid-Pleistocene transition. Science 324(5934), 15511554.Google Scholar
Hooijer, D.A. (1958). Fossil rhinoceroses from the Limeworks Cave, Makapansgat. Palaeontologica Africana 6, 113.Google Scholar
Hooijer, D.A. and Churcher, C.S. (1985). Perissodactyla of the Omo Group deposits, American collections. In: Coppens, Y. and Howell, F.C. (Eds.), Les Faunes Plio-Pléistocènes de la Basse Vallée de l’Omo (Éthiopie): Tome 1: Périssodactyles – Artiodactyles (Bovidae). Paris: CNRS, pp. 97117.Google Scholar
Hooijer, D.A. and Maglio, V.J. (1974). Hipparions from the Late Miocene and Pliocene of northwestern Kenya. Zoologische Verhandelingen 134, 134.Google Scholar
Hooijer, D.A. and Patterson, B. (1972). Rhinoceroses from the Pliocene of northwestern Kenya. Bulletin of the Museum of Comparative Zoology 144, 126.Google Scholar
Hooker, P.J. and Miller, J.A. (1979). K–Ar dating of the Pleistocene fossil hominid site at Chesowanja, North Kenya. Nature 282, 710712.Google Scholar
Hopley, P.J. and Maslin, M.A. (2010). Climate-averaging of terrestrial faunas: an example from the Plio-Pleistocene of South Africa. Paleobiology 36(1), 3250.Google Scholar
Hopley, P., Latham, A. and Marshall, J. (2006). Palaeoenvironments and palaeodiets of mid-Pliocene micromammals from Makapansgat Limeworks, South Africa: a stable isotope and dental microwear approach. Paleogeography, Paleoclimatology, Paleoecology 233, 235251.Google Scholar
Hopley, P.J., Marshall, J.D., Weedon, G.P., et al. (2007a). Orbital forcing and the spread of C4 grasses in the late Neogene: stable isotope evidence from South African speleothems. Journal of Human Evolution 53, 620634.Google Scholar
Hopley, P., Weedon, G., Marshall, J., et al. (2007b). High- and low-latitude orbital forcing of early hominin habitats in South Africa. Earth and Planetary Science Letters 256, 419432.Google Scholar
Hopley, P.J., Herries, A.I.R., Edwards Baker, S., Kuhn, B.F. and Menter, C.G. (2013). Brief communication: Beyond the South Africa cave paradigm – Australopithecus africanus from Plio-Pleistocene paleosol deposits at Taung. American Journal of Physical Anthropology 151, 316324.Google Scholar
Hopley, P.J., Weedon, G.P., Brierley, C.M., et al. (2018). Orbital precession modulates interannual rainfall variability, as recorded in an Early Pleistocene speleothem. Geology 46, 731734.Google Scholar
Hopley, P.J., Reade, H., Parrish, R., De Kock, M. and Adams, J. W. (2019). Speleothem evidence for C3 dominated vegetation during the Late Miocene (Messinian) of South Africa. Review of Palaeobotany and Palynology 264, 7589.Google Scholar
Hopley, P., Vermeesch, P., Parrish, R. and Latham, A. (2021). Clusters of flowstone ages are not supported by statistical evidence. Nature 594(7863), E10E10.Google Scholar
Horwitz, L.K. and Chazan, M. (2015). Past and present at Wonderwerk Cave (Northern Cape Province, South Africa). African Archaeological Review 32, 595612.Google Scholar
Hours, F. (1976). Le Middle Stone Age de Melka-Kunturé. In: Berhanou, A., Chavaillon, J. and Sutton, R. (Eds.), Proceedings of the 7th Congress of Prehistory and Quaternary Studies, 1971, Addis-Ababa: Ministry of Culture, pp. 99104.Google Scholar
Hovers, E. (2003). Treading carefully; site formation processes and Pliocene lithic technology. In: Martinez-Moreno, J.M., Mora, R., and de la Torre, I. (Eds.), Oldowan: Rather More Than Smashing Stones. Barcelona: Universitat Autonoma de Barcelona, pp. 145164.Google Scholar
Howell, F.C. (1968). Omo research expedition. Nature 219, 567572.Google Scholar
Howell, F.C. (1969). Remains of Hominidae from Pliocene/Pleistocene formations in the lower Omo Basin, Ethiopia. Nature 223, 12341239.Google Scholar
Howell, F.C. (1978a). Overview of the Pliocene and earlier Pleistocene of the lower Omo basin, southern Ethiopia. In: Jolly, C.J. (Ed.), Early Hominids of Africa. London: Duckworth, pp. 85130.Google Scholar
Howell, F.C. (1978b). Hominidae. In: Maglio, V.J. and Cooke, H.B.S. (Eds.), Evolution of African Mammals. Cambridge, MA: Harvard University Press, pp. 154248.Google Scholar
Howell, F.C. (1987a). Preliminary observations on Carnivora from the Sahabi formation (Libya). In: Boaz, N.T., El-Arnauti, A., Gaziry, A.W., de Heinzelin, J. and Boaz, D.D. (Eds.), Neogene Paleontology and Geology of As Sahabi. New York: Alan R. Liss, pp. 153181.Google Scholar
Howell, F.C. and Bourlière, F. (1963).African Ecology and Human Evolution. Chicago: Aldine Publishing.Google Scholar
Howell, F.C. and Coppens, Y. (1973). Deciduous teeth of Hominidae from the Pliocene/Pleistocene of the Lower Omo Basin, Ethiopia. Journal of Human Evolution 2, 461472.Google Scholar
Howell, F.C. and Coppens, Y. (1974). Inventory of remains of Hominidae from Pliocene/Pleistocene formations of the lower Omo basin, Ethiopia (1967–1972). American Journal of Physical Anthropology 40, 116.Google Scholar
Howell, F.C. and Coppens, Y. (1976). An overview of Hominidae from the Omo Succession, Ethiopia. In: Coppens, Y., Howell, F.C., Isaac, G.L. and Leakey, R.E. (Eds.), Earliest Man and Environments in the Lake Rudolf Basin. Chicago: University of Chicago Press, pp. 522532.Google Scholar
Howell, F.C. and Coppens, Y. (1983). Introduction. In: de Heinzelin, J. (Ed.), The Omo Group. Tervuren: Musée Royal de l’Afrique Centrale, pp. 15.Google Scholar
Howell, F.C. and Garcia, N. (2007). Carnivora (Mammalia) from Lemudong’o (Late Miocene: Narok District, Kenya). Kirtlandia 56, 121139.Google Scholar
Howell, F.C., Haesaerts, P. and de Heinzelin, J. (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
Hrdlička, A. (1925). The Taungs ape. American Journal of Physical Anthropology 8, 379392.Google Scholar
Hsieh, T.C., Ma, K.H. and Chao, A. (2016). iNEXT: an R package for rarefaction and extrapolation of species diversity (Hill numbers). Methods in Ecology and Evolution 7(12), 14511456.Google Scholar
Hubbell, S.P. (2001). The Unified Neutral Theory of Biodiversity and Biogeography. Monographs in Population Biology 32. Princeton: Princeton University Press.Google Scholar
Hublin, J.-J. (1992). Le fémur humain du pléistocène moyen de l’Ain Maarouf près de El Hajeb (Maroc). Comptes Rendus de l’Académie des Sciences sér. II 314, 975980.Google Scholar
Hublin, J.-J. (2001). North-western African Middle Pleistocene hominids and their bearing on the origin of Homo sapiens. In: Barham, L. and Robson-Brown, K. (Eds.), Human Roots; Africa and Asia in the Middle Pleistocene. Bristol: CHERUB, pp. 99121.Google Scholar
Hublin, J.-J. Tillier, A.M. and Tixier, J. (1987). L’humérus d’enfant moustérien (Homo 4) du Djebel Irhoud (Maroc) dans son contexte archéologique. Bulletins et Mémoires de la Société d’Anthropologie de Paris series 14, 4, 115141.Google Scholar
Hublin, J.-J., Ben-Ncer, A., Bailey, S.E., et al. (2017). New fossils from Jebel Irhoud, Morocco and the pan-African origin of Homo sapiens. Nature 546, 289292.Google Scholar
Hughes, A.R. and Tobias, P.V. (1977). A fossil skull probably of the genus Homo from Sterkfontein, Transvaal. Nature 265(5592), 310.Google Scholar
Hughes, R.H. and Hughes, J.S. (1992). A Directory of African Wetlands. Gland: IUCN/Cambridge: UK/UNEP/Nairobi: WCMC.Google Scholar
Hujer, W., Kuiper, K., Viola, T.B., Wagreich, M. and Faupl, P. (2015). Lithostratigraphy of the Late Miocene to Early Pleistocene hominid-bearing Galili Formation, southern Afar Depression, Ethiopia. Australian Journal of Earth Sciences 108(2), 105127.Google Scholar
Hull, P.M. and Norris, R.D. (2009). Evidence for abrupt speciation in a classic case of gradual evolution. Proceedings of the National Academy of Sciences 106(50), 2122421229.Google Scholar
Humphreys, A.J.B. and Thackeray, A.I. (1983). Ghaap and Gariep; later stone age studies in the Northern Cape. South African Archaeological Society Monograph Series 2.Google Scholar
Hunter, J.P. and Fortelius, M. (1994). Comparative dental occlusal morphology, facet development and microwear in two sympatric species of Listriodon (Mammalia, Suidae) from the middle Miocene of western Anatolia (Turkey). Journal of Vertebrate Paleontology 14, 105126.Google Scholar
Hurlbert, S.H. (1971). The nonconcept of species diversity: a critique and alternative parameters. Ecology 52, 577586.Google Scholar
Hutchinson, G.E. and MacArthur, R.H. (1959). A theoretical ecological model of size distributions among species of animals. The American Naturalist 93, 117125.Google Scholar
Hutson, J.M. (2012). Neotraphonomic measures of carnivore serial predation at Ngamo. Pan as an analog for interpreting open-air faunal assemblages. Journal of Archaeological Science 39, 440457.Google Scholar
Hutson, J.M. (2018). The faunal remains from Bundu Farm and Pniel 6: examining the problematic Middle Stone Age archaeological record within the southern African interior. Quaternary International 466, 178193.Google Scholar
Hutterer, R. (2010). The Middle Palaeolithic vertebrate fauna of Ifri n’Ammar. In: Nami, M. and Moser, J. (Eds.), La Grotte d’Ifri n’Ammar, Le Paléolithique Moyen. Boon: Kommission fur Archaologie Aubereuropaisher Kulturen, pp. 307314.Google Scholar
Icole, M., Taieb, M., Perinet, G., Manega, P. and Robert, C. (1987). Minéralogie des sédiments du Groupe Peninj (Lac Natron, Tanzanie). Reconstitution des paléoenvironnements lacustres. Sciences Géologiques, Bulletins et Memoires, 40, 7182.Google Scholar
Inizan, M.L., Reduron-Ballinger, M., Roche, H. and Tixier, J. (1999). Technology and Terminology of Knapped Stone (Préhistoire de la Pierre taillée 5). Nanterre: CREP.Google Scholar
Inskeep, R.R. (1987). Nelson Bay Cave. Cape Province, South Africa: The Holocene Levels. Oxford: Archaeopress.Google Scholar
Isaac, G.L. (1965). The stratigraphy of the Peninj Beds and the provenance of the Natron australopithecine mandible. Quaternaria 7, 101130.Google Scholar
Isaac, G.L. (1967). The stratigraphy of the Peninj Group – Early Middle Pleistocene formations west of Lake Natron, Tanzania. In: Bishop, W.W. and Clark, J.D. (Eds.), Background to Evolution in Africa. Chicago: University of Chicago Press, pp. 229257.Google Scholar
Isaac, G.L. (1971). The diet of early man: aspects of archaeological evidence from Lower and Middle Pleistocene sites in Africa. World Archaeology 2, 278299.Google Scholar
Isaac, G.L. (1977). Olorgesailie: Archaeological Studies of a Middle Pleistocene Lake Basin in Kenya. Chicago: Chicago University Press.Google Scholar
Isaac, G.L. (1978a). The food-sharing behavior of protohuman hominids. In: Isaac, G. and Leakey, R.E.F. (Eds.), Human Ancestors: Readings from Scientific American. San Francisco: W.H. Freeman and Company, pp. 110123.Google Scholar
Isaac, G.L. (1978b). The Olorgesailie Formation: stratigraphy, tectonics and the palaeogeographic context of the middle Pleistocene archaeological sites. In: Bishop, W.W. (Ed.), Geological Background to Fossil Man. Edinburgh: Scottish Academic Press, pp. 173206.Google Scholar
Isaac, G.L. (1981–1982). Natron Site Notes. Unpublished manuscript, Smithsonian Institution Archives.Google Scholar
Isaac, G.L. (1997). Koobi Fora Research Project Volume 5: Plio-Pleistocene archaeology. In: Leakey, R.E. (Ed.), Koobi Fora: Researches into Geology, Palaeontology, and Human Origins. Oxford: Clarendon Press, p. 596.Google Scholar
Isaacs, G.L. and Curtis, G.H. (1974). Age of early Acheulian industries from the Peninj Group, Tanzania. Nature 249(5458), 624627.Google Scholar
Isaac, G.L. and Harris, J.W.K. (1978). Archaeology. In: Leakey, M.G. and Leakey, R.E. (Eds.), Koobi Fora Research Project, Volume 1: The Fossil Hominids and an Introduction to Their Context, 1968–1974. Oxford: Clarendon Press, pp. 6485.Google Scholar
Isaac, G.L. and Harris, J.W.K. (1997). Sites stratified within the KBS tuff. In: Isaac, G.L. (Ed.), Koobi Fora Research Project Volume 5: Plio-Pleistocene Archaeology. Oxford: Clarendon Press, pp. 7199.Google Scholar
Ishida, H. and Pickford, M. (1997). A new Late Miocene hominoid from Kenya: Samburupithecus kiptalami gen. et sp. nov. Comptes Rendus de l’Académie des Sciences – Series IIA – Earth and Planetary Science 325(10), 823829.Google Scholar
Ivanovic, R.F., Valdes, P.J., Flecker, R. and Gutjahr, M. (2014). Modelling global-scale climate impacts of the late Miocene Messinian Salinity Crisis. Climate of the Past 10, 607622.Google Scholar
Jablonski, N.G. (Ed.) (1993). The phylogeny of Theropithecus. In: Theropithecus: The Rise and Fall of a Primate Genus. Cambridge: Cambridge University Press, pp. 209224.Google Scholar
Jablonski, N.G. (1994). New fossil cercopithecid remains from the Humpata Plateau, southern Angola. American Journal of Physical Anthropology 94(4), 435464.Google Scholar
Jablonski, N.G. (2002). Fossil Old World monkeys: the late Neogene radiation. In: Hartwig, W.C. (Ed.), The Primate Fossil Record (Vol. 33). Cambridge: Cambridge University Press, pp. 255299.Google Scholar
Jablonski, N.G. and Frost, S. (2010). Cercopithecoidea. In:Werdelin, L. and Sanders, W. (Eds.), Cenozoic Mammals of Africa. Berkeley: University of California Press, pp. 393428.Google Scholar
Jablonski, N.G. and Leakey, M.G. (2008). Koobi Fora Research Project: the fossil monkeys. In: Leakey, R.E. (Ed.), Koobi Fora: Researches into Geology, Paleontology, and Human Origins. San Francisco: California Academy of Sciences, p. 469.Google Scholar
Jablonski, N.G., Leakey, M.G. and Antón, M. (2008). Systematic paleontology of the cercopithecines. In: Jablonski, N.G. and Leakey, M.G. (Eds.), Koobi Fora Research Project: The Fossil Monkeys. San Francisco: California Academy of Sciences, pp. 103300.Google Scholar
Jackman, B. and Scott, J. (1983). The Marsh Lions. Boston, MA: David R. Godine.Google Scholar
Jackson, A.L., Inger, R., Parnell, A.C. and Bearhop, S. (2011). Comparing isotopic niche widths among and within communities: SIBER – Stable Isotope Bayesian Ellipses in R. Journal of Animal Ecology 80(3), 595602.Google Scholar
Jacobs, B.F. (1999). Estimation of rainfall variables from leaf characters in tropical Africa. Palaeogeography, Palaeoclimatology, Palaeoecology 145, 231250.Google Scholar
Jacobs, B.F. (2001). Estimation of low-latitude paleoclimates using fossil angiosperm leaves: examples from the Miocene Tugen Hills, Kenya. Paleobiology 28, 399421.Google Scholar
Jacobs, B.F. and Deino, A.L. (1996). Test of climate–leaf physiognomy regression models, their application to two Miocene floras from Kenya, and 40Ar/39Ar dating of the Late Miocene Kapturo site. Palaeogeography, Palaeoclimatology, Palaeoecology 123, 259271.Google Scholar
Jacobs, B.F. and Kabuye, C.H. (1987). A middle Miocene (12.2 my old) forest in the East African Rift Valley, Kenya. Journal of Human Evolution 16, 147155.Google Scholar
Jacobs, B.F., Kingston, J.D. and Jacobs, L.L. (1999). The origin of grass-dominated ecosystems. Annals of the Missouri Botanical Garden 86, 590643.Google Scholar
Jacobs, B.F., Pan, A.D. and Scotese, C.D. (2010). A review of the Cenozoic vegetation history of Africa. In: Werdelin, L. and Sanders, W.J. (Eds.), Cenozoic Mammals of Africa. Berkeley: University of California Press, pp. 5772.Google Scholar
Jacobs, Z. (2010). An OSL chronology for the sedimentary deposits from Pinnacle Point Cave 13B – a punctuated presence. Journal of Human Evolution 59 (3–4), 289305.Google Scholar
Jacobs, Z., Duller, G.A., Wintle, A.G. and Henshilwood, C.S. (2006). Extending the chronology of deposits at Blombos Cave, South Africa, back to 140 ka using optical dating of single and multiple grains of quartz. Journal of Human Evolution 51(3), 255273.Google Scholar
Jaeger, J.-J. (1971). Les Micrommamifères du “Villafranchien” inférieur du Lac Ichkeul (Tunisie): données stratigraphiques et biogéographiques nouvelles. Comptes rendus de l’Académie des Sciences D 273, 562565.Google Scholar
Jaeger, J.-J. (1973). Les faunes de Mammifères et les Hominidés fossiles du Pléistocène moyen du Maghreb. Travaux de la R.C.P. 29, 265290.Google Scholar
Jaeger, J.-J. (1975). Les Muridae (Mammalia, Rodentia) du Pliocène et du Pléistocène du Maghreb. Origine, évolution, données biogéographiques et paléoclimatiques. Thèse, Université des Sciences et Techniques du Languedoc, Montpellier.Google Scholar
Jaeger, J.J. (1976). Les Rongeurs (Mammalia, Rodentia) du Pléistocène inférieur d’Olduvai Bed I (Tanzanie). 1ère partie: les Muridés. In: Savage, R.J.G. and Coryndon, S.C. (Eds.), Fossil Vertebrates of Africa. London: Academic Press, pp. 58120.Google Scholar
Jaeger, J.J. (1977a). Les rongeurs du Miocène moyen et supérieur du Maghreb. Paleovertebrata 8, 1164.Google Scholar
Jaeger, J.-J. (1977b). Rongeurs (Mammalia, Rodentia) du Miocène de Beni-Mellal. Palaeovertebrata 7, 91132.Google Scholar
Jaeger, J.J. (1979). Les faunes de rongeurs et de lagomorphes du Pliocène et du Pléistocène d’Afrique orientale. Bulletin du Société Géologique de France 21(3), 301308.Google Scholar
Jaeger, J.-J. (1981). Les hommes fossiles du Pléistocène moyen du Maghreb. In: Sigmon, B.A., and Cybulski, J.S. (Eds.), Homo erectus: Papers in Honor of Davidson Black. Toronto: University of Toronto Press, pp. 159187.Google Scholar
Jaeger, J.-J. (1988). Origine et évolution du genre Ellobius (Mammalia, Rodentia) en Afrique Nord-Occidentale. Folia Quaternaria 57, 350.Google Scholar
Janis, C.M. (1988). An estimation of tooth volume and hypsodonty indices in ungulate mammals, and the correlation of these factors with dietary preference. Mémoires du Muséum national d’histoire naturelle, Paris 53, 367387.Google Scholar
Jarman, P.J. (1974). The social organization of antelope in relation to their ecology. Behaviour 48(3/4), 215267.Google Scholar
Jerardino, A. (2012). Large shell middens and hunter-gatherer resource intensification along the West Coast of South Africa: the Elands Bay case study. The Journal of Island and Coastal Archaeology 7(1), 76101.Google Scholar
Jerardino, A. (2016). On the origins and significance of Pleistocene coastal resource use in southern Africa with particular reference to shellfish gathering. Journal of Anthropological Archaeology 41, 213230.Google Scholar
Jerardino, A. and Marean, C.W. (2010). Shellfish gathering, marine paleoecology and modern human behavior: perspectives from cave PP13B, Pinnacle Point, South Africa. Journal of Human Evolution 59 (3–4), 412424.Google Scholar
Jernvall, J. and Fortelius, M. (2002). Common mammals drive the evolutionary increase of hypsodonty in the Neogene. Nature 417, 538540.Google Scholar
Johanson, D. (2017). The paleoanthropology of Hadar, Ethiopia. Comptes Rendus Palevol 16, 140154.Google Scholar
Johanson, D.C. (1976). Ethiopia yields first “family” of early man. National Geographic 150, 790811.Google Scholar
Johanson, D.C. and Coppens, Y. (1976). A preliminary anatomical diagnosis of the first Plio-Pleistocene hominid discoveries in the central Afar, Ethiopia. American Journal of Physical Anthropology 45, 217234.Google Scholar
Johanson, D.C. and Taieb, M. (1976). Plio-Pleistocene hominid discoveries in Hadar, Ethiopia. Nature 260, 293297.Google Scholar
Johanson, D.C., Taieb, M., Gray, B.T. and Coppens, Y. (1978a). Geological framework of the Pliocene Hadar Formation (Afar, Ethiopia) with notes on paleontology including hominids. In: Bishop, W.W. (Ed.), Geological Background to Fossil Man. Edinburgh: Scottish Academic Press, pp. 549564.Google Scholar
Johanson, D.C., White, T.D. and Coppens, Y. (1978b). A new species of the genus Australopithecus (Primates: Hominidae) from the Pliocene of eastern Africa. Kirtlandia 28, 114.Google Scholar
Johanson, D.C., Taieb, M., and Coppens, Y. (1982a). Pliocene hominids from the Hadar Formation, Ethiopia (1973–1977): stratigraphic, chronologic, and paleoenvironmental contexts, with notes on hominid morphology and systematics. American Journal of Physical Anthropology 57, 373402.Google Scholar
Johanson, D.C., Lovejoy, C.O., Kimbel, W.H., et al. (1982b). Morphology of the Pliocene partial hominid skeleton (A.L. 288–1) from the Hadar Formation, Ethiopia. American Journal of Physical Anthropology 57, 403451.Google Scholar
Johanson, D.C., White, T.D. and Coppens, Y. (1982c). Dental remains from the Hadar Formation, Ethiopia: 1974–1977 collections. American Journal of Physical Anthropology 57, 545603.Google Scholar
Johanson, D.C., Masao, F.T., Eck, G.G., et al. (1987). New partial skeleton of Homo habilis from Olduvai Gorge, Tanzania. Nature 327, 205209.Google Scholar
Johnson, B.J., Miller, G.H., Fogel, M.L. and Beaumont, P.B. (1997). The determination of late Quaternary paleoenvironments at Equus Cave, South Africa, using stable isotopes and amino acid racemization in ostrich eggshell. Palaeogeography, Palaeoclimatology, Palaeoecology 136, 121137.Google Scholar
Johnson, C.R. and McBrearty, S. (2012). Archaeology of middle Pleistocene lacustrine and spring paleoenvironments in the Kapthurin Formation, Kenya. Journal of Anthropological Archaeology 31(4), 485499.Google Scholar
Jolly, C.J. (1970). The seed eaters: a new model for hominid differentiation based on a baboon analogy. Man 5, 526.Google Scholar
Jones, S.C. and Stewart, B.A. (Eds.) (2016). Africa from MIS 6–2: Population Dynamics and Paleoenvironments. Dordrecht: Springer.Google Scholar
Jones, T., Ehardt, C.L., Butynski, T.M., et al. (2005). The highland mangabey Lophocebus kipunji: a new species of African monkey. Science 308, 11611164.Google Scholar
Jost, L. (2006). Entropy and diversity. Oikos 113(2), 363375.Google Scholar
Jost, L. (2007). Partitioning diversity into independent alpha and beta components. Ecology 88(10), 24272439.Google Scholar
Kadima, E., Delvaux, D., Sebagenzi, S.N., Tack, L. and Kabeya, S.M. (2011). Structure and geological history of the Congo Basin: an integrated interpretation of gravity, magnetic and reflection seismic data. Basin Research 23, 499527.Google Scholar
Kaiser, T.M. (2000). Proposed fossil insect modification to fossil mammalian bone from Plio-Pleistocene hominid-bearing deposits of Laetoli (northern Tanzania). Annals of theEntomological Society of America 93, 693700.Google Scholar
Kaiser, T.M. (2011). Feeding ecology and niche partitioning of the Laetoli ungulate faunas. In: Harrison, T. (Ed.), Paleontology and Geology of Laetoli, Tanzania: Human Evolution in Context. Vol. 1: Geology, Geochronology, Paleoecology and Paleoenvironment. Dordrecht: Springer, pp. 329354.Google Scholar
Kaiser, T.M. and Fortelius, M. (2003). Differential mesowear in occluding upper and lower molars: opening mesowear analysis for lower molars and premolars in hypsodont horses. Journal of Morphology 258, 6783.Google Scholar
Kaiser, T., Bromage, T.G. and Schrenk, F. (1995). Hominid Corridor Research Project update: new Pliocene fossil localities at Lake Manyara and putative oldest Early Stone Age occurrences at Laetoli (Upper Ndolanya Beds), northern Tanzania. Journal of Human Evolution 28, 117120.Google Scholar
Kaiser, T.M., Seiffert, C., Hertler, C., et al. (2010). Makuyuni, a new Lower Palaeolithic hominid site in Tanzania. Mitteilungen aus dem Hamburgischen Zoologischen Museum und Institut 106, 69110.Google Scholar
Kalb, J.E. (1993). Refined stratigraphy of the hominid-bearing Awash Group, Middle Awash valley, Afar depression, Ethiopia. Newsletters on Stratigraphy, 29, 2162.Google Scholar
Kalb, J.E. (1995). Fossil elephantoids, Awash paleolake basins, and the Afar triple junction, Ethiopia. Palaeogeography, Palaeoclimatology, Palaeoecology 114, 357368.Google Scholar
Kalb, J.E. and Mebrate, A. (1993). Fossil elephantoids from the hominid-bearing Awash Group, Middle Awash Valley, Afar depression, Ethiopia. Transactions of the American Philosophical Society 83(1), 1110.Google Scholar
Kalb, J.E., Wood, C.B., Smart, C., et al. (1980). Preliminary geology and palaeontology of the Bodo D’ar hominid Site, Afar, Ethiopia. Palaeogeography, Palaeoclimatology, Palaeoecology 30, 107120.Google Scholar
Kalb, J.E., Jolly, C.J., Mebrate, A., et al. (1982a). Fossil mammals and artefacts from the Middle Awash Valley, Ethiopia. Nature 298, 2529.Google Scholar
Kalb, J.E., Jolly, C.J., Tebedge, S., et al. (1982b). Vertebrate faunas from the Awash Group, Middle Awash Valley, Afar, Ethiopia. Journal of Vertebrate Paleontology 2, 237258.Google Scholar
Kalb, J.E., Oswald, E.B., Mebrate, A., Tebedge, C., Jolly, C. (1982c). Stratigraphy of the Awash Group, Middle Awash Valley, Afar, Ethiopia. Newsletters on Stratigraphy 11, 95127.Google Scholar
Kandel, A.W. and Conard, N.J. (2012). Settlement patterns during the earlier and Middle stone age around Langebaan Lagoon, western Cape (South Africa). Quaternary International 270, 1529.Google Scholar
Kappelman, J. (1984). Plio-Pleistocene environments of Bed I and Lower Bed II, Olduvai Gorge, Tanzania. Palaeogeography, Palaeoclimatology, Palaeoecology 48, 171196.Google Scholar
Kappelman, J. (1986). Plio-Pleistocene marine–continental correlation using habitat indicators from Olduvai Gorge, Tanzania. Quaternary Research 25, 141149.Google Scholar
Kappelman, J. (1991). The paleoenvironment of Kenyapithecus at Fort Ternan. Journal of Human Evolution 20, 95129.Google Scholar
Kappelman, J., Swisher, C.C., Fleagle, J.G., et al. (1996). Age of Australopithecus afarensis from Fejej, Ethiopia. Journal of Human Evolution 30, 139146.Google Scholar
Kappelman, J., Plummer, T., Bishop, L., Duncan, A. and Appleton, S. (1997). Bovids as indicators of Plio-Pleistocene paleoenvironments in East Africa. Journal of Human Evolution 32, 229256.Google Scholar
Karl, H.V. (2012). Human consumption of turtles of the Homo rudolfensis site Uraha (Malawi, East Africa). Archaeofauna 21, 267279.Google Scholar
Kaszycka, K.A. (2002). Status of Kromdraai. Cranial, Mandibular and Dental Morphology, Systematic Relationships, and Significance of the Kromdraai Hominids. Paris: Editions du CNRS.Google Scholar
Katoh, S., Nagaoka, S., WoldeGabriel, G., et al. (2000). Chronostratigraphy and correlation of the Plio-Pleistocene tephra layers of the Konso Formation, southern Main Ethiopian Rift, Ethiopia. Quaternary Science Reviews 19, 13051317.Google Scholar
Katoh, S., Suwa, G., Nakaya, H. and Beyene, Y. (2014). Stratigraphic and chronological context of the Konso Formation paleontology. In: Suwa, G., Beyene, Y. and Asfaw, B. (Eds.), The Konso-Gardula Research Project Volume 1. Paleonotological Collections: Background and Fossil Aves, Cercopithecidae, and Suidae. Tokyo: The University Museum, The University of Tokyo, Bulletin, 47, 1123.Google Scholar
Katoh, S., Beyene, Y., Itaya, T., et al. (2016). New geological and palaeontological age constraint for the gorilla–human lineage split. Nature 530, 215218.Google Scholar
Kaufulu, Z.M., Vrba, E.S. and White, T.D. (1981). Age of the Chiwondo Beds, northern Malawi. Annals of the Transvaal Museum 33, 18.Google Scholar
Kent, P.E. (1941). The recent history and Pleistocene deposits of the plateau north of Lake Eyasi, Tanganyika. Geological Magazine 78, 173184.Google Scholar
Kent, P.E. (1942). The Pleistocene beds of Kanam and Kanjera, Kavirondo, Kenya. Geological Magazine 79(2), 117132.Google Scholar
Kerr, J.T. and Packer, L. (1997). Habitat heterogeneity as a determinant of mammal species richness in high-energy regions. Nature 385, 252254.Google Scholar
Keyser, A.W. and Martini, J.E.J. (1991). Haasgat, a new Plio-Pleistocene fossil locality. Palaeontologia Africana 21, 119129.Google Scholar
Keyser, A.W., Menter, C.G., Moggi-Cecchi, J., Pickering, T.R. and Berger, L.R. (2000). Drimolen: a new hominid-bearing site in Gauteng, South Africa. South African Journal of Science 96(4), 193197.Google Scholar
Kiberd, P. (2006). Bundu farm: a report on archaeological and paleoenvironmental assemblages from a pan site in Bushmanland, Northern Cape, South Africa. South African Archaeological Bulletin 61, 189201.Google Scholar
Kibii, J.M. (2000). The macrofauna from Jacovec Cavern, Sterkfontein. MSc thesis, University of the Witwatersand.Google Scholar
Kibii, J.M. (2004). Comparative taxonomic, taphonomic and palaeoenvironmental analysis of 4–2.3 million year old australopithecine cave infills at Sterkfontein. PhD thesis, University of the Witwatersrand, Johannesburg.Google Scholar
Kibunjia, M. (1994). Pliocene archaeological occurrences in the Lake Turkana basin. Journal of Human Evolution 27, 159171.Google Scholar
Kidane, T., Platzman, E., Ebinger, C., Abebe, B. and Rochette, P. (2006). Paleomagnetic constraints on continental break-up processes: observations from the Main Ethiopian rift. In: Yirgu, G., Ebinger, C.J. and Maguire, P.K.H. (Eds.), The Afar Volcanic Province within the East African Rift. London: Geological Society Special Publication 259, 165183.Google Scholar
Kidane, T., Brown, F.H. and Kidney, C. (2014). Magnetostratigraphy of the Fossil-Rich Shungura Formation, southwest Ethiopia. Journal of African Earth Sciences 97, 207223.Google Scholar
Kieffer, G., Raynal, J.-P. and Bardin, G. (2002). Cadre structural et volcanologiques des sites du Paléolithique ancien de Melka Kunture (Awash, Ethiopie): Premiers résultats. In: Raynal, J.-P., Albore-Livadie, C. and Piperno, M. (Eds.), Hommes et Volcans. De l’éruption à l’objet. Les Dossiers de l’Archéologie 2, 7792.Google Scholar
Kieffer, G., Raynal, J.-P. and Bardin, G. (2004). Volcanic markers in coarse alluvium at Melka Kunture (Upper Awash, Ethiopia). In: Chavaillon, J., and Piperno, M. (Eds.), Studies on the Early Paleolithic site of Melka Kunture, Ethiopia. Florence: Istituto Italiano di Preistoria e Protostoria, pp. 93101.Google Scholar
Kimbel, W.H. (1988). Identification of a partial cranium of Australopithecus afarensis from the Koobi Fora Formation, Kenya. Journal of Human Evolution 17, 647656.Google Scholar
Kimbel, W.H. and Delezene, L.K. (2009). “Lucy” redux: a review of research on Australopithecus afarensis. American Journal of Physical Anthropology 140, 248.Google Scholar
Kimbel, W.H. and White, T.D. (1988). Variation, sexual dimorphism and the taxonomy of Australopithecus. In: Grine, F.E. (Ed.), Evolutionary History of the “Robust” Australopithecines. New York: Aldine de Gruyter, pp. 175192.Google Scholar
Kimbel, W.H., Johanson, D.C. and Coppens, Y. (1982). Pliocene hominid cranial remains from the Hadar Formation, Ethiopia. American Journal of Physical Anthropology 57, 453499.Google Scholar
Kimbel, W.H., White, T.D. and Johanson, D.C. (1984). Cranial morphology of Australopithecus afarensis: a comparative study based on a composite reconstruction of the adult skull. American Journal of Physical Anthropology 64, 337388.Google Scholar
Kimbel, W.H., Johanson, D.C. and Rak, Y. (1994). The first skull and other new discoveries of Australopithecus afarensis at Hadar, Ethiopia. Nature 368, 449451.Google Scholar
Kimbel, W.H., Walter, R.C., Johanson, D.C., et al. (1996). Late Pliocene Homo and Oldowan tools from the Hadar Formation (Kada Hadar Member), Ethiopia. Journal of Human Evolution 31, 549561.Google Scholar
Kimbel, W.H., Johanson, D.C. and Rak, Y. (1997). Systematic assessment of a maxilla of Homo from Hadar, Ethiopia. American Journal of Physical Anthropology 103, 235262.Google Scholar
Kimbel, W.H., Rak, Y. and Johanson, D.C. (2004). The Skull of Australopithecus afarensis. New York: Oxford University Press.Google Scholar
Kimbel, W. H., Lockwood, C. A., Ward, C. V., et al. (2006). Was Australopithecus anamensis ancestral to A. afarensis? A case of anagenesis in the hominin fossil record. Journal of Human Evolution 51(2), 134152.Google Scholar
King, L.C. (1951). The geology of Makapan and other caves. Transactions of the Royal Society of South Africa 33, 121151.Google Scholar
Kingdon, J. (1990). Island Africa: The Evolution of Africa’s Rare Animals and Plants. Princeton: Princeton University Press.Google Scholar
Kingdon, J. (2003). Lowly Origin: When, Where, and Why Our Ancestors First Stood Up. Princton: Princeton University Press.Google Scholar
Kingdon, J. (2015). The Kingdon Field Guide to African Mammals. London: Bloomsbury Publishing.Google Scholar
Kingdon, J. and Hoffmann, M. (2013). Mammals of Africa. Volume VI: Pigs, Hippopotamuses, Chevrotain, Giraffes, Deer and Bovids. London: Bloomsbury Publishing.Google Scholar
Kingdon, J., Happold, D., Butynski, T., et al. (2013). Mammals of Africa,Vol. 3. London: A&C Black.Google Scholar
Kingston, J.D. (1999). Environmental determinants in early hominid evolution; issues and evidence from the Tugen Hills, Kenya. In: Andrews, P. and Banham, P. (Eds.), Late Cenozoic Environments and Hominid Evolution: a Tribute to Bill Bishop. London: Geological Society of London, pp. 6984.Google Scholar
Kingston, J.D. (2007). Shifting adaptive landscapes: progress and challenges in reconstructing early hominid environments. American Journal of Physical Anthropology 134(S45), 2058.Google Scholar
Kingston, J. (2011). Stable isotopic analyses of Laetoli fossil herbivores. In: Harrison, T. (Ed.), Paleontology and Geology of Laetoli, Tanzania: Human Evolution in Context. Vol. 1: Geology, Geochronology, Paleoecology and Paleoenvironment. Dordrecht: Springer, pp. 293328.Google Scholar
Kingston, J.D. and Harrison, T. (2007). Isotopic dietary reconstructions of Pliocene herbivores at Laetoli: implications for early hominin paleoecology. Palaeogeography, Palaeoclimatology, Palaeoecology 243(3–4), 272306.Google Scholar
Kingston, J. and Hill, A. (1999). Late Miocene palaeoenvironments in Arabia: a synthesis. In: Whybrow, P.J. and Hill, A.P. (Eds.), Fossil Vertebrates of Arabia: With Emphasis on the Late Miocene Faunas, geology, and Palaeoenvironments of the Emirate of Abu Dhabi, United Arab Emirates. New Haven: Yale University Press, pp. 389407.Google Scholar
Kingston, J., Marino, B. and Hill, A. (1994). Isotopic evidence for Neogene hominins paleoenvironments in the Kenya Rift-Valley. Science 264, 955959.Google Scholar
Kingston, J.D., Jacobs, B.F., Hill, A. and Deino, A.L. (2002). Stratigraphy, age and environments of the late Miocene Mpesida Beds, Tugen Hills, Kenya. Journal of Human Evolution 42, 95116.Google Scholar
Kingston, J.D., Ditchfield, P.W., Plummer, T.W., et al. (2006). Isotopic approaches to paleoecological reconstruction at Olduvai Bed I (Tanzania) and Kanjera (Kenya). American Journal of Physical Anthropology, S42, 113–114.Google Scholar
Kingston, J.D., Deino, A.L., Edgar, R.K. and Hill, A. (2007). Astronomically forced climate change in the Kenyan Rift Valley 2.7–2.55 Ma: implications for the evolution of early hominin ecosystems. Journal of Human Evolution 53(5), 487503.Google Scholar
Kingswood, S.C. and Kumamoto, A.T. (1997). Madoqua kirkii. Mammalian Species 569, 110.Google Scholar
Kirsanow, K.O. (2009). Animal physiology, biomineral diagenesis, and the isotopic reconstruction of palaeoenvironment. Unpublished PhD dissertation, Harvard University, Cambridge, MA.Google Scholar
Kitching, I.J. and Sadler, S. (2011). Lepidoptera, Insecta. In: Harrison, T. (Ed.), Paleontology and Geology of Laetoli: Human Evolution in Context. Vol. 2: Fossil Hominins and the Associated Fauna. Dordrecht: Springer, pp. 549554.Google Scholar
Kitching, J.W. (1951). A new species of hippopotamus from Potgietersrus. South African Journal of Science 477, 209211.Google Scholar
Kitching, J.W. (1963). A fossil Orycteropus from the limeworks quarry, Makapanasgat, Potgietersrus. Palaeontologica Africana 8, 119121.Google Scholar
Kitching, J.W. (1965). A new giant hyracoid from the limeworks quarry, Makapanasgat, Potgietersrus. Palaeontologica Africana 9, 9196.Google Scholar
Kitching, J.W. (1980). On some fossil Arthropoda from the Limeworks, Makapansgat, Potgietersrus. Palaeontologica Africana 23, 6368.Google Scholar
Klein, R.G. (1972). The late Quaternary mammalian fauna of Nelson Bay Cave (Cape Province, South Africa): its implications for megafaunal extinctions and environmental and cultural change. Quaternary Research 2, 135142.Google Scholar
Klein, R.G. (1973). Geological antiquity of Rhodesian Man. Nature 244, 311312.Google Scholar
Klein, R.G. (1975). Paleoanthropological implications of the nonarcheological bone assemblage from Swartklip I, south-western Cape Province, South Africa. Quaternary Research 5, 275288.Google Scholar
Klein, R.G. (1976). The mammalian fauna of the Klasies River Mouth sites, Southern Cape Province, South Africa. South African Archaeological Bulletin 31, 7598.Google Scholar
Klein, R.G. (1977). The mammalian fauna from the Middle and Later Stone Age (Later Pleistocene) levels of Border Cave, Natal Province, South Africa. South African Archaeological Bulletin 34, 1427.Google Scholar
Klein, R.G. (1978a). A preliminary report on the larger mammals from the Boomplaas stone age cave site, Cango Valley, Oudtshoorn District, South Africa. South African Archaeological Bulletin 33, 6675.Google Scholar
Klein, R.G. (1978b). The fauna and overall interpretation of the “Cutting 10” Acheulean site at Elandsfontein (Hopefield), Southwestern Cape Province, South Africa. Quaternary Research 10, 6983.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. (1983). Palaeoenvironmental implications of Quaternary large mammals in the fynbos region. In: Deacon, H.J., Hendey, Q.B. and Lambrechts, J.J.N. (Eds.), Fynbos Palaeoecology: A Preliminary Synthesis, South African National Scientific Programmes Report No 75. Cape Town: Mills Litho, pp. 116138.Google Scholar
Klein, R.G. (1984a). Later Stone Age faunal samples from Heuningneskrans Shelter (Transvaal) and Leopard’s Hill Cave (Zambia). South African Archaeological Bulletin 39, 109116.Google Scholar
Klein, R.G. (1984b). Mammalian extinctions and Stone Age people in Africa. In: Martin, P.S. and Klein, R.G. (Eds.), Quaternary Extinctions: A Prehistoric Revolution. Tucson: University of Arizona Press, pp. 553573.Google Scholar
Klein, R.G. (Ed.) (1984c). The large mammals of southern Africa: Late Pliocene to recent. In: Southern African Prehistory and Paleoenvironments. Rotterdam: Balkema, pp. 107146.Google Scholar
Klein, R.G. (1988). The archaeological significance of animal bones from Acheulian sites in southern Africa. African Archaeological Review 6, 326.Google Scholar
Klein, R.G. (1989). Why does skeletal part representation differ between smaller and larger bovids at Klasies River Mouth and other archeological sites? Journal of Archaeological Science 16(4), 363381.Google Scholar
Klein, R.G. (1991). Size variation in the Cape dune molerat (Bathyergus suillus) and Late Quaternary climatic change in the southwestern Cape Province, South Africa. Quaternary Research 36(3), 243256.Google Scholar
Klein, R.G. and Cruz-Uribe, K. (1984). The Analysis of Animal Bones from Archaeological Sites. Chicago: University of Chicago Press.Google Scholar
Klein, R.G. and Cruz-Uribe, K. (1987). Large mammal and tortoise bones from Eland’s Bay Cave and nearby sites, Western Cape Province, South Africa. In: Parkington, J.E. and Hall, M. (Eds.), Papers in the Prehistory of the Western Cape, South Africa. Cambridge: BAR International Series 322, pp. 132163.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
Klein, R.G. and Cruz-Uribe, K. (2016). Large mammal and tortoise bones from Elands Bay Cave (South Africa): implications for Later Stone Age environment and ecology. Southern African Humanities 29(1), 259282.Google Scholar
Klein, R.G. and Cruz-Uribe, K. (2000). Middle and Later Stone Age large mammal and tortoise remains from Die Kelders Cave 1, Western Cape Province, South Africa. Journal of Human Evolution 38, 169195.Google Scholar
Klein, R.G. and Scott, K. (1986). Re-analysis of faunal assemblages from the Haua Fteah and other late Quaternary archaeological sites in Cyernaican Libya. Journal of Archaeological Science 13, 515542.Google Scholar
Klein, R.G., Cruz-Uribe, K. and Beaumont, P.B. (1991). Environmental, ecological, and paleoanthropological implications of the Late Pleistocene mammalian fauna from Equus Cave, Northern Cape Province, South Africa. Quaternary Research 36, 94119.Google Scholar
Klein, R.G., Avery, G., Cruz-Uribe, K., et al. (1999). Duinefontein 2: an Acheulean site in the western Cape province of South Africa. Journal of Human Evolution 37(2), 153190.Google Scholar
Klein, R.G., Avery, G., Cruz-Uribe, K., et al. (2004). The Ysterfontein 1 Middle Stone Age site, South Africa, and early human exploitation of coastal resources. Proceedings of the National Academy of Sciences 101(16), 57085715.Google Scholar
Klein, R.G., Avery, G., Cruz-Uribe, K. and Steele, T.E. (2007). The mammalian fauna associated with an archaic hominin skullcap and later Acheulean artifacts at Elandsfontein, Western Cape Province, South Africa. Journal of Human Evolution 52, 164186.Google Scholar
Kleindienst, M.R. (1962). Component of the East African Acheulean assemblage: an analytic approach. In: Mortelmans, G. and Nenquin, J. (Eds.), Actes du IV Congrès Panafricain de Préhistoire et de l0 Etude du Quaternaire, Leopoldville, 1959. Belgie Annalen, Musée Royal de l0 Afrique Centrale, Tervuren, pp. 81e108.Google Scholar
Kleinsasser, L.L., Quade, J., Mcintosh, W.C., et al. (2008). Stratigraphy and geochronology of the late Miocene Adu-Asa Formation at Gona, Ethiopia. In: Quade, J. and Wynn, J.G. (Eds.), The Geology of Early Humans in the Horn of Africa. Geological Society of America Special Paper 446. Boulder: Geological Society of America, pp. 3365.Google Scholar
Koch, C.F. (1987). Prediction of sample size effects on the measured temporal and geographic distribution patterns of species. Paleobiology 3(1), 100107.Google Scholar
Kohl-Larsen, L. (1943). Auf den Sporen des Vormenschen. Stuttgart: Strecker und Schröder.Google Scholar
Köhler, C.M., Heslop, D., Krijgsman, W. and Dekkers, M.J. (2010). Late Miocene paleoenvironmental changes in North Africa and the Mediterranean recorded by geochemical proxies (Monte Gibliscemi section, Sicily). Palaeogeography, Palaeoclimatology, Palaeoecology 285, 6673.Google Scholar
Köhler, M. (1993). Skeleton and Habitat of Recent and Fossil Ruminants. Münchner geowissenschaftliche Abhandlungen Reihe A, Geologie und Paläontologie 25, pp. 188.Google Scholar
Kohn, M.J., Schoeninger, M.J. and Valley, J.W. (1996). Herbivore tooth oxygen isotope compositions: effects of diet and physiology. Geochimica et Cosmochimica Acta 60, 38893896.Google Scholar
Kovarovic, K. (2004). Bovids as palaeoenvironmental indicators. An ecomorphological analysis of bovid postcranial remains from Laetoli, Tanzania. Ph.D. Dissertation, University of London.Google Scholar
Kovarovic, K. and Andrews, P. (2007). Bovid postcranial ecomological survey of the Laetoli paleoenvironment. Journal of Human Evolution 52, 663680.Google Scholar
Kovarovic, K. and Andrews, P. (2011). Environmental change within the Laetoli fossiliferous sequence: vegetation catenas and bovid ecomorphology. In: Harrison, T. (Ed.), Paleontology and Geology of Laetoli, Tanzania: Human Evolution in Context. Vol. 1: Geology, Geochronology, Paleoecology and Paleoenvironment. Dordrecht: Springer, pp. 367380.Google Scholar
Kovarovic, K., Andrews, P. and Aiello, L. (2002). The palaeoecology of the Upper Ndolanya Beds at Laetoli, Tanzania. Journal of Human Evololution 43, 395418.Google Scholar
Kovarovic, K., Slepkov, R. and McNulty, K.P. (2013). Ecological continuity between Lower and Upper Bed II, Olduvai Gorge, Tanzania. Journal of Human Evolution 64, 538555.Google Scholar
Krell, F.-T. and Schawaller, W. (2011). Beetles (Insecta: Coleoptera). In: Harrison, T. (Ed.), Paleontology and Geology of Laetoli: Human Evolution in Context. Vol. 2: Fossil Hominins and the Associated Fauna. Dordrecht: Springer, pp. 535554.Google Scholar
Krigbaum, J., Berger, M.H., Daegling, D.J. and McGraw, W.S. (2013). Stable isotope canopy effects for sympatric monkeys at Tai Forest, Cote d’Ivoire. Biology Letters 9, 20130466.Google Scholar
Krijgsman, W., Hilgen, F.J., Raffi, I., Sierro, F.J. and Wilson, D.S. (1999). Chronology, causes and progression of the Messinian salinity crisis. Nature 400, 652655.Google Scholar
Kröpelin, S., Verschuren, D., Lézine, A.-M., et al. (2008). Climate-driven ecosystem succession in the Sahara: the past 6000 years. Science 320, 765768.Google Scholar
Kruger, A. and Badenhorst, S. (2018). Remains of a barn owl (Tyto alba) from the Dinaledi Chamber, Rising Star Cave, South Africa. South African Journal of Science 114(11–12), 15.Google Scholar
Kruuk, H. (1972). The Spotted Hyena. Chicago: The University of Chicago Press.Google Scholar
Kruuk, H. and Turner, M. (1967). Comparative notes on predation by lion, leopard, cheetah and wild dog in the Serengeti area, East Africa. Mammalia 31, 127.Google Scholar
Ksepka, D.T. and Thomas, D.B. (2011). Multiple cenozoic invasions of Africa by penguins (Aves, Sphenisciformes). Proceedings of the Royal Society B: Biological Sciences 279(1730), 10271032.Google Scholar
Kuhn, B.F., Berger, L.R. and Skinner, J.D. (2008). Examining criteria for identifying and differentiating fossil faunal assemblages accumulated by hyenas and hominins using extant hyenid accumulations. International Journal of Osteoarchaeology 20, 1535.Google Scholar
Kuhn, B.F., Werdelin, L., Hartstone-Rose, A., Lacruz, R.S. and Berger, L.R. (2011). Carnivoran remains from the Malapa hominin site, South Africa. PLoS ONE, 6(11), e26940.Google Scholar
Kuhn, B.F., Carlson, K., Hopley, P.J., Zipfel, B. and Berger, L.R. (2015). Identification of fossilized eggshells from the Taung hominin locality, Taung, Northwest Province, South Africa. Palaeontologia Electronica 18.1.12A, 116. palaeo-electronica.org/content/2015/1026-identification-of-fossil-eggsGoogle Scholar
Kuhn, B.F., Herries, A.I.R., Pickering, R., et al. (2016). Renewed investigations at Taung; 90 years after Australopithecus africanus. Palaeontologia Africana 51, 117.Google Scholar
Kuhn, B.F., Werdelin, L. and Steininger, C. (2017). Fossil Hyaenidae from Cooper’s Cave, South Africa, and the palaeoenvironmental implications. Palaeobiodiversity and Palaeoenvironments 97(2), 355365.Google Scholar
Kullmer, O. (1999). Evolution of African Plio-Pleistocene suids (Suidae; Artiodactyla) based on tooth pattern analysis. Kaupia 9, 134.Google Scholar
Kullmer, O. (2008). The fossil suidae from the Plio-Pleistocene Chiwondo Beds of Northern Malawi. African Journal of Vertebrate Paleontology 208, 208216.Google Scholar
Kullmer, O., Sandrock, O., Schrenk, F. and Bromage, T.G. (1999a). The Malawi Rift: biogeography, ecology and coexistence of Homo and Paranthropus. Anthropologie 37(3), 221231.Google Scholar
Kullmer, O., Sandrock, O., Abel, R., et al. (1999b). The first Paranthropus from the Malawi Rift. Journal of Human Evolution 37, 121127.Google Scholar
Kullmer, O., Sandrock, O., Viola, T.B., et al. (2008). Suids, elephantoids, paleochronology, and paleoecology of the Pliocene hominid site Galili, Somali region, Ethiopia. Palaios 23, 452464.Google Scholar
Kullmer, O., Sandrock, O., Kupczik, K., et al. (2011). New primate remains from Mwenirondo, Chiwondo Beds in northern Malawi. Journal of Human Evolution 61, 617623.Google Scholar
Kuman, K. (1994a). The archaeology of Sterkfontein: preliminary findings on site formation and cultural change. South African Journal of Science 90, 215219.Google Scholar
Kuman, K. (1994b). The archaeology of Sterkfontein – past and present. Journal of Human Evolution 27, 471495.Google Scholar
Kuman, K. (2003). Site formation in the early South African Stone Age sites and its influence on the archaeological record. South African Journal of Science 99, 251254.Google Scholar
Kuman, K. and Clarke, R.J. (1986). Florisbad – new investigations at a Middle Stone Age hominid site in South Africa. Geoarchaeology, 1(2), 103125.Google Scholar
Kuman, K. and Clarke, R.J. (2000). Stratigraphy, artefact industries and hominid associations for Sterkfontein Member 5. Journal of Human Evolution 38, 827847.Google Scholar
Kuman, K., Field, A. and Thackeray, J.F. (1997). Discovery of new artefacts at Kromdraai. South African Journal of Science 93, 187193.Google Scholar
Kuman, K., Inbar, M. and Clarke, R.J. (1999). Palaeoenvironments and cultural sequence of the Florisbad Middle Stone Age hominid site, South Africa. Journal of Archaeological Science 26, 14091425.Google Scholar
Kumar, S., Filipski, A., Swarna, V., Walker, A. and Hedges, S. B. (2005). Placing confidence limits on the molecular age of the human–chimpanzee divergence. Proceedings of the National Academy of Sciences of the United States of America 102(52), 1884218847.Google Scholar
Kunimatsu, Y., Nakatsukasa, M., Sawada, Y., et al. (2007). A new Late Miocene great ape from Kenya and its implications for the origins of African great apes and humans. Proceedings of the National Academy of Sciences 104(49), 1922019225.Google Scholar
Kunimatsu, Y., Nakatsukasa, M., Sawada, Y., et al. (2016). A second hominoid species in the early Late Miocene fauna of Nakali (Kenya). Anthropological Science 124, 7583.Google Scholar
Kunimatsu, Y., Nakatsukasa, M., Sakai, T., et al. (2017a). A newly discovered galagid fossil from Nakali, an early Late Miocene locality of East Africa. Journal of Human Evolution 105, 123126.Google Scholar
Kunimatsu, Y., Sawada, Y., Sakai, T., et al. (2017b). The latest occurrence of the nyanzapithecines from the early Late Miocene Nakali Formation in Kenya, East Africa. Anthropological Science 125, 4551.Google Scholar
Kunneriath, M. and Gaillard, C. (2010). Techno-typological analysis of lithic collections from Sheppard Island and Pniel, Vaal River Valley, South Africa. Annali dell’Università di Ferrara, Museologia Scientifica e Naturalistica 6, 111116.Google Scholar
Kuper, R. and Kröpelin, S. (2006). Climate-controlled Holocene occupation in the Sahara: motor of Africa’s evolution. Science 313, 803807.Google Scholar
Kurashina, H. (1987). Comparison of Gadeb and other Early Stone Age assemblages from Africa south of the Sahara. African Archaeological Review 5, 1928.Google Scholar
Kurtén, B. (1976). Fossil Carnivora from the late Tertiary of Bled Douarah and Cherichira, Tunisia. Notes du Service Géologique de Tunisie 42, 177214.Google Scholar
Kuykendall, K.L. (2012). On the evolutionary development of early hominid molar teeth and the Gondolin Paranthropus molar. In: Reynolds, S.C. and Gallagher, A. (Eds.), African Genesis: Perspectives on Hominin Evolution. Cambridge: Cambridge University Press, pp. 280297.Google Scholar
Kuykendall, K.L., Toich, C.A. and McKee, J.K. (1995). Preliminary analysis of the fauna from Buffalo Cave, northern Transvaal, South Africa. Palaeontologia Africana 32, 2731.Google Scholar
Kyauka, P.S. (1994). A comparative study of the Laetoli Hominid 21 skeleton and its implications for the developmental biology of Australopithecus afarensis. PhD dissertation, University of California, Berkeley.Google Scholar
Kyauka, P.S. and Ndessokia, P. (1990). A new hominid tooth from Laetoli, Tanzania. Journal of Human Evolution 19, 747750.Google Scholar
L’Abbé, E.N., Symes, S.A., Pokines, J.T., et al. (2015). Evidence of fatal skeletal injuries on Malapa Hominins 1 and 2. Scientific Reports 5, 15120.Google Scholar
Lacomba, J.I., Morales, J., Robles, F., Santiesteban, C. and Alberdi, M.T. (1986). Sedimentología y paleontología del yacimiento finimioceno de La Portera (Valencia). Estudios geológicos 42(2–3), 167180.Google Scholar
Lacruz, R. (2007). Enamel microstructure of the hominid KB 5223 from Kromdraai, South Africa. American Journal of Physical Anthropology 132, 175182.Google Scholar
Lacruz, R.S. and Maude, G. (2005). Bone accumulations at brown hyaena (Parahyaena brunnea) den sites in the Makgadikgadi Pans, northern Botswana: taphonomic, behavioral and palaeoecological implications. Journal of Taphonomy 3, 4354.Google Scholar
Lacruz, R.S., Brink, J.S., Hancox, P.J., et al. (2002). Palaeontology and geological context of a middle Pleistocene faunal assemblage from Gladysvale Cave, South Africa. Palaeontologia Africana 38, 99114.Google Scholar
Lague, M.R., Chirchir, H., Green, D.J., et al. (2019a). Cross-sectional properties of the humeral diaphysis of Paranthropus boisei: implications for upper limb function. Journal of Human Evolution 126, 5170.Google Scholar
Lague, M.R., Chirchir, H., Green, D.J., et al. (2019b). Humeral anatomy of the KNM-ER 47000 upper limb skeleton from Ileret, Kenya: implications for taxonomic identification. Journal of Human Evolution 126, 2438.Google Scholar
Lahr, M.M. and Foley, R. (1998). Toward a theory of modern human origins: geography, demography, and diversity in recent human evolution. Yearbook of Physical Anthropology 41, 137176.Google Scholar
Lanzarone, P., Garrison, E., Bobe, R. and Getahun, A. (2016). Examining fluvial stratigraphic architecture using ground-penetrating radar at the Fanta Stream fossil and archaeological site, Central Ethiopia. Geoarchaeology 31, 577591.Google Scholar
Laplace, G. (1956). Découverte de galets taillés (Pebble culture) dans le Quaternaire ancien de Mansourah (Constantine). Comptes Rendus de l’Académie des Sciences 247, 184185.Google Scholar
Larbey, C., Mentzer, S.M., Ligouis, B., Wurz, S. and Jones, M.K. (2019). Cooked starchy food in hearths ca. 120 kya and 65 kya (MIS 5e and MIS 4) from Klasies River Cave, South Africa. Journal of Human Evolution 131, 210227.Google Scholar
Larrasoaña, J., Roberts, A., Rohling, E., Winklhofer, M. and Wehausen, R. (2003). Three million years of monsoon variability over the northern Sahara. Climate Dynamics 21, 689698.Google Scholar
Larrasoaña, J.C., Roberts, A.P. and Rohling, E.J. (2013). Dynamics of green Sahara periods and their role in hominin evolution. PLoS ONE 8(10), e76514.Google Scholar
Laskar, J., Robutel, P., Joutel, F., et al. (2004). A long-term numerical solution for the insolation quantities of the Earth. Astronomy & Astrophysics 428(1), 261285.Google Scholar
Latham, A. and Warr, G. (2008). Makapansgat Limeworks stratigraphy and the singular case of Member X. Palaeontologica Africana 43, 4550.Google Scholar
Latham, A.G, Herries, A.I.R., Quinney, P., Sinclair, A. and Kuykendall, K. (1999). The Makapansgat australopithecine site from a speleological perspective. In: Pollard, A.M. (Ed.), Geoarchaeology: Exploration, Environments, Resources. Royal Geological Society Special Publications 165. London: Royal Geological Society, pp. 6177.Google Scholar
Latham, A.G., Herries, A.I.R. and Kuykendall, K. (2003). The formation and sedimentary infilling of the Limeworks Cave, Makapansgat, South Africa. Palaeontological Africana 39, 6982.Google Scholar
Latham, A.G., Herries, A.I. and Kuykendall, K.L. (2004). The formation and sedimentary infilling of the Limeworks Cave, Makapansgat, South Africa. Geoarchaeology 19, 323342.Google Scholar
Latham, A.G., McKee, J.K. and Tobias, P.V. (2007). Bone breccias, bone dumps, and sedimentary sequences of the western Limeworks, Makapansgat, South Africa. Journal of Human Evolution 52, 388400.Google Scholar
Latham, T.S., Sutton, P.A. and Verosub, K.L. (1992). Non-destructive XRF characterization of basaltic artifacts from Truckee, California. Geoarchaeology 7, 81100.Google Scholar
Latimer, B. (1991). Locomotor adaptations in Australopithecus afarensis: the issue of arboreality. In: Senut, B. and Coppens, Y. (Eds), Origine(s) de la Bipédie chez les Hominidés. Paris: CNRS, pp. 169176.Google Scholar
Latimer, B.M., Ohman, J.C. and Lovejoy, C.O. (1987). Talocrural joint in African hominoids. Implications for Australopithecus afarensis. American Journal of Physical Anthropology 74, 155175.Google Scholar
Lavocat, R. (1952). Sur une faune de mammifères miocènes découverte à Beni Mellal (Atlas marocain). Comptes Rendus de l’Académie des Sciences 235, 189191.Google Scholar
Lazagabaster, I.A., Brophy, J., Sanisidro, O., Pineda-Munoz, S. and Berger, L. (2018a). A new partial cranium of Metridiochoerus (Suidae, Mammalia) from Malapa, South Africa. Journal of African Earth Sciences 145, 4952.Google Scholar
Lazagabaster, I.A., Souron, A., Rowan, J., et al. (2018b). Fossil Suidae (Mammalia, Artiodactyla) from Lee Adoyta, Ledi-Geraru, lower Awash Valley, Ethiopia: implications for late Pliocene turnover and paleoecology. Palaeogeography, Palaeoclimatology, Palaeoecology 504, 186200.Google Scholar
Le Bas, M.J. (1977). Carbonatite–Nephelinite Volcanism: An African Case History. London: John Wiley & Sons.Google Scholar
Le Fur, S., Fara, E., Mackaye, H.T., Vignaud, P. and Brunet, M. (2009). The mammal assemblage of the hominid site TM266 (Late Miocene, Chad Basin): ecological structure and paleoenvironmental implications. Naturwissenschaften 96, 565574.Google Scholar
Leakey, L.S.B. (1928). The Oldoway skull. Nature 121, 499500.Google Scholar
Leakey, L.S.B. (1936). Fossil human remains from Kanam and Kanjera, Kenya Colony. Nature 138, 643643.Google Scholar
Leakey, L.S.B. (1951). Olduvai Gorge: A Report on the evolution of the Hand-Axe Culture in Beds I–IV. Cambridge: Cambridge University Press.Google Scholar
Leakey, L.S.B. (1958). Some East African Pleistocene Suidae. Fossil Mammals of Africa 14, 1133.Google Scholar
Leakey, L.S.B. (1959). A new fossil skull from Olduvai. Nature 189, 491493.Google Scholar
Leakey, L.S.B. (1960). Recent discoveries at Olduvai Gorge. Nature 188, 10501052.Google Scholar
Leakey, L.S.B. (1961a). New finds at Olduvai Gorge. Nature 189, 649650.Google Scholar
Leakey, L.S.B. (1961b). The juvenile mandible from Olduvai. Nature 191, 417418.Google Scholar
Leakey, L.S.B. (1965). Olduvai Gorge. Vol. 1, A Preliminary Report on the Geology and Fauna, 1951–1961. London: Cambridge University Press.Google Scholar
Leakey, L.S.B. (1972). Homo sapiens in the Middle Pleistocene and the evidence of Homo sapiens’ evolution. In: Bordes, F. (Ed.), The Origin of Homo sapiens. Paris: UNESCO, pp. 2529.Google Scholar
Leakey, L.S.B. and Leakey, M.D. (1963). Archaeological Excavations at Olduvai Gorge, Tanzania. Washington, DC: National Geographic Society Research Reports, pp. 179182.Google Scholar
Leakey, L.S.B. and Leakey, M.D. (1964). Recent discoveries of fossil Hominids in Tanganyika: at Olduvai and near Lake Natron. Nature 202, 57.Google Scholar
Leakey, L.S.B., Hopwood, A.T. and Reck, H. (1931). New yields from the Oldoway Bone Beds, Tanganyika Territory. Nature 128, 10751075.Google Scholar
Leakey, L.S.B., Evernden, J.F. and Curtis, G.H. (1961). Age of Bed I, Olduvai Gorge, Tanganyika. Nature 191, 478479.Google Scholar
Leakey, L.S.B., Tobias, P.V. and Napier, J.R. (1964). A new species of the genus Homo from Olduvai Gorge. Nature 202, 79.Google Scholar
Leakey, L.S.B., Savage, R.J.G. and Coryndon, S.C. (1973). Fossil Vertebrates of Africa, Volume 3. London: Academic Press.Google Scholar
Leakey, M.D. (1966). A review of the Oldowan Culture from Olduvai Gorge, Tanzania. Nature 210, 462466.Google Scholar
Leakey, M.D. (1969). Recent discoveries of hominid remains at Olduvai Gorge, Tanzania. Nature 223, 756756.Google Scholar
Leakey, M.D. (1970). New hominid remains and early artefacts from northern Kenya: early artefacts from the Koobi Fora area. Nature, 226, 228230.Google Scholar
Leakey, M.D. (1971). Olduvai Gorge. Vol.3, Excavations in Beds 1 and 2, 1960–1963. London: Cambridge University Press.Google Scholar
Leakey, M.D. (1978). Pliocene footprints at Laetoli, northern Tanzania. Antiquity 52, 133.Google Scholar
Leakey, M.D. (1979). 3.6 million years old footprints in the ashes of time. National Geographic Magazine 155, 446457.Google Scholar
Leakey, M.D. (1981). Tracks and tools. Philosophical Transactions of the Royal Society of London, B, 292, 95102.Google Scholar
Leakey, M.D. (1984). Disclosing the Past. An Autobiography. New York: Doubleday.Google Scholar
Leakey, M.D. (1987a). The hominid footprints: introduction. In: Leakey, M.D. and Harris, J.M. (Eds.), Laetoli: A Pliocene Site in Northern Tanzania. Oxford: Clarendon Press, pp. 490496.Google Scholar
Leakey, M.D. (1987b). Introduction. In: Leakey, M.D. and Harris, J.M. (Eds.), Laetoli: A Pliocene Site in Northern Tanzania. Oxford: Clarendon Press, pp. 122.Google Scholar
Leakey, M.D. (1987c). The Laetoli hominid remains. In: Leakey, M.D. and Harris, J.M. (Eds.), Laetoli: A Pliocene Site in Northern Tanzania. Oxford: Clarendon Press, pp. 108117.Google Scholar
Leakey, M.D. (1987d). Animal prints and trails. In: Leakey, M.D. and Harris, J.M. (Eds.), Laetoli: A Pliocene Site in Northern Tanzania. Oxford: Clarendon Press, pp. 451489.Google Scholar
Leakey, M.D. (1994). Olduvai Gorge. Volume 5, Excavations in Beds III and IV, and the Masek Beds, 1968–1971. Cambridge: Cambridge University Press.Google Scholar
Leakey, M.D. and Harris, J.M. (Eds.) (1987). Laetoli: A Pliocene Site in Northern Tanzania. Oxford: Clarendon Press.Google Scholar
Leakey, M.D. and Hay, R.L. (1979). Pliocene footprints in the Laetolil Beds at Laetoli, northern Tanzania. Nature 278, 317323.Google Scholar
Leakey, M.D. and Roe, D.A. (1994). Olduvai Gorge: Geology and Dating of Beds III, IV and the Masek Beds, 1968–1971. Cambridge: Cambridge University Press.Google Scholar
Leakey, M.D., Clarke, R.J. and Leakey, L.S.B. (1971). New hominid skull from Bed I, Olduvai Gorge, Tanzania. Nature 232, 308312.Google Scholar
Leakey, M.D., Hay, R.L., Curtis, G.H., et al. (1976). Fossil hominids from the Laetolil Beds. Nature 262, 460466.Google Scholar
Leakey, M.G. and Harris, J.M. (2003a). Lothagam: The Dawn of Humanity In Eastern Africa. New York: Columbia University Press.Google Scholar
Leakey, M.G. and Harris, J.M. (2003b). Lothagam: its significance and contributions. In:Lothagam: The Dawn of Humanity in Eastern Africa. New York: Columbia University Press, pp. 625660.Google Scholar
Leakey, M.G. and Leakey, R.E.F. (1976). Further Cercopithecinae (Mammalia, Primates) from the Plio/Pleistocene of East Africa. In: Savage, R.J.G. and Coryndon, S.C. (Eds.), Fossil Vertebrates of Africa, vol. 4. London: Academic Press, pp. 121146.Google Scholar
Leakey, M.G. and Leakey, R.E. (1978). Koobi Fora Research Project, Volume 1: The fossil hominids and an introduction to their context, 1968–1974. In: Leakey, R.E. and Isaac, G.L. (Eds.), Koobi Fora: Researches into Geology, Palaeontology, and Human Origins. Oxford: Clarendon Press, p. 191.Google Scholar
Leakey, M.G. and Walker, A. (2003). The Lothagam hominids. In:Lothagam: The Dawn of Humanity in Eastern Africa. New York: Columbia University Press, pp. 249260.Google Scholar
Leakey, M.G. and Werdelin, L. (2010). Early Pleistocene mammals of Africa: background to dispersal. In Fleagle, J.G., Shea, J.J., Grine, F.E., Baden, A.L. and Leakey, R.E. (Eds.), Out of Africa I. Dordrecht: Springer, pp. 311.Google Scholar
Leakey, M.G., Feibel, C.S., McDougall, I. and Walker, A. (1995). New four-million-year-old hominid species from Kanapoi and Allia Bay, Kenya. Nature 376, 565571.Google Scholar
Leakey, M.G., Feibel, C.S., Bernor, R.L., et al. (1996). Lothagam: a record of faunal change in the Late Miocene of East Africa. Journal of Vertebrate Paleontology 16(3), 556570.Google Scholar
Leakey, M.G., Feibel, C.S., McDougall, I., Ward, C.V. and Walker, A. (1998). New specimens and confirmation of an early age for Australopithecus anamensis. Nature 393, 6266.Google Scholar
Leakey, M.G., Spoor, F., Brown, F.H., et al. (2001). New hominin genus from eastern Africa shows diverse middle Pliocene lineages. Nature 410, 433440.Google Scholar
Leakey, M.G., Teaford, M.F. and Ward, C.V. (2003). Cercopithecidae from Lothagam. In:Lothagam: The Dawn of Humanity in Eastern Africa. New York: Columbia University Press, pp. 201248.Google Scholar
Leakey, M.G., Spoor, F., Dean, M.C., et al. (2012). New fossils from Koobi Fora in northern Kenya confirm taxonomic diversity in early Homo. Nature 488, 201204.Google Scholar
Leakey, R.E.F. (1969). Early Homo sapiens remains from the Omo River region of south-west Ethiopia: faunal remains from the Omo Valley. Nature, 222, 11321133.Google Scholar
Leakey, R.E.F. (1970). New hominid remains and early artefacts from northern Kenya: fauna and artefacts from a new Plio-Pleistocene locality near Lake Rudolf in Kenya. Nature 226, 223224.Google Scholar
Leakey, R.E. (1973). Evidence for an advanced Plio-Pleistocene hominid from East Rudolf, Kenya. Nature 242, 447450.Google Scholar
Leakey, R.E.F. (1976). New hominid fossils from the Koobi Fora Formation in Northern Kenya. Nature 261, 574576.Google Scholar
Lebatard, A.-E., Bourlès, D.L., Duringer, P., et al. (2008). Cosmogenic nuclide dating of Sahelanthropus tchadensis and Australopithecus bahrelghazali: Mio-Pliocene hominids from Chad. Proceedings of the National Academy of Sciences 105, 32263231.Google Scholar
Lebatard, A.E., Bourlès, D.L., Braucher, R., et al. (2010). Application of the authigenic 10Be/9Be dating method to continental sediments: reconstruction of the Mio-Pleistocene sedimentary sequence in the early hominid fossiliferous areas of the northern Chad Basin. Earth and Planetary Science Letters 297(1–2), 5770.Google Scholar
Leblanc, M., Favreau, G., Tweed, S., et al. (2007). Remote sensing for groundwater modelling in large semiarid areas: Lake Chad Basin, Africa. Hydrogeology Journal, 15(1), 97100.Google Scholar
Le Cabec, A., Colard, T., Charabidze, D. et al. (2021). Insights into the palaeobiology of an early Homo infant: multidisciplinary investigation of the GAR IVE hemi-mandible, Melka Kunture, Ethiopia. Scientific Reports, 11, 23087.Google Scholar
Lee, R. and Roberts, D. (1997). Last interglacial (c. 117 kyr) human footprints from South Africa. South African Journal of Science, 93(8), 349350.Google Scholar
Leece, A.B., Kegley, A.D., Lacruz, R.S., et al. (2016). The first hominin from the early Pleistocene paleocave of Haasgat, South Africa. PeerJ, 4, e2024.Google Scholar
Lee-Thorp, J.A. and Beaumont, P.B. (1995). Vegetation and seasonality shifts during the Late Quaternary deduced from 13C/12C ratios of grazers at Equus Cave, South Africa. Quaternary Research 43(3), 426432.Google Scholar
Lee-Thorp, J. and Ecker, M. (2015). Holocene environmental change at Wonderwerk Cave, South Africa: insights from stable light isotopes in ostrich egg shell. African Archaeological Review 32, 793811.Google Scholar
Lee-Thorp, J. and Sponheimer, M. (2003). Three case studies used to reassess the reliability of fossil bone and enamel isotope signals for paleodietary studies. Journal of Anthropological Archaeology, 22(3), 208216.Google Scholar
Lee-Thorp, J.A. and Sponheimer, M. (2013). Hominin ecology from hard-tissue biogeochemistry. In: Ungar, P., Kaye, E., Lee-Thorp, J.A. and Sponheimer, M. (Eds.), Early Hominin Paleoecology. Boulder: University Press of Colorado, pp. 281324.Google Scholar
Lee-Thorp, J.A. and Talma, A.S. (2000). Stable light isotopes and environments in the southern African Quaternary and Late Pliocene. Oxford Monographs on Geology and Geophysics 40, 236251.Google Scholar
Lee-Thorp, J.A. and van der Merwe, N.J. (1993). Stable carbon isotope studies of Swartkrans fossils. In: Brain, C.K. (Ed.), Swartkrans: A Cave’s Chronicle of Early Man. Transvaal Museum Monograph No. 8. Pretoria: Transvaal Museum, pp. 251256.Google Scholar
Lee-Thorp, J.A., van der Merwe, N.J. and Brain, C.K. (1989). Isotopic evidence for dietary differences between two extinct baboon species from Swartkrans. Journal of Human Evolution 18, 183190.Google Scholar
Lee-Thorp, J., van der Merwe, N. and Brain, C. (1994). Diet of Australopithecus robustus at Swartkrans from stable carbon isotope analysis. Journal of Human Evolution 27, 361372.Google Scholar
Lee-Thorp, J.A., Manning, L. and Sponheimer, M. (1997). Problems and prospects for carbon isotope analysis of very small samples of fossil tooth enamel. Bulletin de la Société géologique de France 168, 767773.Google Scholar
Lee-Thorp, J.A., Thackeray, J.F. and van der Merwe, N. (2000). The hunters and the hunted revisited. Journal of Human Evolution 39, 565576.Google Scholar
Lee-Thorp, J.A., Sponheimer, M. and Van der Merwe, N.J. (2003). What do stable isotopes tell us about Hominid dietary and ecological niches in the Pliocene? International Journal of Osteoarchaeology 13, 104113.Google Scholar
Lee-Thorp, J.A., Sponheimer, M. and Luyt, J. (2007). Tracking changing environments using stable carbon isotopes in fossil tooth enamel: an example from the South African hominin sites. Journal of Human Evolution 53(5), 595601.Google Scholar
Lee-Thorp, J., Likius, A., Mackaye, H.T., et al. (2012). Isotopic evidence for an early shift to C4 resources by Pliocene hominins in Chad. Proceedings of the National Academy of Sciences 109, 2036920372.Google Scholar
Lefèvre, D. (2000). Du continent à l’océan. Morphostratigraphie et paléogéographie du Quaternaire du Maroc atlantique. Le modèle casablancais. Thèse HDR, vol. 3(2), pp. 100308, Univ. Montpellier 3.Google Scholar
Lefèvre, D. and Raynal, J.P. (2002). Les formations plio-pléistocènes de Casablanca et la chronostratigraphie du Quaternaire marin du Maroc revisitées. Quaternaire 13, 921.Google Scholar
Lefèvre, D. (2016). Le Quaternaire d’Oulad Hamida. In: Raynal, J.-P. and Mohib, A. (Eds.), Préhistoire de Casablanca. 1 ‒ La Grotte des Rhinocéros (fouilles 1991 et 1996). Villes et Sites Archéologiques du Maroc, volume 6. Rabat: Ministère de la Culture, INSAP, pp. 4559.Google Scholar
Lefèvre, D., Texier, J.P., Raynal, J.P., Occhietti, S. and Evin, J. (1994). Enregistrements-réponses des variations climatiques du Pléistocène supérieur et de l’Holocène sur le littoral de Casablanca (Maroc). Quaternaire 5, 173180.Google Scholar
Legendre, P. and De Cáceres, M. (2013). Beta diversity as the variance of community data: dissimilarity coefficients and partitioning. Ecology Letters 16(8), 951963.Google Scholar
Legendre, P. and Gallagher, E. D. (2001). Ecologically meaningful transformations for ordination of species data. Oecologia 129(2), 271280.Google Scholar
Legendre, S. (1989). Les communautés de mammifères du Paléogène (Eocène supérieur et Oligocène) d’Europe occidentale: structures, milieux et evolution. Münchner Geowissenschaftliche Abhandlungen (A) 16, 1110.Google Scholar
Lehmann, S.B., Braun, D.R., Dennis, K.J., et al. (2016). Stable isotopic composition of fossil mammal teeth and environmental change in southwestern South Africa during the Pliocene and Pleistocene. Palaeogeography, Palaeoclimatology, Palaeoecology, 457, 396408.Google Scholar
Lehmann, T. (2004). Fossil aardvark (Orycteropus) from Swartkrans Cave, South Africa. South African Journal of Science, 100(5), 311314.Google Scholar
Lehmann, T., Vignaud, P., Mackaye, H.T. and Brunet, M. (2004). A fossil aardvark (Mammalia, Tubulidentata) from the lower Pliocene of Chad. Journal of African Earth Sciences 40, 201217.Google Scholar
Lehmann, T., Vignaud, P., Likius, A. and Brunet, M. (2005). A new species of Orycteropodidae (Mammalia, Tubulidentata) in the Mio-Pliocene of northern Chad. Zoological Journal of the Linnean Society 143, 109131.Google Scholar
Leichliter, J.N. (2018). Early Hominin environments in Southern Africa: a micromammalian perspective. Doctoral dissertation, University of Colorado at Boulder.Google Scholar
Leichliter, J.N., Sponheimer, M., Avenant, N.L., et al. (2016). Small mammal insectivore stable carbon isotope compositions as habitat proxies in a South African savanna ecosystem. Journal of Archaeological Science: Reports 8, 335345.Google Scholar
Leichliter, J., Sandberg, P., Passey, B., et al. (2017). Stable carbon isotope ecology of small mammals from the Sterkfontein Valley: implications for habitat reconstruction. Palaeogeography, Palaeoclimatology, Palaeoecology, 485, 5767.Google Scholar
Lemorini, C., Plummer, T.W., Braun, D.R., et al. (2014). Old stones’ song: use-wear experiments and analysis of the Oldowan quartz and quartzite assemblage from Kanjera South (Kenya). Journal of Human Evolution 72, 1025.Google Scholar
Lemorini, C., Plummer, T.W., Braun, D.R., et al. (2019). Old stones’ song: use-wear experiments and analysis of the Oldowan quartz and quartzite assemblage from Kanjera South (Kenya). Archaeological and Anthropological Sciences 11, 47294754.Google Scholar
Lennox, S.J. and Bamford, M. (2015). Use of wood anatomy to identify poisonous plants: charcoal of Spirostachys africana. South African Journal of Science, 111(3–4), 19.Google Scholar
Lennox, S.J. and Wadley, L. (2019). A charcoal study from the Middle Stone Age, 77,000 to 65,000 years ago, at Sibudu, KwaZulu-Natal. Transactions of the Royal Society of South Africa, 74(1), 3854.Google Scholar
Lepre, C.J. (2014). Early Pleistocene lake formation and hominin origins in the Turkana–Omo rift. Quaternary Science Reviews 102, 181191.Google Scholar
Lepre, C.J., Roche, H., Kent, D.V., et al. (2011). An earlier origin for the Acheulian. Nature 477, 8285.Google Scholar
Leroy, S. and Dupont, L. (1994). Development of vegetation and continental aridity in northwestern Africa during the Late Pliocene: the pollen record of ODP site 658. Palaeogeography, Palaeoclimatology, Palaeoecology 109, 295316.Google Scholar
Leroy, S.A.G. and Dupont, L.M. (1997). Marine palynology of the ODP 658 (N-W Africa) and its contribution to the stratigraphy of Late Pliocene. Geobios 30, 351359.Google Scholar
Lesur, J., Vigne, J.-D. and Gutherz, X. (2007). Exploitation of wild mammals in South-west Ethiopia during the Holocene (4000 BC–500 AD): the finds from Moche Borago shelter (Wolayta). Environmental Archaeology 12, 139159.Google Scholar
Lesur, J., Faith, J.T., Bon, F., et al. (2016). Paleoenvironmental and biogeographic implications of terminal Pleistocene large mammals from the Ziway-Shala Basin, Main Ethiopian Rift, Ethiopia. Palaeogeography, Palaeoclimatology, Palaeoecology 449, 567579.Google Scholar
Levin, N.E. (2013). Compilation of East Africa Soil Carbonate Stable Isotope Data. Integrated Earth Data Applications. doi:10.1594/IEDA/100231.Google Scholar
Levin, N.E. (2015). Environment and climate of early human evolution. Annual Review of Earth and Planetary Sciences 43, 405429.Google Scholar
Levin, N.E., Quade, J., Simpson, S.W., Semaw, S. and Rogers, M. (2004). Isotopic evidence for Plio-Pleistocene environmental change at Gona, Ethiopia. Earth and Planetary Science Letters 219, 93110.Google Scholar
Levin, N.E., Cerling, T.E., Passey, B., Harris, J. and Ehleringer, J. (2006). A stable isotope aridity index for terrestrial environments. Proceedings of the National Academy of Science 103, 1120111205.Google Scholar
Levin, N.E., Simpson, S.W., Quade, J., Cerling, T.E. and Frost, S.R. (2008). Herbivore enamel carbon isotopic composition and the environmental context of Ardipithecus at Gona, Ethiopia. In: Quade, J. and Wynn, J.G. (Eds.), The Geology of Early Humans in the Horn of Africa. Geological Society of America Special Paper 446.Boulder: Geological Society of America, pp. 215234.Google Scholar
Levin, N.E., Brown, F.H., Behrensmeyer, A.K., Bobe, R. and Cerling, T.E. (2011). Paleosol carbonates from the Omo Group: isotopic records of local and regional environmental change in East Africa. Palaeogeography, Palaeoclimatology, Palaeoecology 307, 7589.Google Scholar
Levin, N.E., Haile-Selassie, Y., Frost, S.R. and Saylor, B.Z. (2015). Dietary change among hominins and cercopithecids in Ethiopia during the early Pliocene. Proceedings of the National Academy of Sciences of the United States of America 117, 12304–1230.Google Scholar
Levin, S.A., Carpenter, S.R., Godfray, H.C.J., et al. (2009). The Princeton Guide to Ecology. Princeton: Princeton University Press, p. 848.Google Scholar
Levinson, M. (1985). Are fossil rodents useful in palaeo-ecological interpretations. Annals of the Geological Survey of South Africa 19, 5364.Google Scholar
Lewin, R. (1983). Ethiopia halts prehistoric research. Science 219, 147149.Google Scholar
Lewis, A.R., Marchant, D.R., Ashworth, A.C., Hemming, S.R. and Machlus, M.L. (2007). Major middle Miocene global climate change: evidence from East Antarctica and the Transantarctic Mountains. Geological Society of America Bulletin 119, 14491461.Google Scholar
Lewis, M.E. (1997). Carnivoran paleoguilds of Africa: implications for hominid food procurement strategies. Journal of Human Evolution 32(2–3), 257288.Google Scholar
Lézine, A.-M. (1989). Late Quaternary vegetation and climate of the Sahel. Quaternary Research 32, 317334.Google Scholar
Lézine, A.-M., Casanova, J. and Hillaire-Marce, C. (1990). Across an early Holocene humid phase in western Sahara: pollen and isotope stratigraphy. Geology 18, 264267.Google Scholar
Lihoreau, F., Boisserie, J.-R., Viriot, L., et al. (2006). Anthracothere dental anatomy reveals a late Miocene Chado-Libyan bioprovince. Proceedings of the National Academy of Sciences, 103(23), 87638767.Google Scholar
Lihoreau, F., Boisserie, J.R., Blondel, C., et al. (2014). Description and palaeobiology of a new species of Libycosaurus (Cetartiodactyla, Anthracotheriidae) from the Late Miocene of Toros-Menalla, northern Chad. Journal of Systematic Palaeontology 12, 761798.Google Scholar
Lihoreau, F., Sarr, R., Chardon, D., et al. (2021). A fossil terrestrial fauna from Tobène (Senegal) provides a unique early Pliocene window in western Africa. Gondwana Research 99, 2135.Google Scholar
Linseele, V., Van Neer, W., Thys, S., et al. (2014). New archaeozoological data from the Fayum “Neolithic” with a critical assessment of the evidence for early stock keeping in Egypt. PLoS ONE 9, e108517.Google Scholar
Lisiecki, L.E. and Raymo, M.E. (2007). Plio-Pleistocene climate evolution: trends and transitions in glacial cycle dynamics. Quaternary Science Reviews 26, 5669.Google Scholar
Livingston, D. and Kingdon, J. (2013). The evolution of a continent: geography and geology. In: Kingdon, J., Happold, D.C., Butynski, T.M., et al. (Eds.), Mammals of Africa. London: Bloomsbury Publishing, pp. 2742.Google Scholar
Lockwood, C.A. and Tobias, P.V. (1999). A large male hominin cranium from Sterkfontein, South Africa, and the status of Australopithecus. Journal of Human Evolution 36, 637685.Google Scholar
Lockwood, C.A., Kimbel, W.H. and Johanson, D.C. (2000). Temporal trends and metric variation in the mandibles and dentition of Australopithecus afarensis. Journal of Human Evolution 39, 2355.Google Scholar
Loftus, E., Roberts, P. and Lee-Thorp, J.A. (2016). An isotopic generation: four decades of stable isotope analysis in African archaeology. Azania: Archaeological Research in Africa 51(1), 88114.Google Scholar
Loget, N., Driessche, J.V.D. and Davy, P. (2005). How did the Messinian salinity crisis end? Terra Nova, 17(5), 414419.Google Scholar
Lombard, M., Malmström, H., Schlebusch, C., et al. (2019). Genetic data and radiocarbon dating question Plovers Lake as a Middle Stone Age hominin-bearing site. Journal of Human Evolution 131, 203209.Google Scholar
Longinelli, A. (1984). Oxygen isotopes in mammal bone phosphate: a new tool for paleohydrological and paleoclimatological research. Geochimica et Cosmochimica Acta 48, 385390.Google Scholar
López-Martínez, N., Likius, A., Mackaye, H.T., Vignaud, P. and Brunet, M. (2007). A new Lagomorph from the Late Miocene of Chad (Central Africa). [Un nuevo Lagomorfo del Mioceno superior de Chad (África central).] Revista Española de Paleontología 22(1), 120.Google Scholar
López-Sáez, J.-A. and Domínguez-Rodrigo, M. (2009). Palynology of OGS-6a and OGS-7, two new 2.6 Ma archaeological sites from Gona, Afar, Ethiopia: insights on aspects of Late Pliocene habitats and the beginnings of stone-tool use. Geobios 42, 503511.Google Scholar
Lorenzen, E.D., Heller, R. and Siegismund, H.R. (2012). Comparative phylogeography of African savannah ungulates. Molecular Ecology 21, 36563670.Google Scholar
Lotter, M.G. and Kuman, K. (2018). The Acheulean in South Africa, with announcement of a new site (Penhill Farm) in the lower Sundays River Valley, Eastern Cape Province, South Africa. Quaternary International 480, 4365.Google Scholar
Lotter, M.G., Gibbon, R.J., Kuman, K., et al. (2016). A geoarchaeological study of the middle and upper Pleistocene levels at Canteen Kopje, Northern Cape Province, South Africa. Geoarchaeology 31(4), 304323.Google Scholar
Louchart, A. (2011). Aves. In: Harrison, T. (Ed.), Paleontology and Geology of Laetoli: Human Evolution in Context. Vol. 2: Fossil Hominins and the Associated Fauna. Dordrecht: Springer, pp. 505533.Google Scholar
Louchart, A. (2014). Fossil birds of the Konso Formation. In: Suwa, G., Beyene, Y. and Asfaw, B. (Eds.), The Konso-Gardula Research Project Volume 1. Paleonotological Collections: Background and Fossil Aves, Cercopithecidae, and Suidae. Tokyo: The University Museum, The University of Tokyo, Bulletin, 47, 2539.Google Scholar
Louchart, A., Mourer-Chauviré, C., Vignaud, P., Taisso MacKaye, H. and Brunet, M. (2005a). A finfoot from the Late Miocene of Toros Menalla (Chad, Africa): palaeobiogeographical and palaeoecological implications. Palaeogeography, Palaeoclimatology, Palaeoecology 222, 19.Google Scholar
Louchart, A., Vignaud, P., Mackaye, H.T. and Brunet, M. (2005b). A new swan (Aves: Anatidae) in Africa, from the latest Miocene of Chad and Libya. Journal of Vertebrate Paleontology 25, 384392.Google Scholar
Louchart, A., Haile-Selassie, Y., Vignaud, P., Likius, A. and Brunet, M. (2008). Fossil birds from the late Miocene of Chad and Ethiopia and zoogeographical implications. Oryctos 7, 147167.Google Scholar
Louchart, A., Wesselman, H., Blumenschine, R.J., et al. (2009). Taphonomic, avian, and small-vertebrate indicators of Ardipithecus ramidus habitat. Science, 326, 66.Google Scholar
Louys, J. (2012). Paleontology in Ecology and Conservation. New York: Springer.Google Scholar
Louys, J., Meloro, C., Elton, S., Ditchfield, P. and Bishop, L.C. (2015). Analytical framework for reconstructing heterogeneous environmental variables from mammal community structure. Journal of Human Evolution 78, 111.Google Scholar
Lovejoy, C. (2009). Reexamining human origins in light of Ardipithecus ramidus. Science 326(5949), 74.Google Scholar
Lovejoy, C.O., Meindl, R.S., Ohman, J.C., Heiple, K.G. and White, T.D. (2002). The Maka femur and its bearing on the antiquity of human walking: applying contemporary concepts of morphogenesis to the human fossil record. American Journal of Physical Anthropology 119, 97133.Google Scholar
Lucas, P.W., Omar, R., Al-Fadhalah, K., 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
Lüdecke, T., Schrenk, F., Thiemeyer, H., et al. (2016a). Persistent C3 vegetation accompanied Plio-Pleistocene hominin evolution in the Malawi Rift (Chiwondo Beds, Karonga Basin, East African Rift System). Journal of Human Evolution 90, 163175.Google Scholar
Lüdecke, T., Mulch, A., Kullmer, O., et al. (2016b). Stable isotope dietary reconstructions of herbivore enamel reveal heterogenous savanna ecosystems in the Plio-Pleistocene Malawi Rift. Paleogeography, Paleoclimatology, Paleoecology 459, 170181.Google Scholar
Lüdecke, T., Kullmer, O., Wacker, U., et al. (2018). Dietary versatility of Early Pleistocene hominins. Proceedings of the National Academy of Sciences 115(52), 1333013335.Google Scholar
Ludwig, A.J. and Reynolds, J.F. (1988). Statistical Ecology: A Primer on Methods and Computing. New York: Wiley.Google Scholar
Luiselli, L., Politano, E. and Angelici, F.M. (2000). Ecological correlates of the distribution of terrestrial and freshwater Chelonians in the Niger Delta, Nigeria: a biodiversity assessment with conservation implications. Revue D’Ecologie – La Terre et la Vie 55, 323.Google Scholar
Luiselli, L., Akani, G.C., Ebere, N., et al. (2011). Food habits of a pelomedusid turtle, Pelomedusa subrufa, in tropical Africa (Nigeria): the effects of sex, body size, season, and site. Chelonian Conservation and Biology 10, 138144.Google Scholar
Lukich, V., Porat, N., Faershtein, G., Cowling, S. and Chazan, M. (2019). New chronology and stratigraphy for Kathu Pan 6, South Africa. Journal of Paleolithic Archaeology 2(3), 235257.Google Scholar
Lukich, V., Cowling, S. and Chazan, M. (2020). Palaeoenvironmental reconstruction of Kathu Pan, South Africa, based on sedimentological data. Quaternary Science Reviews 230, 106153.Google Scholar
Lupien, R.L., Russell, J.M., Yost, C.L., et al. (2019). Vegetation change in the Baringo Basin, East Africa across the onset of Northern Hemisphere glaciation 3.3–2.6 Ma. Palaeogeography, Palaeoclimatology, Palaeoecology 570, 109426.Google Scholar
Luque, L., Alcalá, L. and Domínguez-Rodrigo, M. (2009). The Peninj Group: tectonics, volcanism, and sedimentary paleoenvironments during the Lower Pleistocene in the Lake Natron Basin (Tanzania). In: Domínguez-Rodrigo, M., Alcalá, L. and Luque, L. (Eds.), Peninj. A Research Project on Human Origins (1995–2005). Oxford: Oxbow, pp. 1548.Google Scholar
Luyt, C.J. (2001). Revisiting palaeoenvironments from the hominid-bearing Plio-Pleistocene sites: new isotopic evidence from Sterkfontein. Unpublished masters thesis. University of Cape Town.Google Scholar
Luyt, C.J. and Lee-Thorp, J.A. (2003). Carbon isotope ratios of Sterkfontein fossils indicate a marked shift to open environments c. 1.7 Myr ago. South African Journal of Science 99, 271273.Google Scholar
Luz, B., Kolodny, Y. and Horowitz, M. (1984). Fractionation of oxygen isotopes between mammalian bone phosphate and environmental drinking water. Geochimica et Cosmochimica Acta 48, 16891693.Google Scholar
Lyman, R.L. (1994). Vertebrate Taphonomy. New York: Cambridge University Press.Google Scholar
Lyman, R.L. (2006). Paleozoology in the service of conservation biology. Evolutionary Anthropology 15, 1119.Google Scholar
Lyman, R.L. (2010). What taphonomy is, what it isn’t, and why taphonomists should care about the difference. Journal of Taphonomy 8, 116.Google Scholar
Lyman, R.L. and O’Brien, M.J. (1987). Plow-zone zooarchaeology: fragmentation and identifiability. Journal of Field Archaeology 14, 493498.Google Scholar
Macchiarelli, R., Bondioli, L., Coppa, A., et al. (2002). The one-million-year-old human remains from the Danakil (Afar) depression of Eritrea. American Journal of Physical Anthropology Supplement 34, 104 (abstract).Google Scholar
Macchiarelli, R., Bondioli, L., Falk, D., et al. (2004a). Early Pliocene hominid tooth from Galili, Somali Region, Ethiopia. Collegium Anthropologicum 28, 6576.Google Scholar
Macchiarelli, R., Bondioli, L., Chech, M., et al. (2004b). The late Early Pleistocene Human Remains from Buia Danakil Depression, Eritrea. Rivista Italiana di Stratigrafia e Paleontologia 110 (Supplement), 133144.Google Scholar
Macchiarelli, R., Bondioli, L., Coppa, A., et al. (2007). Extensive variation in the east African Early Pleistocene human fossil record: additional evidence from Uadi Aalad, Eritrea. Bulletins et Memoires de la Société d’Anthropologie de Paris 19, 279280 (abstract).Google Scholar
Macchiarelli, R., Bondioli, L., Bruner, E., et al. (2014). The 1 Ma old human assemblage from the Homo site at Uadi Aalad, Buia (Danakil Depression of Eritrea): an updated record. The African Human Fossil Record (Toulouse, 26–27 September).Google Scholar
MacFadden, B.J. (1998). Tale of two rhinos: isotopic ecology, paleodiet, and niche differentiation of Aphelops and Teleoceras from the Florida Neogene. Paleobiology 24(2), 274286.Google Scholar
MacFadden, B.J. (2005). Diet and habitat of toxodont megaherbivores (Mammalia, Notoungulata) from the late Quaternary of South and Central America. Quaternary Research 64(2), 113124.Google Scholar
MacGregor, D. (2015). History of the development of the East African Rift System: a series of interpreted maps through time. Journal of African Earth Sciences 101, 232252.Google Scholar
Macho, G.A. (2014). An ecological and behavioral approach to hominin evolution during the Pliocene. Quaternary Science Reviews 96, 2331.Google Scholar
MacIntyre, R.M., Mitchell, J.G. and Dawson, J.B. (1974). Age of fault movements in Tanzanian sector of East African Rift System. Nature 247, 354356.Google Scholar
MacLatchy, L., DeSilva, J., Sanders, W.J. and Wood, B. (2010). Hominini. In: Werdelin, L. and Sanders, W.J. (Eds.), Cenozoic Mammals of Africa. Berkeley: University of California Press, pp. 471540.Google Scholar
Madella, M., Alexandre, A. and Ball, T. (2005). International code for phytolith nomenclature 1.0. Annals of Botany 96, 253260.Google Scholar
Madden, R.H. (2014). Hypsodonty in Mammals: Evolution, Geomorphology, and the Role of Earth Surface Processes. Cambridge: Cambridge University Press.Google Scholar
Maddock, L. (1979). The migration and grazing succession. In: Sinclair, A.R.E. and Norton-Griffiths, M. (Eds.), Serengeti. Chicago: Chicago University Press.Google Scholar
Magill, C.R., Ashley, G.M. and Freeman, K.H. (2013a). Ecosystem variability and early human habitats in eastern Africa. Proceedings of the National Academy of Sciences 110, 11671174.Google Scholar
Magill, C.R., Ashley, G.M. and Freeman, K.H. (2013b). Water, plants, and early human habitats in eastern Africa. Proceedings of the National Academy of Sciences 110, 11751180.Google Scholar
Maglio, V.J. (1970). Four new species of Elephantidae from the Plio-Pleistocene of northwestern Kenya. Breviora 341, 143.Google Scholar
Maglio, V.J. (1973). Origin and evolution of the Elephantidae. Transactions of the American Philosophical Society 63, 1149.Google Scholar
Magori, C.C. and Day, M.H. (1983). Laetoli Hominid 18: an early Homo sapiens skull. Journal of Human Evolution 12, 747753.Google Scholar
Maguire, J.M. (1979). A taxonomic and ecological study of the living and fossil Hystricidae with particular reference to southern Africa. PhD thesis, University of the Witwatersrand.Google Scholar
Maguire, J.M. (1985). Recent geological, stratigraphic and palaeontological studies at Makapansgat Limeworks. In: Tobias, P.V. (Ed.), Hominid Evolution: Past, Present and Future. New York: Alan R. Liss, pp. 151164.Google Scholar
Maguire, J.M., Pemberton, D. and Collett, M.H. (1980). The Makapansgat Limeworks grey breccia: Hominids, hyenas or hillwash? Palaeontologica Africana 23, 7598.Google Scholar
Maguire, J.M., Schrenk, F. and Stanistreet, I.G. (1985). The lithostratigraphy of the Makapansgat Limeworks Australopithecine site: some matters arising. Annals of the Geological Survey of South Africa 19, 3751.Google Scholar
Magurran, A.E. (1988). Ecological Diversity and its Measurement. Princeton: Princeton University Press.Google Scholar
Maier, W. (1970). New fossil Cercopithecoidea from the Pleistocene cave deposits of the Makapansgat limeworks, South Africa. Palaeontologica Africana 12, 69107.Google Scholar
Maier, W. (1972). The first complete skull of Simopithecus [Theropithecus] darti from Makapansgat, SA, and its systematic position. Journal of Human Evolution 1, 395405.Google Scholar
Maitima, J.M. (1991). Vegetation response to climatic change in the Central Rift Valley, Kenya. Quaternary Research 35, 234245.Google Scholar
Makinouchi, T., Ishida, S., Sawada, Y., et al. (1992). Geology of the Sinda-Mohari Region, Haut-Zaïre Province, Eastern Zaïre. African Studies, Monographs 17, 318.Google Scholar
Malan, B.D. and Cooke, H.B.S. (1941). A preliminary account of the Wonderwerk Cave, Kuruman. South African Journal of Science 37, 300312.Google Scholar
Malan, B.D. and Wells, L.H. (1943). A further report on the Wonderwerk Cave, Kuruman. South African Journal of Science 40, 258270.Google Scholar
Mallon, J.C., Henderson, D.M., McDonough, C.M. and Loughry, W. J. (2018). A ‘bloat-and-float’ taphonomic model best explains the upside-down preservation of ankylosaurs. Palaeogeography, Palaeoclimatology, Palaeoecology 497, 117127.Google Scholar
Maloiy, G.M.O., Rugangazi, B.M. and Clemens, E.T. (1988). Physiology of the dik-dik antelope. Comparative Biochemistry and Physiology 91A, 18.Google Scholar
Manega, P.C. (1993). Geochronology, geochemistry and isotopic study of the Plio-Pleistocene hominid sites and the Ngorongoro volcanic highland in northern Tanzania. Unpublished PhD, University of Chicago, Boulder.Google Scholar
Manegold, A. (2013). Two new parrot species (Psittaciformes) from the early Pliocene of Langebaanweg, South A frica, and their palaeoecological implications. Ibis 155(1), 127139.Google Scholar
Manegold, A. and Louchart, A. (2012). Biogeographic and paleoenvironmental implications of a new woodpecker species (Aves, Picidae) from the early Pliocene of South Africa. Journal of Vertebrate Paleontology 32(4), 926938.Google Scholar
Manegold, A., Pavia, M. and Haarhoff, P. (2014). A new species of Aegypius vulture (Aegypiinae, Accipitridae) from the early Pliocene of South Africa. Journal of Vertebrate Paleontology 34(6), 13941407.Google Scholar
Mannion, P.D., Upchurch, P., Carrano, M.T. and Barrett, P.M. (2011). Testing the effect of the rock record on diversity: a multidisciplinary approach to elucidating the generic richness of sauropodomorph dinosaurs through time. Biological Reviews 86(1), 157181.Google Scholar
Manthi, F.K. and Winkler, A.J. (2020). Rodents and other terrestrial small mammals from Kanapoi, north-western Kenya. Journal of Human Evolution 140, 102694.Google Scholar
Manthi, F.K., Brown, F.H., Plavcan, M.J. and Werdelin, L. (2018). Gigantic lion, Panthera leo, from the Pleistocene of Natodomeri, eastern Africa. Journal of Paleontology 92, 305312.Google Scholar
Manthi, F.K., Cerling, T.E., Chritz, K.L. and Blumenthal, S.A. (2020a). Diets of mammalian fossil fauna from Kanapoi, northwestern Kenya. Journal of Human Evolution 140, 102338.Google Scholar
Manthi, F.K., Plavcan, J.M. and Ward, C.V. (2020b). Introduction to special issue Kanapoi: Paleobiology of a Pliocene Hominin Site in Kenya. Journal of Human Evolution 140, 102718.Google Scholar
Marçais, J. (1934). Découverte de restes humains fossiles dans les grès quaternaires de Rabat (Maroc). L’Anthropologie 44, 579583.Google Scholar
Marean, C.W. (1991). Measuring the post-depositional destruction of bone in archaeological assemblages. Journal of Archeological Science 18, 677694.Google Scholar
Marean, C.W. (1992a). Hunter to herder: the large mammal remains from Enkapune Ya Muto rockshelter (Central Rift, Kenya). African Archaeological Review 10, 65127.Google Scholar
Marean, C.W. (1992b). Implications of late Quaternary mammalian fauna from Lukenya Hill (south-central Kenya) for paleoenvironmental change and faunal extinctions. Quaternary Research 37, 239255.Google Scholar
Marean, C.W. (1997). Hunter-gatherer foraging strategies in tropical grasslands: model building and testing in the East African Middle and Later Stone Age. Journal of Anthropological Archaeology 16, 189225.Google Scholar
Marean, C.W. (2010). Pinnacle Point Cave 13B (Western Cape Province, South Africa) in context: the Cape floral kingdom, shellfish, and modern human origins. Journal of Human Evolution 59(3–4), 425443.Google Scholar
Marean, C.W. and Assefa, Z. (2005). The Middle and Upper Pleistocene African record for the biological and behavioral origins of modern humans. In: Stahl, A. (Ed.), African Archaeology: A Critical Introduction. Malden: Blackwell, pp. 93129.Google Scholar
Marean, C.W. and Gifford-Gonzalez, D. (1991). Late Quaternary extinct ungulates of East Africa and palaeoenvironmental implications. Nature 350, 418420.Google Scholar
Marean, C.W., Nilssen, P.J., Brown, K., Jerardino, A. and Stynder, D. (2004). Paleoanthropological investigations of Middle Stone Age sites at Pinnacle Point, Mossel Bay (South Africa): archaeology and hominid remains from the 2000 field season. Paleoanthropology, 2(1).Google Scholar
Marean, C.W., Bar-Matthews, M., Bernatchez, J., et al. (2007). Early human use of marine resources and pigment in South Africa during the Middle Pleistocene. Nature 449(7164), 905908.Google Scholar
Marean, C.W., Anderson, R.J., Bar‐Matthews, M., et al. (2015). A new research strategy for integrating studies of paleoclimate, paleoenvironment, and paleoanthropology. Evolutionary Anthropology: Issues, News, and Reviews 24(2), 6272.Google Scholar
Mares, M.A. (1992). Neotropical mammals and the myth of Amazonian biodiversity. Science 255, 976979.Google Scholar
Marin-Monfort, M.D., García-Morato, S., Andrews, P., et al. (2022). The owl that never left! Taphonomy of Earlier Stone Age small mammal assemblages from Wonderwerk Cave (South Africa). Quaternary International 614,111–125.Google Scholar
Marker, M.E. and Gamble, F.M. (1987). Karst in southern Africa. Endins: publicació d’espeleologia 13, 9398.Google Scholar
Marks, A.E., Peters, J. and Van Neer, W. (1987). Late Pleistocene and early Holocene occupations in the Upper Atbara River Valley, Sudan. In: Close, A.E. (Ed.), Prehistory of Arid North Africa: Essays in Honor of Fred Wendorf. Dallas: Southern Methodist University Press, pp. 137162.Google Scholar
Marshall, C.R. (1990). Confidence intervals on stratigraphic ranges. Paleobiology 16, 110.Google Scholar
Marshall, C.R. (1994). Confidence intervals on stratigraphic ranges: partial relaxation of the assumption of randomly distributed fossil horizons. Paleobiology 20(4), 459469.Google Scholar
Marshall, C.R. (1997). Confidence intervals on stratigraphic ranges with nonrandom distributions of fossil horizons. Paleobiology 23(2), 165173.Google Scholar
Martin, F., Plastiras, C.A., Merceron, G., Souron, A. and Boisserie, J.R. (2018). Dietary niches of terrestrial cercopithecines from the Plio-Pleistocene Shungura Formation, Ethiopia: evidence from dental microwear texture analysis. Scientific Reports 8(1), 113.Google Scholar
Martin, J.W. and Trautwein, S. (2003). Fossil crabs (Crustacea, Decapoda, Brachyura) from Lothagam. In: Leakey, M.G. and Harris, J.M. (Eds.), Lothagam: The Dawn of Humanity in Eastern Africa. New York: Columbia University Press, pp. 6774.Google Scholar
Martínez-Navarro, B. (2010). Early Pleistocene faunas of Eurasia and hominin dispersals. In: Fleagle, J.G., Shea, J.J., Grine, F.E., Baden, A.L. and Leakey, R.E. (Eds.), Out of Africa I: The First Hominin Colonization of Eurasia. Dordrecht: Springer, pp. 207224.Google Scholar
Martínez-Navarro, B., Rook, L., Segid, A., et al. (2004). The large fossil mammals from Buia (Eritrea): systematics, biochronology and paleoenvironments. Rivista Italiana di Paleontologia e Stratigrafia 110, 6188.Google Scholar
Martínez-Navarro, B., Rook, L., Papini, M. and Libsekal, Y. (2010). A new species of bull from the Early Pleistocene paleoanthropological site of Buia (Eritrea): parallelism on the dispersal of the genus Bos and the Acheulian culture. Quaternary International 212, 169175.Google Scholar
Martínez-Navarro, B., Karoui-Yaakoub, N., Oms, O., et al. (2014). The early Middle Pleistocene archeopaleontological site of Wadi Sarrat (Tunisia) and the earliest record of Bos primigenius. Quaternary Science Reviews 90, 3746.Google Scholar
Martini, F., Libsekal, Y., Filippi, O., et al. (2004). Characterization of lithic complexes from Buia (Dandiero Basin, Danakil Depression, Eritrea). Rivista Italiana di Paleontologia e Stratigrafia 110 (Supplement), 99132.Google Scholar
Martyn, J. (1967). Pleistocene deposits and new fossil localities in Kenya. Nature 215(5100), 476.Google Scholar
Marzocchi, A., Lunt, D.J., Flecker, R., et al. (2015). Orbital control on late Miocene climate and the North African monsoon: insight from an ensemble of sub-precessional simulations. Climate of the Past 11, 12711295.Google Scholar
Masao, F.T., Anton, S.C. and Njau, J.K. (2013). Results of recent investigations of the Oldowan and associated hominid remains at the DK site, Olduvai Gorge: a conservation exercise. Proceedings of the 22 International Symposium Africa, Cradle of Humanity: Recent Discoveries, Travaux du Centre National de Recherches Préhistoriques, Anthropologiques et Historiques 18, 147168.Google Scholar
Maslin, M.A. and Christensen, B. (2007). Tectonics, orbital forcing, global climate change, and human evolution in Africa: introduction to the African paleoclimate special volume. Journal of Human Evolution 53(5), 443464.Google Scholar
Maslin, M.A. and Trauth, M.H. (2009). Plio-Pleistocene East African pulsed climate variability and its influence on early human evolution. In: Grine, F.E., Fleagle, J.G. and Leakey, R.E. (Eds.), The First Humans: Origin and Early Evolution of the Genus Homo. Dordrecht: Springer, pp. 151158.Google Scholar
Maslin, M.A., Brierley, C., Milner, A., et al. (2014). East African climate pulses and early human evolution. Quaternary Science Reviews 101, 117.Google Scholar
Maslin, M.A., Shultz, S. and Trauth, M.H. (2015). A synthesis of the theories and concepts of early human evolution. Philosophical Transactions of the Royal Society B 370, 2014.0064.Google Scholar
Mason, R.J. (1962). Prehistory of the Transvaal. Johannesburg: Witwatersrand University Press.Google Scholar
Mason, R.J. (1988a). A Middle Stone Age faunal site at Kalkbank, northerrn Transvaal: archaeology and interpretation. Paleoecology of Africa 19, 201203.Google Scholar
Mason, R.J. (1988b). Cave of Hearths, Makapansgat Transvaal. Occasional Paper No. 21. Johannesburg: Archaeological Research Unit, University of the Witwatersrand.Google Scholar
Masters, J.C. and Rayner, R.J. (1993). Competition and macroevolution: the ghost of competition yet to come? Biological Journal of the Linnean Society 49, 8798.Google Scholar
Matthews, T. (1999). Taphonomy and the micromammals from Elands Bay Cave. The South African Archaeological Bulletin 54(170), pp.133140.Google Scholar
Matthews, T., Parkington, J.E. and Denys, C. (2006a). The taphonomy of the micromammals from the late Middle Pleistocene site of Hoedjiespunt 1 (Cape Province, South Africa). Journal of Taphonomy 4(1), 1126.Google Scholar
Matthews, T., Denys, C. and Parkington, J.E. (2006b). An analysis of the mole rats (Mammalia: Rodentia) from Langebaanweg (Mio-Pliocene, South Africa). Geobios 39(6), 853864.Google Scholar
Matthews, T., Rector, A., Jacobs, Z., Herries, A.I. and Marean, C.W. (2011). Environmental implications of micromammals accumulated close to the MIS 6 to MIS 5 transition at Pinnacle Point Cave 9 (Mossel Bay, Western Cape Province, South Africa). Palaeogeography, Palaeoclimatology, Palaeoecology 302(3–4), 213229.Google Scholar
Matthews, T., van Dijk, E., Roberts, D.L. and Smith, R.M. (2015). An early Pliocene (5.1 Ma) fossil frog community from Langebaanweg, south-western Cape, South Africa. African Journal of Herpetology 64(1), 3953.Google Scholar
Matthews, T., John Measey, G. and Roberts, D.L. (2016). Implications of summer breeding frogs from Langebaanweg, South Africa: regional climate evolution at 5.1 mya. South African Journal of Science 112(9–10), 17.Google Scholar
Matthews, T., Marean, C.W. and Cleghorn, N. (2019). Past and present distributions and community evolution of Muridae and Soricidae from MIS 9 to MIS 1 on the edge of the Palaeo-Agulhas Plain (south coast, South Africa). Quaternary Science Reviews 235, 105774.Google Scholar
Matmon, A., Ron, H., Chazan, M., Porat, N. and Horwitz, L.K. (2012). Reconstructing the history of sediment deposition in caves: a case study from Wonderwerk Cave. Geological Society of America Bulletin 124, 611625.Google Scholar
Matmon, A., Hidy, A.J., Vainer, S., et al. (2015). New chronology for the southern Kalahari Group sediments with implications for sediment-cycle dynamics and early hominin occupation. Quaternary Research 84, 118132.Google Scholar
Mawby, I.E. (1970). Fossil vertebrates from northern Malawi: preliminary report. Quaternaria 13, 319323.Google Scholar
Maxwell, S.J. (2018). The quality of the early hominin fossil record: implications for evolutionary analyses. PhD thesis, Department of Earth and Planetary Sciences, Birkbeck, University of London.Google Scholar
Maxwell, S.J., Hopley, P.J., Upchurch, P. and Soligo, C. (2018). Sporadic sampling, not climatic forcing, drives observed early hominin diversity. Proceedings of the National Academy of Sciences 115(19), 48914896.Google Scholar
Mayr, E. (1942). Systematics and the Origin of Species. New York: Columbia University Press.Google Scholar
Mayr, E. (1963). Animal Species and Evolution. Cambridge, MA: Harvard University Press.Google Scholar
Mayr, E. (1970). Populations, Species, and Evolution. Cambridge, MA: Harvard University Press.Google Scholar
Mbua, E., Kusaka, S., Kunimatsu, Y., et al. (2016). Kantis: a new Australopithecus site on the shoulders of the Rift Valley near Nairobi, Kenya. Journal of Human Evolution 94, 2844.Google Scholar
McBrearty, S. and Brooks, A.S. (2000). The revolution that wasn’t: a new interpretation of the origin of modern human behavior. Journal of Human Evolution 39, 453463.Google Scholar
McBrearty, S. and Jablonski, N.G. (2005). First fossil chimpanzee. Nature 437, 105108.Google Scholar
McBrearty, S., Tryon, C. (2006). From Acheulean to Middle Stone Age in the Kapthurin Formation, Kenya. In: Hovers, E. and Kuhn, S.L. (Eds.), Transitions before the Transition. New York: Springer, pp. 257277.Google Scholar
McBrearty, S., Bishop, L. and Kingston, J.D. (1996). Variability in traces of Middle Pleistocene hominid behavior in the Kapthurin Formation, Baringo, Kenya. Journal of Human Evolution 30, 563580.Google Scholar
McCall, G.J.H., Baker, B.H. and Walsh, J. (1967). Late Tertiary and Quaternary sediments of the Kenya Rift Valley. In: Bishop, W.W. and Clark, J.D. (Eds.), Background to Evolution in Africa. Chicago: Chicago University Press, pp. 191220.Google Scholar
McCarthy, T.C. and Hancox, P.J. (2000). Wetlands. In: Partridge, T.C. and Maud, R.R. (Eds.), Cenozoic of Southern Africa. Oxford: Oxford University Press, pp. 218235.Google Scholar
McDougall, 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. and Brown, F.H. (2006). Precise 40Ar/39Ar geochronology for the upper Koobi Fora Formation, Turkana Basin, northern Kenya. Journal of the Geological Society of London 163, 205220.Google Scholar
McDougall, I. and Brown, F.H. (2008). Geochronology of the pre-KBS Tuff sequence, Omo Group, Turkana Basin. Journal of the Geological Society 165, 549562.Google Scholar
McDougall, I. and Feibel, C.S. (1999). Numerical age control for the Miocene–Pliocene succession at Lothagam, a hominoid-bearing sequence in the northern Kenya Rift. Journal of the Geological Society 156(4), 731745.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
McDougall, I., Brown, F.H. and Fleagle, J.G. (2008). Sapropels and the age of hominins Omo I and II, Kibish, Ethiopia. Journal of Human Evolution 55, 409420.Google Scholar
McDougall, I., Brown, F.H., Vasconcelos, P.M., et al. (2012). New single crystal 40Ar/39Ar ages improve time scale for deposition of the Omo Group, Omo–Turkana Basin, East Africa. Journal of the Geological Society 169, 213226.Google Scholar
McFadden, P.L., Brock, A. and Partridge, T.C. (1979). Paleomagnetism and the age of the Makapansgat hominid site. Earth and Planetary Science Letters 44, 411415.Google Scholar
McGraw, W.S., Cooke, C. and Shultz, S. (2006). Primate remains from African crowned eagle Stephanoaetus coronatus nests in Ivory Coast’s Tai Forest: implications for primate predation and early hominid taphonomy in South Africa. American Journal of Physical Anthropology 131, 151165.Google Scholar
McHenry, H.M. (1986). The first bipeds: a comparison of the Australopithecus afarensis and Australopithecus africanus postcranium and implications for the evolution of bipedalism. Journal of Human Evolution 15, 177191.Google Scholar
McHenry, H.M. (1991). First steps? Analyses of the postcranium of early hominids. In: Senut, B. and Coppens, Y. (Eds.), Origine(s) de la Bipédie chez les Hominidés. Paris: CNRS, pp. 133142.Google Scholar
McHenry, H.M. (1994). Early hominid postcrania: phylogeny and function. In: Corruccini, R.S. and Ciochon, R.L. (Eds.), Integrative Paths to the Past: Paleoanthropological Advances in Honor of F. Clark Howell. Englewood Cliffs: Prentice Hall, pp. 251268.Google Scholar
McHenry, L.J. (2005). Phenocryst composition as a tool for correlating fresh and altered tephra, Bed I, Olduvai Gorge, Tanzania. Stratigraphy 2, 101115.Google Scholar
McHenry, L.J. (2011). Geochemistry and mineralogy of the Laetoli area tuffs: Lower Laetolil through Naibadad Beds. In: Harrison, T. (Ed.), Paleontology and Geology of Laetoli, Tanzania: Human Evolution in Context. Volume 1: Geology, Geochronology, Paleoecology and Paleoenvironment. Dordrecht: Springer, pp. 121141.Google Scholar
McHenry, L.J. (2012). A revised stratigraphic framework for Olduvai Gorge Bed I based on tuff geochemistry. Journal of Human Evolution 63, 284299.Google Scholar
McHenry, L.J. and Stanistreet, I.G. (2018). Tephrochronology of Bed II, Olduvai Gorge, Tanzania, and placement of the Oldowan–Acheulean transition. Journal of Human Evolution 120, 718.Google Scholar
McHenry, L.J., Mollel, G.F. and Swisher, C.C. (2008). Compositional and textural correlations between Olduvai Gorge Bed I tephra and volcanic sources in the Ngorongoro Volcanic Highlands, Tanzania. Quaternary International 178, 306319.Google Scholar
McHenry, L.J., Luque, L., Gómez, J.Á. and Diez-Martín, F. (2011). Promise and pitfalls for characterizing and correlating the zeolitically altered tephra of the Pleistocene Peninj Group, Tanzania. Quaternary Research 75, 708720.Google Scholar
McHenry, L.J., Stollhofen, H. and Stanistreet, I.G. (2013). Use of single-grain geochemistry of cryptic tuffs and volcaniclastic sandstones improves the tephrostratigraphic framework of Olduvai Gorge, Tanzania. Quaternary Research 80, 235249.Google Scholar
McHenry, L.J., Njau, J.K., de la Torre, I. and Pante, M.C. (2016). Geochemical “fingerprints” for Olduvai Gorge Bed II tuffs and implications for the Oldowan–Acheulean transition. Quaternary Research 85, 147158.Google Scholar
McKee, J.K. (1991). Palaeo-ecology of the Sterkfontein hominids: a review and synthesis. Palaeontologia Africana 28, 4151.Google Scholar
McKee, J.K. (1993a). The faunal age of the Taung hominid deposit. Journal of Human Evolution 14, 252265.Google Scholar
McKee, J.K. (1993b). Formation and geomorphology of caves in calcareous tufas and implications for the study of Taung fossil deposits. Transactions of the Royal Society of South Africa 48, 307322.Google Scholar
McKee, J.K. (1994). Catalogue of fossil sites at the Buxton Limeworks, Taung. Palaeontologia Africana 31, 7381.Google Scholar
McKee, J.K. (1996). Faunal turnover patterns in the Pliocene and Pleistocene of southern Africa. South African Journal of Science 92(3), 111113.Google Scholar
McKee, J.K. (2001). Faunal turnover rates and mammalian biodiversity of the late Pliocene and Pleistocene of eastern Africa. Paleobiology 27, 500511.Google Scholar
McKee, J.K. (2010). Taphonomic processes of bone distribution and deposition in the tufa caves of Taung, South Africa. Journal of Taphonomy 8, 203213.Google Scholar
McKee, J.K. (2016). Return to the Taung cave paradigm. American Journal of Physical Anthropology 159(2), 348351.Google Scholar
McKee, J.K. and Keyser, A.W. (1994). Craniodental remains of Papio angusticeps from the Haasgat cave site, South Africa. International Journal of Primatology 15(6), 823841.Google Scholar
McKee, J.K. and Kuykendall, K.L. (2016). The Dart deposits of the Buxton Limeworks, Taung, South Africa, and the context of the Taung Australopithecus fossil. Journal of Vertebrate Paleontology 36(2), e1054937.Google Scholar
McKee, J.K. and Tobias, P.V. (1990). New fieldwork at the Taung Hominid Site: 1988–1989. American Journal of Physical Anthropology 81, 266267.Google Scholar
McKee, J.K. and Tobias, P.V. (1994). Taung stratigraphy and taphonomy: preliminary results based on the 1988–89 excavations. South African Journal of Science 90, 233235.Google Scholar
McKee, J.K., Thackeray, J.F. and Berger, L.R. (1995). Faunal assemblage seriation of Southern African Pliocene and Pleistocene fossil deposits. American Journal of Physical Anthropology 106, 235250.Google Scholar
McKee, J.K., von Mayer, A. and Kuykendall, K.L. (2011). New species of Cercopithecoides from Haasgat, North West Province, South Africa. Journal of Human Evolution 60(1), 8393.Google Scholar
McKenzie, D.P., Davies, D. and Molnar, P. (1970). Plate tectonics of the Red Sea and East Africa. Nature 226, 243248.Google Scholar
McNabb, J. and Beaumont, P. (2011). A Report on the Archaeological Assemblages from Excavations by Peter Beaumont at Canteen Koppie, Northern Cape, South Africa. Southampton Series in Archaeology 4. Oxford: Archaeopress.Google Scholar
McPherron, S.P., Alemseged, Z., Marean, C.W., et al. (2010). Evidence for stone-tool-assisted consumption of animal tissues before 3.39 million years ago at Dikika, Ethiopia. Nature 466, 857860.Google Scholar
Meadows, M.E. and Baxter, A.J. (1999). Late Quaternary palaeoenvironments of the southwestern Cape, South Africa: a regional synthesis. Quaternary International 57, 193206.Google Scholar
Medin, T., Martínez-Navarro, B., Rivals, F., Libsekal, Y. and Rook, L. (2015). The late Early Pleistocene suid remains from the paleoanthropological site of Buia (Eritrea): systematics, biochronology and eco-geographical context. Palaeogeography, Palaeoclimatology, Palaeoecology 431, 2642.Google Scholar
Mehlman, M.J. (1987). Provenience, age and associations of archaic Homo sapiens crania from Lake Eyasi, Tanzania. Journal of Archaeological Science 14, 133162.Google Scholar
Mehlman, M.J. (1989). Later Quaternary archaeological sequences in northern Tanzania. PhD thesis, University of Illinois, Urbana-Champaign.Google Scholar
Meijaard, E., Oliver, W.L.R. and d’Huart, J.-P. (2011). Suidae. In: Wilson, D.E. and Mittermeier, R. (Eds.), Handbook of the Mammals of the World. Vol. 2. Hoofed Mammals. Paris: Lynx Edicions, pp. 248291.Google Scholar
Meldrum, D.J. (2004). Fossilized Hawaiian footprints compared with Laetoli hominid footprints. In: Meldrum, D.J. and Hilton, C.E. (Eds.), From Biped to Strider: The Emergence of Modern Human Walking, Running, and Resource Transport. New York: Kluwer Academic, pp. 6383.Google Scholar
Meldrum, D.J., Lockley, M.G., Lucas, S.G. and Musiba, C. (2011). Ichnotaxonomy of the Laetoli trackways: the earliest hominin footprints. Journal of African Earth Sciences 60, 112.Google Scholar
Melson, W.G. and Potts, R. (2002). Origin of reddened and melted zones in Pleistocene sediments of the Olorgesailie basin, southern Kenya rift. Journal of Archaeological Science 29, 307316.Google Scholar
Menter, C.G., Kuykendall, K.L., Keyser, A.W. and Conroy, G.C. (1999). First record of hominid teeth from the Plio-Pleistocene site of Gondolin, South Africa. Journal of Human Evolution 37, 299307.Google Scholar
Merceron, G. and Ungar, P. (2005). Dental microwear and palaeoecology of bovids from the Early Pliocene of Langebaanweg, Western Cape Province, South Africa. South African Journal of Science 101(7), 365370.Google Scholar
Merrick, H.V., Brown, F.H. and Nash, W.P. (1994). Use and movement of obsidian in the Early and Middle Stone Ages of Kenya and Northern Tanzania. In: Childs, S.T. (Ed.), Society, Culture, and Technology in Africa, vol. 11. Philadelphia: University of Pennsylvania Press, pp. 2944.Google Scholar
Merritt, S.R. (2017). Investigating hominin carnivory in the Okote Member of Koobi Fora, Kenya with an actualistic model of carcass consumption and traces of butchery on the elbow. Journal of Human Evolution 112, 105133.Google Scholar
Merzoug, S. and Sari, L. (2008). Re-examination of the Zone I material from Tamar Hat (Algeria): zooarchaeological and technofunctional analyses. African Archaeological Review 25, 5773.Google Scholar
Metcalfe, C.R. (1971). Anatomy of the Monocotyledons: Volume V, Cyperaceae. Oxford: Clarendon Press.Google Scholar
Meyer, A., Kocher, T.D., Basasibwaki, P. and Wilson, A.C. (1990). Monophyletic origin of Lake Victoria cichlid fishes suggested by mitochondrial DNA sequences. Nature 347(6293), 550553.Google Scholar
Meylan, P. (1990). Fossil turtles from the upper Semliki, Zaire. In: Boaz, N.T. (Ed.), Evolution of Environments and Hominidae in the African Western Rift Valley. Martinsville: Virginia Museum of Natural History, pp. 163170.Google Scholar
Miall, D. (1996). The Geology of Fluvial Deposits. Sedimentary Facies, Basin Analysis, and Petroleum Geology. Berlin: Springer.Google Scholar
Michel, P. (1992). Pour une meilleure connaissance du Quaternaire Continental Marocain: les vertébrés fossiles du Maroc Atlantique, Central et Oriental. L’Anthropologie 96, 643656.Google Scholar
Michel, P. and Wengler, L. (1993). Le site paléontologique et archéologique de Doukkala II (Maroc, Pléistocène moyen et supérieur): Premier jalon en Afrique du Nord d’un comportement humain assimilable à “charognage contrôlé et actif’. Comptes Rendu de l’Académie des Sciences Paris 317, 557562.Google Scholar
Michel, P., Campmas, E., Stoetzel, E., et al. (2009). La macrofaune du Pléistocène supérieur d’El Harhoura 2 (Témara, Maroc): données préliminaires. L’Anthropologie 113, 283312.Google Scholar
Michel, P., Campmas, E., Stoetzel, E., et al. (2010). Upper Palaeolithic (layer 2) and Middle Palaeolithic (layer 3) large faunas from El Harhoura 2 Cave (Témara, Morocco): paleontological, paleoecological and paleoclimatic data. Historical Biology 22, 327340.Google Scholar
Mihlbachler, M.C., Rivals, F., Solounias, N. and Semprebon, G.M. (2011). Dietary change and evolution of horses in North America. Science 331(6021), 11781181.Google Scholar
Miller, D.G.M. (1979). Daily basking patterns of the fresh-water turtle. South African Journal of Zoology 14, 139142.Google Scholar
Miller, E.R., Benefit, B.R., McCrossin, M.L., et al. (2009). Systematics of early and middle Miocene Old World monkeys. Journal of Human Evolution 57, 195211.Google Scholar
Miller, E.R., Gunnell, G.F., Abdel Gawad, M.K., et al. (2014). Anthracotheres from Wadi Moghra, Early Miocene, Egypt. Journal of Paleontology 88, 967981.Google Scholar
Miller, J.H., Behrensmeyer, A.K., Du, A., et al. (2014). Ecological fidelity of functional traits based on species presence–absence in a modern mammalian bone assemblage (Amboseli, Kenya). Paleobiology 40, 560583.Google Scholar
Miller, J.M., Hallager, S., Monfort, S.L., et al. (2011). Phylogeographic analysis of nuclear and mtDNA supports subspecies designations in the ostrich (Struthio camelus). Conservation Genetics 12, 423431.Google Scholar
Miller, K.G., Fairbanks, R.G. and Mountain, G.S. (1987) Tertiary oxygen isotope synthesis, sea level history, and continental margin erosion. Paleoceanography 2(1), 119.Google Scholar
Milne, G. (1935). Some suggested units of classification and mapping, particularly for East African soils. Soil Research 4, 183198.Google Scholar
Milne, G. (1947). A soil reconnaissance journey through parts of Tanganyika Territory December 1935 to February 1936. Journal of Ecology 35, 192265.Google Scholar
Mitchell, P. (2002). The Archaeology of Southern Africa. Cambridge: Cambridge University Press.Google Scholar
Moehlman, P.D., Kebede, F. and Yohannes, H. (2015). Equus africanus. The IUCN Red List of Threatened Species 2015, e.T7949A451070994.Google Scholar
Moggi-Cecchi, J. (2003). The elusive’second species’ in Sterkfontein Member 4: the dental metrical evidence: research articles: human origins research in South Africa. South African Journal of Science 99(5–6), 268270.Google Scholar
Moggi-Cecchi, J., Menter, C., Boccone, S. and Keyser, A. (2010). Early hominin dental remains from the Plio-Pleistocene site of Drimolen, South Africa. Journal of Human Evolution 58, 374405.Google Scholar
Mohr, P. (1999). Le système des rifts Africains. Environnement géologique et géographique. In: Gallay, A. (Ed.), Comment l’Homme? A la découverte des premiers hominidés d’Afrique de l’Est. Paris: Errance, pp. 231288.Google Scholar
Mokokwe, W.D. (2006). Goldsmith’s: Preliminary Study of a newly discovered Pleistocene site near Sterkfontein. Unpublished PhD thesis, University of the Witwatersrand.Google Scholar
Mokokwe, D.W. (2016). Taxonomy, taphonomy and spatial distribution of the cercopithecoid postcranial fossils from Sterkfontein caves. Doctoral dissertation, University of the Witwatersrand.Google Scholar
Moleón, M., Sánchez-Zapata, J.A., Margalida, A., et al. (2014). Humans and scavengers: the evolution of interactions and ecosystem services. BioScience 64, 394403.Google Scholar
Mollel, G.F. (2007). Petrochemistry and geochemistry of Ngorongoro volcanic highland complex (NVHC) and its relationship to Laetoli and Olduvai Gorge, Tanzania. PhD dissertation, Rutgers University, New Brunswick, NJ.Google Scholar
Mollel, G.F. and Swisher, C.C. III (2012). The Ngorongoro Volcanic Highland and its relationships to volcanic deposits at Olduvai Gorge and East African Rift volcanism. Journal of Human Evolution 63, 274283.Google Scholar
Mollel, G., Swisher, C., Feigenson, M. and Carr, M.J. (2008). Geochemical evolution of Ngorongoro Caldera, Northern Tanzania: implications for crust–magma interaction. Earth and Planetary Science Letters 271, 337347.Google Scholar
Mollel, G., Swisher, C., Feigenson, M. and Carr, M.J. (2009). Petrogenesis of basalt-trachyte lavas from Olmoti crater, Tanzania. Journal of African Earth Sciences 54, 127143.Google Scholar
Mollel, G.F., Swisher, C.C. III, McHenry, L.J., Feigenson, M.D. and Carr, M.J. (2011). Petrology, geochemistry and age of Satiman, Lemagurut and Oldeani: sources of the volcanic deposits of the Laetoli area. In: Harrison, T. (Ed.), Paleontology and Geology of Laetoli, Tanzania: Human Evolution in Context. Volume 1: Geology, Geochronology, Paleoecology and Paleoenvironment. Dordrecht: Springer, pp. 99119.Google Scholar
Molnar, P., England, P. and Martinod, J. (1993). Mantle dynamics, uplift of the Tibetan Plateau, and the Indian Monsoon. Reviews of Geophysics 31, 357396.Google Scholar
Molnar, P., Boos, W.R. and Battisti, D.S. (2010). Orographic controls on climate and paleoclimate of Asia: thermal and mechanical roles for the Tibetan Plateau. Annual Review of Earth and Planetary Sciences 38, 77102.Google Scholar
Monchot, H. and Aouraghe, H. (2009). Deciphering the taphonomic history of an Upper Paleolithic faunal assemblage from Zouhrah Cave/El Harhoura 1, Morocco. Quaternaire 20, 239253.Google Scholar
Monson, T.A., Brasil, M.F. and Hlusko, L.J. (2015). Materials collected by the southern branch of the UC Africa Expedition with a report on previously unpublished Plio-Pleistocene fossil localities. Paleobios 32(1).Google Scholar
Moore, T., Loutit, T. and Greenlee, S. (1987). Estimating short‐term changes in Eustatic sea level. Paleoceanography 2(6), 625637.Google Scholar
Moorjani, P., Amorim, C.E.G., Arndt, P.F. and Przeworski, M. (2016). Variation in the molecular clock of primates. Proceedings of the National Academy of Sciences 113(38), 1060710612.Google Scholar
Mora, R., Domínguez-Rodrigo, M., de la Torre, I., Luque, L. and Alcalá, L. (2003). The archaeology of the Peninj “ST Complex” (Lake Natron, Tanzania). In: Martínez-Moreno, J., Mora, R. and de la Torre, I. (Eds.), Oldowan: Rather More than Smashing Stones. First Hominid Technology Workshop. Bellaterra, December 2001. Barcelona: Treballs d’Arqueologia, 9, 77116.Google Scholar
Morales, J. and Pickford, M. (2005). Carnivores from the Middle Miocene Ngorora Formation (13–12 Ma), Kenya. Estudios Geológicos 61, 271284.Google Scholar
Morales, J. and Pickford, M. (2006). A large percrocutid carnivore from the Late Miocene (ca. 10–9 Ma) of Nakali, Kenya. Annales de Paléontologie 92, 359366.Google Scholar
Morales, J., Brewer, P. and Pickford, M. (2010). Carnivores (Creodonta and Carnivora) from the basal middle Miocene of Gebel Zelten, Libya, with a note on a large amphicyonid from the middle Miocene of Ngorora, Kenya. Bulletin of the Tethys Geological Society 5, 4354.Google Scholar
Morales, J., Cantalapiedra, J. L., Valenciano, A., et al. (2015). The fossil record of the Neogene carnivore mammals from Spain. Palaeobiodiversity and Palaeoenvironments 95(3), 373386.Google Scholar
Morgan, L.E., Renne, P.R., Kieffer, G., et al. (2012). A chronological framework for a long and persistent archaeological record: Melka Kunture, Ethiopia. Journal of Human Evolution 62, 104115.Google Scholar
Morgan, M.E., Kingston, J.D. and Marino, B.D. (1994). Carbon isotopic evidence for the emergence of C4 plants in the Neogene from Pakistan and Kenya. Nature 367(6459), 162165.Google Scholar
Morison, C.G.T., Hoyle, A.C. and Hope-Simpson, J.F. (1948). Tropical soil–vegetation catenas and mosaics: a study in the south-western part of the Anglo-Egyptian Sudan. Journal of Ecology 36(1), 184.Google Scholar
Morley, R.J. (2000). Origin and Evolution of Tropical Rain Forests. New York: John Wiley & Sons.Google Scholar
Morley, R.J. and Kingdon, J. (2013). Africa’s environmental and climatic past. In: Kingdon, J., Happold, D., Hoffmann, M., et al. (Eds.) Mammals of Africa. Volume I: Introductory Chapters and Afrotheria. London: Bloomsbury Publishing, pp. 4356.Google Scholar
Morley, R.J. and Richards, K. (1993). Gramineae cuticle: a key indicator of Late Cenozoic climatic change in the Niger Delta. Review of Paleobotany and Palynology 77, 119127.Google Scholar
Morlo, M., Miller, E.R. and El-Barkooky, A.N. (2007). Creodonta and Carnivora from Wadi Moghra, Egypt. Journal of Vertebrate Paleontology 27, 145159.Google Scholar
Morlo, M., Miller, E.R., Bastl, K., et al. (2019). New Amphicyonids (Mammalia, Carnivora) from Moghra, Early Miocene, Egypt. Geodiversitas 41, 731745.Google Scholar
Morris, S.F. (1976). A new fossil freshwater crab from the Ngorora Formation (Miocene) of Kenya. Bulletin of the British Museum (Natural History) Geology 27, 295300.Google Scholar
Morton, W.H., Rex, D.C., Mitchell, J.G. and Mohr, P. (1979). Riftward younging of volcanic units in the Addis Ababa region, Ethiopian rift valley. Nature 280, 284288.Google Scholar
Mourer-Chauviré, C. (2016). Les Oiseaux. In: Raynal, J.-P. and Mohib, A. (Eds.), Préhistoire de Casablanca. 1 ‒ La Grotte des Rhinocéros (fouilles 1991 et 1996). Villes et Sites Archéologiques du Maroc, volume 6. Rabat: Ministère de la Culture, INSAP, pp. 8990.Google Scholar
Mourer-Chauviré, C. and Geraads, D. (2008). The Struthionidae and Pelagornithidae (Aves: Struthioniformes, Odontopterygiformes) from the late Pliocene of Ahl Al Oughlam, Morocco. Oryctos 7, 169194.Google Scholar
Mourer-Chauviré, C. and Geraads, D. (2010). The upper Pliocene avifauna of Ahl al Oughlam, Morocco. Systematics and biogeography. Records of the Australian Museum 62(1), 157184.Google Scholar
Moussa, A., Novello, A., Lebatard, A.-E., et al. (2016). Lake Chad sedimentation and environments during the late Miocene and Pliocene: new evidence from mineralogy and chemistry of the Bol core sediments. Journal of African Earth Sciences 118, 192204.Google Scholar
Moyà-Solà, S., Agustí, J. and Pons, J. (1984). The Mio-Pliocene insular faunas from the west Mediterranean origin and distribution factors. Paléobiologie continentale 14(2), 347357.Google Scholar
Mucina, L. and Rutherford, M.C. (2006). The Vegetation of South Africa, Lesotho and Swaziland. Strelitzia, vol. 19. Pretoria: South African National Biodiversity Institute.Google Scholar
Mucina, L., Rutherford, M.C. and Powrie, L.W. (2006). Vegetation atlas of South Africa, Lesotho and Swaziland. In: Mucina, L. and Rutherford, M.C. (Eds.), The Vegetation of South Africa, Lesotho and Swaziland. Pretoria: South African National Biodiversity Institute, pp. 748789.Google Scholar
Mulholland, S.C. (1989). Phytolith shape frequencies in North Dakota grasses: a comparison to general patterns. Journal of Archaeological Science 16, 489511.Google Scholar
Mullin, S.K., Pillay, N. and Taylor, P.J. (2005). The distribution of the water rat Dasymys (Muridae) in Africa: a review. South African Journal of Science 101(3), 117124.Google Scholar
Musiba, C.M. (1999). Laetoli Pliocene paleoecology: a reanalysis via morphological and behavioral approaches. PhD dissertation, University of Chicago, Chicago.Google Scholar
Musiba, C., Magori, C., Stoller, M., et al. (2007). Taphonomy and paleoecological context of the Upper Laetolil Beds (Localities 8 and 9), Laetoli in northern Tanzania. In: Bobe, R., Alemseged, Z. and Behrensmeyer, A.K. (Eds.), Hominin Environments in the East African Pliocene: An Assessment of the Faunal Evidence. Dordrecht: Springer, pp. 257278.Google Scholar
Musiba, C.M., Mabulla, A., Selvaggio, M. and Magori, C.C. (2008). Pliocene animal trackways at Laetoli: research and conservation potential. Ichnos 15, 166178.Google Scholar
Mussi, M., Altamura, F., Macchiarelli, R., Melis, R.T. and Spinapolice, E. (2014). Garba III (Melka Kunture, Ethiopia): a MSA site with archaic Homo sapiens remains revisited. Quaternary International 343, 2839.Google Scholar
Mussi, M., Altamura, F., Di Bianco, L., et al. (2021). After the emergence of the Acheulean at Melka Kunture (Upper Awash, Ethiopia): from Gombore IB (1.6 Ma) to Gombore Iγ (1.4 Ma), Gombore Iδ (1.3 Ma) and Gombore II OAM Test Pit C (1.2 Ma). Quaternary International. DOI 10.1016/j.quaint.2021.02.031Google Scholar
Mutakyahwa, M. (1997). Mineralogy of the Wembere-Manonga Formation, Manonga Valley, Tanzania, and the possible provenance of the sediments. In: Harrison, T. (Ed.), Neogene Paleontology of the Manonga Valley, Tanzania. New York: Plenum, pp. 6778.Google Scholar
Myers, N., Mittermeier, R.A., Mittermeier, C.G., de Fonesca, G.A.B. and Kent, J. (2000). Biodiversity hotspots for conservation priorities. Nature 403, 853858.Google Scholar
Nagaoka, S., Katoh, S., WoldeGabriel, G., et al. (2005). Lithostratigraphy and sedimentary environments of the hominid-bearing Pliocene–Pleistocene Konso Formation in the southern Main Ethiopian Rift, Ethiopia. Palaeogeography, Palaeoclimatology, Palaeoecology 216, 333457.Google Scholar
Naidoo, J. (1987). The siege of Makapansgat: a massacre? And a Trekker victory? History in Africa 14, 173187.Google Scholar
Nakatsukasa, M., Mbua, E., Sawada, Y., et al. (2010). Earliest colobine skeletons from Nakali, Kenya. American Journal of Physical Anthropology 143, 365382.Google Scholar
Nakaya, H., Pickford, M., Yasui, K. and Nakano, Y. (1987). Additional large mammalian fauna from the Namurungule Formation, Samburu Hills, northern Kenya. African Study Monographs, Supplementary Issue 5, 4798.Google Scholar
Nami, H.G., de la Pefia, P., Vásquez, C.A., Feathers, J. and Wurz, S. (2016). Palaeomagnetic results and new dates of sedimentary deposits from Klasies River Cave 1, South Africa. South African Journal of Science, 112(11–12), 112.Google Scholar
NASA Earth Observatory (2004). Topography of Olduvai Gorge, East Africa. http://earthobservatory.nasa.gov/IOTD/view.php?id=4705.Google Scholar
Nash, D.J. and Endfield, G.H. (2002). A 19th century climate chronology for the Kalahari region of central southern Africa derived from missionary correspondence. International Journal of Climatology 22, 821841.Google Scholar
Ndessokia, P.N.S. (1987). Paleofauna composition and paleoenvironment of Peninj (Marita-Nane site) in West Lake Natron basin, Northern Tanzania. Sciences Géologiques, Bulletins et Memoires 40, 203208.Google Scholar
Ndessokia, P.N.S. (1990). The mammalian fauna and archaeology of the Ndolanya and Olpiro Beds, Laetoli, Tanzania. PhD dissertation, University of California, Berkeley.Google Scholar
Negash, E.W., Alemseged, Z., Wynn, J.G. and Bedaso, Z.K. (2015). Paleodietary reconstruction using stable isotopes and abundance analysis of bovids from the Shungura Formation of South Omo, Ethiopia. Journal of Human Evolution 88, 127136.Google Scholar
Negash, E.W., Alemseged, Z., Bobe, R., et al. (2020). Dietary trends in herbivores from the Shungura Formation, southwestern Ethiopia. Proceedings of the National Academy of Science 117(36), 2192121927.Google Scholar
Nel, T.H. and Henshilwood, C.S. (2016). The small mammal sequence from the c. 76–72 ka still bay levels at Blombos Cave, South Africa – Taphonomic and Palaeoecological implications for human behaviour. PLoS ONE 11(8).Google Scholar
Nel, T.H., Wurz, S. and Henshilwood, C.S. (2018). Small mammals from marine isotope stage 5 at Klasies River, South Africa – reconstructing the local palaeoenvironment. Quaternary International 471, 620.Google Scholar
Nespoulet, R., El Hajraoui, M.A., Amani, F., et al. (2008). Palaeolithic and Neolithic occupations in the Temara region (Rabat, Morocco): recent data on hominin contexts and behavior. African Archaeological Review 25, 2139.Google Scholar
Neukirchen, F., Finkenbein, T. and Keller, J. (2010). The lava sequence of the East African Rift escarpment in the Oldoinyo Lengai–Lake Natron sector, Tanzania. Journal of African Earth Sciences 58, 734751.Google Scholar
Newham, E., Benson, R., Upchurch, P., Goswami, A. (2014). Mesozoic mammaliaform diversity: the effect of sampling corrections on reconstructions of evolutionary dynamics. Palaeogeography, Palaeoclimatology, Palaeoecology 412, 3244.Google Scholar
Nicholson, K.E. (2003). Palaeoenvironment and palaeoecology at Kanjera North: the isotopic evidence. MSc thesis (unpublished), University of Oxford.Google Scholar
Nicholson, S.E. (2000). The nature of rainfall variability over Africa on time scales of decades to millennia. Global and Planetary Change 26(1–3), 137158.Google Scholar
Nicholson, S.E., Kim, J., Ba, M.B. and Lare, A.R. (1997). The mean surface water balance over Africa and its interannual variability. Journal of Climate 10, 29813002.Google Scholar
Nigro, J.D., Ungar, P.S. de Ruiter, D.J. and Berger, L.R. (2003). Developing a geographic information system (GIS) for mapping and analyzing fossil deposits at Swartkrans, Gauteng Province, South Africa. Journal of Archeological Science 30, 317324.Google Scholar
Njau, J.K. (2006). The relevance of crocodile to Oldowan hominin paleoecology at Olduvai Gorge, Tanzania. Doctoral dissertation, Rutgers, the State University of New Jersey, New Brunswick.Google Scholar
Njau, J.K. (2012a). Crocodile Predation and Hominin Evolution: Landscape Paleoanthropology at Olduvai Gorge. Saarbrücken: Lambert Academic Publishing.Google Scholar
Njau, J.K. (2012b). Reading Pliocene bones. Science 336, 4647.Google Scholar
Njau, J.K. and Blumenschine, R.J. (2006). A diagnosis of crocodile feeding traces on larger mammal bone, with fossil examples from the Plio-Pleistocene Olduvai Basin, Tanzania. Journal of Human Evolution 50, 142162.Google Scholar
Njau, J.K. and Blumenschine, R.J. (2012). Crocodylian and mammalian carnivore feeding traces on hominid fossils from FLK 22 and FLK NN 3, Plio-Pleistocene, Olduvai Gorge, Tanzania. Journal of Human Evolution 63, 408417.Google Scholar
Njau, J.K., Stanistreet, I., McHenry, L.J., et al. (2015). New Investigations on Hominin Paleolandscapes, Paleoenvironments and Paleoclimates through Scientific Drilling at Olduvai Gorge, Tanzania. Baltimore, MD: Geological Society of America (GSA).Google Scholar
Norton-Griffiths, M., Herlocker, D. and Pennycuick, L. (1975). The patterns of rainfall in the Serengeti Ecosystem, Tanzania. African Journal of Ecology 13, 347374.Google Scholar
Novello, A., Barboni, D., Sylvestre, F., et al. (2017). Phytoliths indicate significant arboreal cover at Sahelanthropus type locality TM266 in northern Chad and a decrease in later sites. Journal of Human Evolution 106, 6683.Google Scholar
Oakley, K.P., Campbell, B.G. and Molleson, T.I. (1975). Catalogue of Fossil Hominids (Vol. 2). London: British Museum (Natural History).Google Scholar
Occhietti, S., Raynal, J.-P., Pichet, P. and Lefèvre, D. (2002). Aminostratigraphie des formations littorales pléistocènes et holocènes de la région de Casablanca, Maroc. Quaternaire 13, 5563.Google Scholar
Odes, E.J., Parkinson, A.H., Randolph-Quinney, P.S., et al. (2017). Osteopathology and insect traces in the Australopithecus africanus skeleton StW 431. South African Journal of Science 113(1–2), 17.Google Scholar
Ogola, C.A. (2009). The Sterkfontein western breccias: statigraphy, fauna and artefacts. Doctoral dissertation, University of the Witwatersrand.Google Scholar
O’Leary, M.H. (1988). Carbon isotopes in photosynthesis. Bioscience 38(5), 328336.Google Scholar
Oliver, J.S. (1994). Estimates of hominid and carnivore involvement in the FLK-Zinjanthropus fossil assemblage: some socioecological implications. Journal of Human Evolution 27, 267294.Google Scholar
Oliver, J.S., Plummer, T.W., Hertel, F. and Bishop, L.C. (2019). Bovid mortality patterns from Kanjera South, Homa Peninsula, Kenya and FLK-Zinj, Olduvai Gorge, Tanzania: evidence for habitat mediated variability in Oldowan hominin hunting and scavenging behavior. Journal of Human Evolution 131, 6175.Google Scholar
Opperman, H. (1978). Excavations in the Buffelskloof rock shelter near Calitzdorp, Southern Cape. South African Archaeological Bulletin 33, 1838.Google Scholar
O’Regan, H.J. and Reynolds, S.C. (2009). An ecological reassessment of the southern Africa carnivore guild: a case study from Member 4, Sterkfontein, South Africa. Journal of Human Evolution 57, 212222.Google Scholar
O’Regan, H.J., Bishop, L.C., Lamb, A., Elton, S. and Turner, A. (2005). Large mammal turnover in Africa and the Levant between 1.0 and 0.5 Ma. In: Head, M.J. and Gibbard, P.L. (Eds.), Early–Middle Pleistocene Transitions: The Land–Ocean Evidence. Geological Society Special Publications, 247. London: Royal Geological Society, pp. 231349.Google Scholar
O’Regan, H.J., Kuman, K. and Clarke, R.J. (2011a). The likely accumulators of bones: five Cape porcupine den assemblages and the role of porcupines in the Post-Member 6 Infill at Sterkfontein, South Africa. Journal of Taphonomy 9(2), 6987.Google Scholar
O’Regan, H.J., Turner, A., Bishop, L.C., Elton, S. and Lamb, A.L. (2011b). Hominins without fellow travellers? First appearances and inferred dispersals of Afro-Eurasian large-mammals in the Plio-Pleistocene. Quaternary Science Review 30, 13431352.Google Scholar
O’Regan, H.J., Cohen, B.F. and Steininger, C.M. (2013). Mustelid and viverrid remains from the Pleistocene site of Cooper’s D, Gauteng, South Africa. Palaeontologica Africana 48, 1923.Google Scholar
Otero, O., Pinton, A., Mackaye, H.T., et al. (2009). Fishes and palaeogeography of the African drainage basins: relationships between Chad and neighbouring basins throughout the Mio-Pliocene. Palaeogeography, Palaeoclimatology, Palaeoecology 274, 134139.Google Scholar
Otero, O., Pinton, A., Mackaye, H.T., et al. (2010). The early/late Pliocene ichthyofauna from Koro-Toro, Eastern Djurab, Chad. Geobios 43, 241251.Google Scholar
Otero, O., Lécuyer, C., Fourel, F., et al. (2011) Freshwater fish δ18O indicates a Messinian change of the precipitation regime in Central Africa. Geology 39(5), 435438.Google Scholar
Owen, D.F. (1965). A populational study of an Equatorial land snail, Limicolaria martensiana (Achatinidae). Proceedings of the Zoological Society of London 144, 361382.Google Scholar
Owen, R.B. and Renaut, R.W. (1986). Sedimentology, stratigraphy and palaeoenvironments of the Holocene Galana Boi Formation, NE Lake Turkana, Kenya. Geological Society, London, Special Publications 25, 311.Google Scholar
Owen, R.B., Potts, R., Behrensmeyer, A.K., Ditchfield, P. (2008). Diatomaceous sediments and environmental change in the Pleistocene Olorgesailie Formation, southern Kenya Rift Valley. Palaeogeography, Palaeoclimatology, Palaeoecology 269, 1737.Google Scholar
Owen, R.B., Renaut, R.W., Scott, J.J., Potts, R. and Behrensmeyer, A.K. (2009). Wetland sedimentation and associated diatoms in the Pleistocene Olorgesailie Basin, southern Kenya Rift Valley. Sedimentary Geology 222, 124137.Google Scholar
Owen-Smith, R.N. (1988). Megaherbivores: The Influence of Very Large Body Size on Ecology. Cambridge: Cambridge University Press.Google Scholar
Pallas, L., Daver, G., Mackaye, H.T., et al. (2019). A window into the early evolutionary history of Cercopithecidae: Late Miocene evidence from Chad, Central Africa. Journal of Human Evolution 132, 6179.Google Scholar
Palmqvist, P., Pérez-Claros, J.A., Janis, C.M. and Gröcke, D.R. (2008). Tracing the ecophysilogy of ungulates and predator–prey relationships in an early Pleistocene large mammal community. Palaeogeography, Palaeoclimatology, Palaeoecology 266, 95111.Google Scholar
Paquet, H. (1970). Evolution géochimique des minéraux argileux dans les altérations et les sols des climats méditérranéens et tropicaux à saisons contrastées. Bulletin et Memoires du Service de la Carte Geologique d’Alsace et de Lorraine 30, 1210.Google Scholar
Pante, M.C. and Blumenschine, R.J. (2010). Fluvial transport of bovid long bones fragmented by the feeding activities of hominins and carnivores. Journal of Archaeological Science 37, 846854.Google Scholar
Pante, M.C., Blumenschine, R.J., Capaldo, S.D. and Scott, R.S. (2012). Validation of bone surface modification models for inferring fossil hominin and carnivore feeding interactions, with reapplication to FLK 22, Olduvai Gorge, Tanzania. Journal of Human Evolution 63, 395407.Google Scholar
Pante, M.C., Blumenschine, R.J., Capaldo, S.D. and Scott, R.S. (2015). Revalidation of bone surface modification models for inferring fossil hominin and carnivore feeding interactions. Quaternary International 355, 164168.Google Scholar
Parker, J., Hopley, P.J. and Kuhn, B.F. (2016). Fossil carder bee’s nest from the hominin locality of Taung, South Africa. PLoS ONE 11(9), e0161198.Google Scholar
Parkington, J.E. (1972). Seasonal mobility in the late Stone Age. African Studies 31, 223–44.Google Scholar
Parkington, J.E. (1976). Coastal settlement between the mouths of the Berg and the Olifants rivers, Cape Province. South African Archaeological Bulletin 31, 127140.Google Scholar
Parkington, J.E. (1980). The Elands Bay cave sequence: cultural stratigraphy and subsistence strategies. In Proceedings of the 8th Pan-African Congress of Prehistory and Quaternary Studies. Nairobi: International Louis Leakey Memorial Institute for African Prehistory, pp. 315320.Google Scholar
Parkington, J.E. (1981). The effects of environmental change on the scheduling of visits to the Elands Bay Cave, Cape Province, South Africa. In: Hodder, I., Isaac, G. and Hammond, N. (Eds.), Pattern of the Past: Studies in Honour of David Clarke. Cambridge: Cambridge University Press, pp. 341359.Google Scholar
Parkington, J.E. (1987). Changing views of prehistoric settlement in the western Cape. In: Papers in the Prehistory of the Western Cape, South Africa, Vol. 322. Oxford: British Archaeological Reports, pp. 423.Google Scholar
Parkington, J.E. and Poggenpoel, C. (1987). Diepkloof rock shelter. In: Parkington, J.E. and Hall, M., (Eds.), Papers in the Prehistory of the Western Cape, South Africa, Vol. 332. Oxford: British Archaeological Reports, pp. 269293.Google Scholar
Parkington, J., Cartwright, C., Cowling, R.M. and Baxter, A. (2000). Africa: wood charcoal and pollen TS evidence from Elands Bay Cave. South African Journal of Science 96, 543.Google Scholar
Parkinson, A.H. (2016). Traces of insect activity at Cooper’s D fossil site (Cradle of Humankind, South Africa). Ichnos 23(3–4), 322339.Google Scholar
Parkinson, J. (2013). A GIS Image analysis approach to documenting Oldowan hominin carcass acquisition: evidence from Kanjera South, FLK Zinj, and neotaphonomic models of carnivore bone destruction. PhD dissertation, City University of New York.Google Scholar
Parravicini, A. and Pievani, T. (2016). Multi-level human evolution: ecological patterns in hominin phylogeny. Journal of Anthropological Sciences 94, 116.Google Scholar
Partridge, T.C. (1973). Geomorphological dating of cave openings at Makapansgat, Sterkfontein, Swartkrans and Taung. Nature 246, 7579.Google Scholar
Partridge, T.C. (1975). Stratigraphic, geomorphological and palaeoenvironmental studies of the Makapansgat Limeworks and Sterkfontein hominid sites: a progress report on research carried out between 1965 and 1975. In: Conference Report on Recent Progress in Later Cenozoic Studies in Southern Africa, Cape Town.Google Scholar
Partridge, T.C. (1978). Re-appraisal of lithostratigraphy of Sterkfontein hominid site. Nature 275(5678), 282287.Google Scholar
Partridge, T.C. (1979). Re-appraisal of lithostratigraphy of Makapansgat Limeworks hominid site. Nature 279(5713), 484488.Google Scholar
Partridge, T.C. (1982). Some preliminary observations on the stratigraphy and sedimentology of the Kromdraai B hominid site. In: Coetzee, J.A. and Van Zinderen Bakker, E.M. (Eds.), Palaeoecology of Africa and the Surrounding Islands, Vol. 15. Rotterdam: Balkema, pp. 312.Google Scholar
Partridge, T.C. (1985). Spring flow and tufa accretion at Taung. In: Tobias, P.V. (Ed.), Hominid Evolution Past, Present and Future. New York: Alan R. Liss, pp. 171187.Google Scholar
Partridge, T.C. (2000). Hominid-bearing cave and tufa deposits. In: Partridge, T.C. and Maud, R.R. (Eds.), The Cenozoic of Southern Africa. Oxford: Oxford University Press, pp. 100130.Google Scholar
Partridge, T.C. (2005). Dating of the Sterkfontein hominids: progress and possibilities. Transactions of the Royal Society of South Africa 60(2), 107110.Google Scholar
Partridge, T.C. (2010). Tectonics and geomorphology of Africa during the Phanerozoic. In: Werdelin, L. and Sanders, W.J. (Eds.), Cenozoic Mammals of Africa. Berkeley: University of California Press, pp. 317.Google Scholar
Partridge, T.C. and Maud, R.R. (1987). Geomorphic evolution of southern Africa since the Mesozoic. South African Journal of Geology 90, 179208.Google Scholar
Partridge, T.C., Bollen, J., Tobias, P.V. and McKee, J.K. (1991). New light on the provenance of the Taung Skull. South African Journal of Science 87, 8083.Google Scholar
Partridge, T.C., deMenocal, P.B., Lorentz, S., Paiker, M.J. and Vogel, J.C. (1997). Orbital forcing of climate over South Africa: a 200,000-year rainfall record from the Pretoria Saltpan. Quaternary Science Reviews 16, 11251133.Google Scholar
Partridge, T.C., Shaw, J., Heslop, D. and Clarke, R.J. (1999). The new hominid skeleton from Sterkfontein, South Africa: age and preliminary assessment. Journal of Quaternary Science 14(4), 293298.Google Scholar
Partridge, T.C., Granger, D.E., Caffee, M.W. and Clarke, R.J. (2003). Lower Pliocene hominid remains from Sterkfontein. Science 300, 607612.Google Scholar
Passey, B.H., Cerling, T.E., Perkins, M.E., et al. (2002). Environmental change in the Great Plains: an isotopic record from fossil horses. Journal of Geology 110, 123140.Google Scholar
Passey, B.H., Robinson, T.F., Ayliffe, L.K., et al. (2005). Carbon isotope fractionation between diet, breath CO2, and bioapatite in different mammals. Journal of Archaeological Science 32, 14591470.Google Scholar
Passey, B.H., Levin, N.E., Cerling, T.E., Brown, F.H. and Eiler, J.M. (2010). High-temperature environments of human evolution in East Africa based on bond ordering in paleosol carbonates. Proceedings of the National Academy of Sciences 107, 1124511249. doi/10.1073/pnas.1001824107.Google Scholar
Patterson, B. (1966). A new locality for early Pleistocene fossils in north-western Kenya. Nature 212, 577581.Google Scholar
Patterson, B. and Howells, W.W. (1967). Hominid humeral fragment from early Pleistocene of northwestern Kenya. Science 156, 6466.Google Scholar
Patterson, B., Behrensmeyer, A.K. and Sill, W.D. (1970). Geology of a new Pliocene locality in northwestern Kenya. Nature 256, 279284.Google Scholar
Patterson, D.B., Faith, J.T., Bobe, R. and Wood, B. (2014). Regional diversity patterns in African bovids, hyaenids, and felids during the past 3 million years: the role of taphonomic bias and implication for the evolution of Paranthropus. Quaternary Science Reviews 96, 922.Google Scholar
Patterson, D.B., Lehmann, S.B., Matthews, T., et al. (2016). Stable isotope ecology of Cape dune mole-rats (Bathyergus suillus) from Elandsfontein, South Africa: Implications for C 4 vegetation and hominin paleobiology in the Cape Floral Region. Palaeogeography, Palaeoclimatology, Palaeoecology 457, 409421.Google Scholar
Patterson, D.B., Braun, D.R., Behrensmeyer, A.K., et al. (2017a). Landscape scale heterogeneity in the East Turkana ecosystem during the Okote Member (1.56–1.38 Ma). Journal of Human Evolution 112, 148161.Google Scholar
Patterson, D.B., Braun, D.R., Behrensmeyer, A.K., et al. (2017b). Ecosystem evolution and hominin paleobiology at East Turkana, northern Kenya between 2.0 and 1.4 Ma. Palaeogeography, Palaeoclimatology, Palaeoecology 481, 113.Google Scholar
Patterson, D.B., Braun, D.R., Allen, K., et al. (2019). Comparative isotopic evidence from East Turkana supports a dietary shift within the genus Homo. Nature Ecology & Evolution 3, 10481056.Google Scholar
Patterson, N., Richter, D.J., Gnerre, S., Lander, E.S. and Reich, D. (2006). Genetic evidence for complex speciation of humans and chimpanzees. Nature 441(7097), 11031108.Google Scholar
Pavia, M. (2019). Geronticus thackerayi, sp. nov. (Aves, Threskiornithidae), a new ibis from the hominin-bearing locality of Kromdraai (Cradle of Humankind, Gauteng, South Africa). Journal of Vertebrate Paleontology, 39(3), e1647433.Google Scholar
Pavia, M., Davies, G.B., Gommery, D. and Kgasi, L. (2017). Mid-Pliocene bald ibis (Geronticus cf. calvus; Aves: Threskiornithidae) from the Cradle of Humankind, Gauteng, South Africa and its environmental and evolutionary implications. PalZ, 91(2), 237243.Google Scholar
Pavlakis, P.P. (1990). Plio-Pleistocene Hippopotamidae from the upper Semliki. In: Boaz, N.T. (Ed.), Evolution of Environments and Hominidae in the African Western Rift Valley. Martinsville: Virginia Museum of Natural History.Google Scholar
Payne, S. (1972). Partial recovery and sample bias: the results of some sieving experiments. In: Higgs, T.S. (Ed.), Papers in Economic Prehistory. Cambridge: Cambridge University Press, pp. 4964.Google Scholar
Payne, S. (1975). Partial recovery and sample bias. In: Clason, A.T. (Ed.), Archaeozoological Studies. Amsterdam: North-Holland Publishing Company, pp. 717.Google Scholar
Peabody, F.E. (1954). Travertines and cave deposits of the Kaap Escarpment of South Africa, and the type locality of Australopithecus africanus Dart. Bulletin of the Geological Society of America 65, 671706.Google Scholar
Peet, R.K. (1974). The measurement of species diversity. Annual Review of Ecology and Systematics 5, 285307.Google Scholar
Peigné, S., de Bonis, L., Likius, A., et al. (2005a). The earliest modern mongoose (Carnivora, Herpestidae) from Africa (late Miocene of Chad). Naturwissenschaften 92, 287292.Google Scholar
Peigné, S., de Bonis, L., Likius, A., et al. (2005b). A new machairodontine (Carnivora, Felidae) from the Late Miocene hominid locality of TM 266, Toros-Menalla, Chad. Comptes Rendus Palevol 4, 243253.Google Scholar
Peigné, S., de Bonis, L., Mackaye, H.T., et al. (2008). Late Miocene Carnivora from Chad: Herpestidae, Viverridae and small-sized Felidae. Comptes Rendus Palevol 7, 499527.Google Scholar
Pennycuick, C.J. and Western, D. (1972). An investigation of some sources of bias in aerial transect sampling of large mammal populations. African Journal of Ecology 10, 175191.Google Scholar
Péringuey, L. (1911). The Stone Ages of South Africa as represented in the collections of the South African Museum. Annals of the South African Museum 8, 1218.Google Scholar
Perini, S., Muttoni, G., Monesi, E., Melis, R.T., Mussi, M. (2021). Magnetochronology and age models of deposition of the Melka Kunture stratigraphic sequence (Upper Awash, Ethiopia) and age assessments of the main archeological levels therein contained. Quaternary Science Reviews 274, 107259.Google Scholar
Peters, C.R. (1999). African wild plants with rootstocks reported to be eaten raw: the monocotyledons, Part IV. In: Timberlake, J. and Kativu, S. (Eds.), African Plants: Biodiversity, Taxonomy and Uses. Kew: Royal Botanic Gardens, pp. 483503.Google Scholar
Peters, C.R. (2007). Theoretical and actualistic ecobotanical perspectives on early hominin diets and paleoecology. In: Ungar, P.S. (Ed.), Evolution of the Human Diet. New York: Oxford University Press, pp. 233261.Google Scholar
Peters, C.R. and Blumenschine, R.J. (1995). Landscape perspectives on possible land use patterns for Early Pleistocene hominids in the Olduvai Basin, Tanzania. Journal of Human Evolution 29, 321362.Google Scholar
Peters, C.R. and Blumenschine, R.J. (1996). Landscape perspectives on possible land use patterns for early Pleistocene hominids in the Olduvai basin, Tanzania: Part II, expanding the landscape models. Kaupia 6, 175221.Google Scholar
Peters, C.R. and O’Brien, E.M. (2001). Palaeo-Lake Congo: implications for Africa’s late Cenozoic climate – some unanswered questions. In: Heine, K. and Runge, J. (Eds.), Palaeoecology of Africa and the Surrounding Islands. Rotterdam: Balkema, pp. 1118.Google Scholar
Peters, C.R. and Vogel, J.C. (2005). Africa’s wild C4 plant foods and possible early hominid diets. Journal of Human Evolution 48, 219236.Google Scholar
Peters, C.R., Blumenschine, R.J., Hay, R.L., et al. (2008). Paleoecology of the Serengeti–Mara ecosystem. In: Sinclair, A.R.E., Packer, C., Mduma, S.A.R. and Fryxell, J.M. (Eds.), Serengeti III: Human Impacts on Ecosystem Dynamics. Chicago: University of Chicago Press, pp. 4794.Google Scholar
Peters, J. (1990a). Late Palaeolithic ungulate fauna and landscape in the plain of Kom Ombo. Sahara 3, 4552.Google Scholar
Peters, J. (1990b). Late Pleistocene hunter-gatherers at Ishango (Eastern Zaire): the faunal evidence. Revue de Paléobiologie 9, 73112.Google Scholar
Peters, R.H. and Raelson, J.V. (1984). Relations between individual size and mammalian population density. The American Naturalist 124, 498571.Google Scholar
Petrocchi, C. (1934). I ritrovamenti faunistici di es-Sahabi. Rivista delle Colonie Italiane 7, 733742.Google Scholar
Petter, G. (1963). Étude quelques Viverridés (Mammifères, Carnivores) du Pléistocène inférieur du Tanganyika (Afrique orientale). Bulletins du Société Géologiques de France, Series 7 5, 265274.Google Scholar
Petter, G. (1973). Carnivores pléistocènes du ravin d’Olduvai (Tanzanie). In: Leakey, L.S.B., Savage, R.J.G. and Coryndon, S.C. (Eds.), Fossil Vertebrates of Africa, vol. 3. London: Academic Press, pp. 44100.Google Scholar
Petter, G. and Howell, F.C. (1988). Nouveau félidé machairodonte (Mammalia, Carnivora) de la faune pliocène de l’Afar (Éthiopie): Homotherium hadarensis n. sp. Comptes Rendus de l’Académie de Sciences, Paris, Série II 306, 731738.Google Scholar
Petter, G., Pickford, M. and Senut, B. (1994). Presence of the genus Agriotherium (Mammalia, Carnivora, Ursidae) in the Late Miocene of the Nkondo Formation (Uganda, East Africa). Comptes Rendus de l’Académie de Sciences 319, 713717.Google Scholar
Peypouquet, J.P., Carbonel, P., Taieb, M., Tiercelin, J.J. and Perinet, G. (1983). Ostracoda and evolution process of paleohydrologic environments in the Hadar Formation (the Afar depression, Ethiopia). In: Maddocks, R.F. (Ed.), Applications of Ostracoda. Houston: University of Houston Geosciences, pp. 277285.Google Scholar
Philander, S.G. and Fedorov, A.V. (2003). Role of tropics in changing the response to Milankovich forcing some three million years ago. Paleoceanography 18, 112.Google Scholar
Pickering, R. (1969). Regional mapping. 1. The geology of the country around Endulen. Records of the Geological Survey of Tanganyika 11, 19.Google Scholar
Pickering, R. (2015). U–Pb dating small buried stalagmites from Wonderwerk Cave, South Africa: a new chronometer for earlier Stone Age cave deposits. The African Archaeological Review 32(4), 645668.Google Scholar
Pickering, R. and Kramers, J.D. (2010). Re-appraisal of the stratigraphy and determination of new U–Pb dates for the Sterkfontein hominin site, South Africa. Journal of Human Evolution 59(1), 7086.Google Scholar
Pickering, R., Hancox, P.J., Lee-Thorp, J.A., et al. (2007). Stratigraphy, U–Th chronology, and paleoenvironments at Gladysvale Cave: insights into the climatic control of South African hominin-bearing cave deposits. Journal of Human Evolution 53(5), 602619.Google Scholar
Pickering, R., Kramers, J.D., Partridge, T., Kodolanyi, J. and Pettke, T. (2010). U–Pb dating of calcite–aragonite layers in speleothems from hominin sites in South Africa by MC-ICP-MS. Quaternary Geochronology 5, 544558.Google Scholar
Pickering, R., Kramers, J.D., Hancox, J.P., de Ruiter, D.J. and Woodhead, J.D. (2011). Contemporary flowstone development links early hominin bearing cave deposits in South Africa. Earth and Planetary Science Letters 306, 2332.Google Scholar
Pickering, R., Herries, A.I., Woodhead, J.D., et al. (2019). U–Pb-dated flowstones restrict South African early hominin record to dry climate phases. Nature 565(7738), 226.Google Scholar
Pickering, T.R. (1999). Taphonomic interpretations of the Sterkfontein early hominid site (Gauteng, South Africa) reconsidered in light of recent evidence. PhD thesis, Department of Anthropology, University of Wisconsin.Google Scholar
Pickering, T.R. (2002). Reconsideration of criteria for differentiating faunal assemblages accumulated by hyaenas and hominids. International Journal of Osteoarchaeology 12, 127141.Google Scholar
Pickering, T.R. and Egeland, C.P. (2006). Experimental patterns of hammerstone percussion damage on bones: implications for inferences of carcass processing by humans. Journal of Archaeological Science 33, 459469.Google Scholar
Pickering, T.R., White, T.D. and Toth, N. (2000). Cutmarks on a Plio-Pleistocene hominid from Sterkfontein, South Africa. American Journal of Physical Anthropology 111, 579584.Google Scholar
Pickering, T.R., Clarke, R.J. and Heaton, J.L. (2004a). The context of Stw 573, an early hominid skull and skeleton from Sterkfontein Member 2: taphonomy and palaeoenvironment. Journal of Human Evolution 46, 277295.Google Scholar
Pickering, T.R., Clarke, R.J. and Moggi-Cecchi, J. (2004b). The role of carnivores in the accumulation of the Sterkfontein Member 4 hominid fossil assemblage: a taphonomic reassessment of the complete hominid fossil sample (1936–1999). American Journal of Physical Anthropology 125, 115.Google Scholar
Pickering, T.R., Dominguez-Rodrigo, M., Egeland, C.P. and Brain, C.K. (2005). The contribution of limb bone fracture patterns to reconstructing early hominid behaviour at Swartkrans Cave (South Africa): archaeological application of a new analytical method. International Journal of Osteoarchaeology 14, 247260.Google Scholar
Pickering, T.R., Egeland, C.P., Dominguez-Rodrigo, M., Brain, C.K. and Schnell, A.G. (2007). Testing the “shift in the balance of power” hypothesis at Swartkrans, South Africa: hominid cave use and subsistence behavior in the early Pleistocene. Journal of Anthropological Archaeology 27, 3045.Google Scholar
Pickering, T.R., Heaton, J.L., Clarke, R.J., et al. (2012). New hominid fossils from Member 1 of the Swartkrans Formation, South Africa. Journal of Human Evolution 62, 618628.Google Scholar
Pickering, T.R., Heaton, J.L., Sutton, M.B., et al. (2016). New early Pleistocene hominin teeth from the Swartkrans formation, South Africa. Journal of Human Evolution 100, 115.Google Scholar
Pickering, T.R., Heaton, J.L., Throckmorton, Z.J., Prang, T.C. and Brain, C.K. (2017). A burned primate cuboid from Swartkrans Cave, South Africa. Annals of the Ditsong National Museum of Natural History 7(1), 17.Google Scholar
Pickett, S.T.A., Kolasa, J. and Jones, C.G. (1994). Ecological Understanding. San Diego: Academic Press.Google Scholar
Pickford, M.H.L. (1978). Geology, palaeoenvironments and vertebrate faunas of the mid-Miocene Ngorora Formation, Kenya. In: Bishop, W.W. (Ed.), Geological Background to Fossil Man. Special Publications. Edinburgh: Geological Society of London and Scottish Academic Press, pp. 237262.Google Scholar
Pickford, M. (1984). Kenya Palaeontology Gazetteer, Volume 1: Western Kenya. Nairobi: National Museums of Kenya.Google Scholar
Pickford, M. (1987). The geology and palaeontology of the Kanam erosion gullies (Kenya). Mainzer Geowissenschaftliche Mitteilungen 6, 209226.Google Scholar
Pickford, M. (1990). Révision des Suidés de la Formation Beglia (Tunisie). Annales de Paléontologie 76, 133142.Google Scholar
Pickford, M. (1993). Climate change, biogeography and Theropithecus. In: Jablonski, N.G. (Ed.), Theropithecus: The Rise and Fall of a Primate Genus. Cambridge: Cambridge University Press, pp 227243.Google Scholar
Pickford, M. (1994a). Patterns of sedimentation and fossil distribution in the Kenya Rift Valleys. Journal of African Earth Sciences 18, 5160.Google Scholar
Pickford, M. (1994b). Fossil Suidae of the Albertine Rift Valley, Uganda–Zaire. In: Senut, B. and Pickford, M. (Eds.), Geology and Palaeobiology of the Albertine Rift Valley, Uganda–Zaire. Palaeobiology-Paléobiologie vol. 2. Occasional Publication. Orléans: CIFEG, pp. 339373.Google Scholar
Pickford, M. (1994c). A new species of Prohyrax (Mammalia, Hyracoidea) from the middle Miocene of Arrisdrift, Namibia. Communications of the Geological Survey of Namibia 9, 4362.Google Scholar
Pickford, M. (1995). Fossil land snails of East Africa and their palaeoecological significance. Journal of African Earth Sciences 20, 167226.Google Scholar
Pickford, M. (2000). Crocodiles from the Beglia Formation, Middle/Late Miocene boundary, Tunisia, and their significance for Saharan palaeoclimatology. Annales de Paléontologie 86(1), 5967.Google Scholar
Pickford, M. (2001a). Equidae in the Ngorora Formation, Kenya, and the first appearance of the family in East Africa. Revista Española de Paleontología 16, 339345.Google Scholar
Pickford, M. (2001b). New species of Listriodon (Suidae, Mammalia) from Bartule, Member A, Ngorora Formation (ca 13 Ma), Tugen Hills, Kenya. Annales de Paléontologie 87, 207221.Google Scholar
Pickford, M. (2004). Paleoenvironments of early Miocene hominoid-bearing deposits at Napak, Uganda, based on terrestrial molluscs. Annals of Paleontology 90, 112.Google Scholar
Pickford, M. (2006). New suoid specimens from Gebel Zelten, Libya. Estudios Geológicos 62, 499514.Google Scholar
Pickford, M. (2008). Libycosaurus petrocchii Bonarelli, 1947, and Libycosaurus anisae (Black, 1972) (Anthracotheriidae, Mammalia): nomenclatural and geochronological implications. Annales de Paléontologie 94, 3955.Google Scholar
Pickford, M. (2009). Land snails from the early Miocene Legetet Formation, Koru, Kenya. Geo-Pal Kenya 2, 188.Google Scholar
Pickford, M. (2013a). The diversity, age, biogeographic and phylogenetic relationships of Plio-Pleistocene suids from Kromdraai, South Africa. Annals of the Ditsong National Museum of Natural History 3, 1132.Google Scholar
Pickford, M. (2013b). Locomotion, diet, body weight, origin and geochronology of Metridiochoerus andrewsi from the Gondolin karst deposits, Gauteng, South Africa. Annals of the Ditsong National Museum of Natural History 3, 3347.Google Scholar
Pickford, M. (2013c). Gorongosa Palaeontology Survey. Gorongosa: Gorongosa National Park, p. 13.Google Scholar
Pickford, M. (2020). The Fossil Suidae (Mammalia, Artiodactyla) from Ternifine (Tighenif) Algeria. Münchner Geowissenschafltiche Abhandlungen A 50, 166.Google Scholar
Pickford, M. and Fischer, M.S. (1987). Parapliohyrax ngororaensis, a new hyracoid from the Miocene of Kenya, with an outline of the classification of Neogene Hyracoidea. Neues Jahrbuch für Geologie und Paläontologie, Abhandlungen 175, 207234.Google Scholar
Pickford, M. and Gommery, D. (2016). Fossil Suidae (Artiodactyla, Mammalia) from Aves Cave I and nearby sites in Bolt’s Farm Palaeokarst System, South Africa. Estudios Geologicos-Madrid 72(2), e059.Google Scholar
Pickford, M. and Gommery, D. (2020). Fossil suids from Bolt’s Farm Palaeokarst System, South Africa: implications for the taxonomy of Potamochoeroides and Notochoerus and for biochronology. Estudios Geologicos-Madrid 76(1), 127.Google Scholar
Pickford, M. and Morales, J. (1994). Biostratigraphy and palaeobiogeography of East Africa and the Iberian Peninsula. Palaeogeography, Palaeoclimatology, Palaeoecology 112(3–4), 297322.Google Scholar
Pickford, M. and Senut, B. (1994). Palaeobiology of the Albertine Rift Valley: general conclusions and synthesis. In: Senut, B. and Pickford, M. (Eds.), Geology and Palaeobiology of the Albertine Rift Valley, Uganda-Zaire, Vol. II, Palaeobiology. Publication Occasionelle.Orléans: Centre International pour la Formation et les Echanges Géologiques – CIFEG, pp. 409423.Google Scholar
Pickford, M. and Senut, B. (1997). Cainozoic mammals from coastal Namaqualand, South Africa. Palaeontologia Africana 34, 199217.Google Scholar
Pickford, M. and Senut, B. (2001a). ‘Millennium Ancestor’, a 6-million-year-old bipedal hominid from Kenya – recent discoveries push back human origins by 1.5 million years. South African Journal of Science, 97(1–2), 2222.Google Scholar
Pickford, M. and Senut, B. (2001b). The geological and faunal context of Late Miocene hominid remains from Lukeino, Kenya. Comptes Rendus de l’Académie des Sciences-Series IIA – Earth and Planetary Science 332(2), 145152.Google Scholar
Pickford, M. and Senut, B. (2005). Hominoid teeth with chimpanzee-and gorilla-like features from the Miocene of Kenya: implications for the chronology of ape-human divergence and biogeography of Miocene hominoids. Anthropological Science 113(1), 95102.Google Scholar
Pickford, M., Johanson, D.C., Lovejoy, C.O., White, T.D. and Aronson, J.L. (1983). A hominoid humeral fragment from the Pliocene of Kenya. American Journal of Physical Anthropology 60, 337346.Google Scholar
Pickford, M., Ishida, H., Nakano, Y. and Yasui, K. (1987). Fossil terrestrial gastropods from the Namurungule Formation, Kenya. African Study Monographs, Supplementary Issue 5, 155156.Google Scholar
Pickford, M., Mein, P. and Senut, B. (1992). Primate bearing Plio-Pleistocene cave deposits of Humpata, southern Angola. Human Evolution 7(1), 1733.Google Scholar
Pickford, M., Senut, B. and Hadoto, D. (1993). Geology and palaeobiology of the Albertine Rift Valley, Uganda-Zaire, Vol. I, Geology. Publication Occasionelle. Orléans: Centre International pour la Formation et les Echanges Géologiques – CIFEG, p. 190.Google Scholar
Pickford, M., Mein, P. and Senut, B. (1994). Fossiliferous Neogene karst fillings in Angola, Botswana and Namibia. South African Journal of Science 90, 227227.Google Scholar
Pickford, M., Senut, B. and Mourer-Chauvire, C. (2004). Early Pliocene Tragulidae and peafowls in the Rift Valley, Kenya: evidence for rainforest in East Africa. Comptes Rendus Palevol 3, 179189.Google Scholar
Pickford, M., Wanas, H. and Soliman, H. (2006). Indications for a humid climate in the Western Desert of Egypt 11–10 Myr ago: evidence from Galagidae (Primates, Mammalia). Comptes Rendus Palevol 5(8), 935943.Google Scholar
Pickford, M., Coppens, Y., Senut, B., Morales, J. and Braga, J. (2009). Late Miocene hominoid from Niger. Comptes Rendus Palevol 8, 413425.Google Scholar
Pickford, M., Miller, E.R. and El-Barkooky, A.N. (2010). Suidae and Sanitheriidae from Wadi Moghra, early Miocene, Egypt. Acta Palaeontologica Polonica 55, 111.Google Scholar
Pik, R., Marty, B., Carignan, J. and Lavé, J. (2003). Stability of the Upper Nile drainage network (Ethiopia) deduced from (U–Th)/He thermochronometry: implications for uplift and erosion of the Afar plume dome. Earth and Planetary Science Letters 215(1), 7388.Google Scholar
Pilbeam, D. (1974). Hominid-bearing deposits at Kanjera, Nyanza Province, Kenya. Unpublished report.Google Scholar
Pilger, A. and Rösler, A. (Eds.) (1976). Temporal relationships in the tectonic evolution of the Afar Depression (Ethiopia) and the adjacent Afro-Arabian rift system In: Afar Between Continental and Oceanic Rifting: Inter-Union Commission on Geodynamics, Scientific Report No. 16. Stuttgart: Schweizerbart’sche Verlagsbuchhandlung, pp. 125.Google Scholar
Piperno, M. and Bulgarelli-Piperno, G. (1975). First approach to the ecological and cultural significance of the early palaeolithic occupation site of Garba IV at Melka-Kunturé (Ethiopia). Quaternaria 18, 347382.Google Scholar
Piperno, M., Collina, C., Gallotti, R., et al. (2009). Obsidian exploitation and utilization during the Oldowan at Melka Kunture (Ethiopia). In: Hovers, E. and Braun, D.R. (Eds.), Interdisciplinary Approaches to the Oldowan. Dordrecht: Springer, pp. 111128.Google Scholar
Pirrone, C.A., Buatois, L.A. and Bromley, R.G. (2014). Ichnotaxobases for bioerosion trace fossils in bones. Journal of Paleontology 88, 195203.Google Scholar
Pisias, N.G. and Moore, T. Jr. (1981). The evolution of Pleistocene climate: a time series approach. Earth and Planetary Science Letters 52, 450458.Google Scholar
Plana, V. (2004). Mechanisms and tempo of evolution in the African Guineo-Congolian rainforest. Philosophical Transactions: Biological Sciences 359, 15851594.Google Scholar
Plint, T. and Magill, C.R. (2021). Large mammal tracks in 1.8-million-year-old volcanic ash (Tuff IF, Bed I) at Olduvai Gorge, Tanzania. Ichnos 28, 114124.Google Scholar
Plug, I. (1981). Some research results on the late Pleistocene and early Holocene deposits of Bushman Rock Shelter, eastern Transvaal. The South African Archaeological Bulletin 36, 1421.Google Scholar
Plug, I. (1997). Late Pleistocene and Holocene hunter-gatherers in the eastern highlands of South African and Lesotho: a faunal interpretation. Journal of Archaeological Science 24, 715727.Google Scholar
Plug, I. (2004). Resource exploitation: animal use during the Middle Stone Age at Sibudu cave, KwaZulu-Natal: Sibudu cave. South African Journal of Science 100(3–4), 151158.Google Scholar
Plug, I. (2006). Aquatic animals and their associates from the Middle Stone Age levels at Sibudu. Southern African Humanities 18(1), 289299.Google Scholar
Plug, I. and Badenhorst, S. (2001). The distribution of macromammals in southern Africa over the past 30,000 years as reflected in animal remains from archaeological sites. Transvaal Museum Monograph 12, 1234.Google Scholar
Plug, I. and Clark, J.L. (2008). The air: a preliminary report on the birds from Sibudu Cave, KwaZulu-Natal, South Africa. Goodwin Series 10, 133142.Google Scholar
Plug, I. and Engela, R. (1992). The macrofaunal remains from recent excavations at Rose Cottage Cave, Orange Free State. The South African Archaeological Bulletin 47, 1625.Google Scholar
Plug, I. and Keyser, A.W. (1994). Haasgat Cave, a Pleistocene site in the central Transvaal: geomorphological, faunal and taphonomic considerations. Annals of the Transvaal Museum 36(9), 139145.Google Scholar
Plummer, T.W. (1992). Site formation and paleoecology at the early to middle Pleistocene locality of Kanjera. PhD thesis, Yale University.Google Scholar
Plummer, T.W. (2004). Flaked stones and old bones: biological and cultural evolution at the dawn of technology. Yearbook of Physical Anthropology 47, 118164.Google Scholar
Plummer, T. and Bishop, L. (1994). Hominid paleoecology at Olduvai Gorge, Tanzania as indicated by antelope remains. Journal of Human Evolution 27, 4775.Google Scholar
Plummer, T.W. and Bishop, L.C. (2016). Oldowan hominin behavior and ecology at Kanjera South, Kenya. Journal of Anthropological Sciences 94, 2940.Google Scholar
Plummer, T.W. and Potts, R. (1989). Excavations and new findings at Kanjera, Kenya. Journal of Human Evolution 18(3), 269276.Google Scholar
Plummer, T. and Potts, R. (1995). Hominid fossil sample from Kanjera, Kenya: description, provenance, and implications of new and earlier discoveries. American Journal of Physical Anthropology 96(1), 723.Google Scholar
Plummer, T.W. and Stanford, C.B. (2000). Analysis of a bone assemblage made by chimpanzees at Gombe National Park, Tanzania. Journal of Human Evolution 39, 345365.Google Scholar
Plummer, T.W., Kinyua, A.M. and Potts, R. (1994). Provenancing of hominid and mammalian fossils from Kanjera, Kenya, using EDXRF. Journal of Archaeological Science 21(4), 553563.Google Scholar
Plummer, T., Bishop, L.C., Ditchfield, P. and Hicks, J. (1999). Research on late Pliocene Oldowan sites at Kanjera South, Kenya. Journal of Human Evolution, 36(2), 151170.Google Scholar
Plummer, T.W., Bishop, L.C., Ditchfield, P.W., Kingston, J.D. and Hertel, F. (2006). Consensus approach to reconstructing Oldowan hominin paleoecology at Kanjera (Kenya) and Olduvai Gorge Bed I (Tanzania). Presented at the annual meetings of the Society for American Archaeology, held in San Juan, Puerto Rico, 26–30 April.Google Scholar
Plummer, T.W., Bishop, L.C., Ditchfield, P.W., et al. (2009a). The environmental context of Oldowan hominin activities at Kanjera South, Kenya. In: Hovers, E. and Braun, D. (Eds.), Interdisciplinary Approaches to the Oldowan. Dordrecht: Springer, pp. 149160.Google Scholar
Plummer, T.W., Ditchfield, P.W., Bishop, L.C., et al. (2009b). Oldest evidence of toolmaking hominins in a grassland-dominated ecosystem. PLoS One, 4 (9), e7199.Google Scholar
Plummer, T.W., Ferraro, J.V., Louys, J., et al. (2015). Bovid ecomorphology and hominin paleoenvironments of the Shungura Formation, lower Omo River Valley, Ethiopia. Journal of Human Evolution 88, 108126.Google Scholar
Plummer, T.W., Oliver, J.S., Bishop, L.C. and Hertel, F. (2018). Habitat-related variation in Oldowan prey acquisition: comparison of Kanjera South and FLK-Zinj. Presented at the 15th Congress of PanAfrican Archaeological Association, held in Rabat, Morocco, 10–14 September.Google Scholar
Pocock, T.N. (1969). Pleistocene bird fossils from Kromdraai and Sterkfontein. Ostrich 40(S1), 16.Google Scholar
Pocock, T.N. (1985). Plio-Pleistocene mammalian microfauna in southern Africa. Annals of the Geological Survey of South Africa 19, 6567.Google Scholar
Pocock, T.N. (1987). Plio-Pleistocene fossil mammalian microfauna of Southern Africa – a preliminary report including description of two new fossil muroid genera (Mammalia: Rodentia). Palaeontologica Africana 26, 6991.Google Scholar
Pomel, A. (1879). Ossements d’éléphants et d’hippopotames découverts dans une station préhistorique de la plaine d’Eghis (province d’Oran). Bulletin de la Société géologique de France, 3ème sér 7, 4451.Google Scholar
Pomel, A. (1886). Sur la station préhistorique de Ternifine près Mascara. Comptes rendus de l’Association Française pour l’Avancement des Sciences 14, 128.Google Scholar
Pomel, A. (1893–1897). Paléontologie, Monographies (8 fascicles: 1893 – Caméliens et cervidés; 1894 – Les bosélaphes Ray; 1895a – Les antilopes Pallas; 1895b – Les éléphants quaternaires; 1896a – Les rhinocéros quaternaires; 1896b – Les hippopotames; 1897a – Les équidés; 1897b – Les carnassiers). Service de la carte géologique de l’Algérie.Google Scholar
Poole, D.F.G. (1961). Notes on tooth replacement in the Nile crocodile (Crocodilus niloticus). Proceedings of the Zoological Society of London 136, 131140.Google Scholar
Porat, N., Chazan, M., Grün, R. and Horwitz, L.K. (2010). New radiometric ages for the Fauresmith industry from Kathu Pan, southern Africa: implications for the Earlier to Middle Stone Age transition. Journal of Archaeological Science 37, 269283.Google Scholar
Porraz, G., Val, A., Dayet, L., et al. (2015). Bushman Rock Shelter (Limpopo, South Africa): a perspective from the edge of the Highveld. The South African Archaeological Bulletin 70, 166179.Google Scholar
Porraz, G., Schmid, V.C., Miller, C.E., et al. (2016). Update on the 2011 excavation at Elands Bay Cave (South Africa) and the Verlorenvlei stone age. Southern African Humanities 29(1), 3368.Google Scholar
Potts, R. (1984). Home bases and early hominids: reevaluation of the fossil record at Olduvai Gorge suggests that the concentrations of bones and stone tools do not represent fully formed campsites but an antecedent to them. American Scientist 72, 338347.Google Scholar
Potts, R. (1988). Early Hominin Activities at Olduvai. New York: Aldine de Gruyter.Google Scholar
Potts, R. (1989). Olorgesailie: new excavations and findings in Early and Middle Pleistocene contexts, southern Kenya rift valley. Journal of Human Evolution 18, 477484.Google Scholar
Potts, R. (1991). Why the Oldowan? Plio-Pleistocene toolmaking and the transport of resources. Journal of Anthropological Research 47, 153176.Google Scholar
Potts, R. (1994). Variables vs. models of early Pleistocene hominid land use. Journal of Human Evolution 27, 724.Google Scholar
Potts, R. (1996). Evolution and climate variability. Science 273, 922923.Google Scholar
Potts, R. (1998a). Variability selection in hominid evolution. Evolutionary Anthropology 7, 8196.Google Scholar
Potts, R. (1998b). Environmental hypothesis of hominin evolution. Yearbook of Physical Anthropology 41, 93136.Google Scholar
Potts, R. (2007). Environmental hypotheses of Pliocene human evolution. In: Bobe, R., Alemseged, Z. and Behrensmeyer, A.K. (Eds.), Hominin Environments in the East African Pliocene: An Assessment of the Faunal Evidence. Dordrecht, Springer: pp. 2549.Google Scholar
Potts, R. (2013). Hominin evolution in settings of strong environmental variability. Quaternary Science Reviews 73, 113.Google Scholar
Potts, R. and Behrensmeyer, A.K. (1992). Late Cenozoic terrestrial ecosystems. In: Behrensmeyer, A.K., Damuth, J.D., DiMichele, W.A., et al. (Eds.), Terrestrial Ecosystems Through Time: Evolutionary Paleoecology of Terrestrial Plants and Animals. Chicago: University of Chicago Press, pp. 419541.Google Scholar
Potts, R. and Deino, A. (1995). Mid-Pleistocene change in large mammal faunas of the southern Kenya rift. Quaternary Research 43, 106113.Google Scholar
Potts, R. and Faith, J.T. (2015). Alternating high and low climate variability: the context of natural selection and speciation in Plio-Pleistocene hominin evolution. Journal of Human Evolution 87, 520.Google Scholar
Potts, R. and Shipman, P. (1981). Cutmarks made by stone tools on bones from Olduvai Gorge, Tanzania. Nature 291, 577580.Google Scholar
Potts, R., Shipman, P. and Ingall, E. (1988). Taphonomy, paleoecology and hominids of Lainyamok, Kenya. Journal of Human Evolution 17, 597614.Google Scholar
Potts, R., Ditchfield, P., Hicks, J. and Deino, A. (1997). Paleoenvironments of late Miocene and early Pliocene strata of Kanam, western Kenya. American Journal of Physical Anthropology Supplement 24, 188189.Google Scholar
Potts, R., Behrensmeyer, A.K. and Ditchfield, P. (1999). Paleolandscape variation in early Pleistocene hominid activities: Members 1 and 7, Olorgesailie Formation, Kenya. Journal of Human Evolution 37, 747788.Google Scholar
Potts, R., Behrensmeyer, A.K., Deino, A., Ditchfield, P. and Clark, J. (2004). Small mid-Pleistocene hominin associated with East African Acheulean technology. Science 305, 7578.Google Scholar
Potts, R., Behrensmeyer, A.K., Faith, J.T., et al. (2018). Environmental dynamics during the onset of the Middle Stone Age in eastern Africa. Science 360, 8690.Google Scholar
Pound, M.J., Haywood, A.M., Salzmann, U., et al. (2011). A Tortonian (Late Miocene, 11.61–7.25 Ma) global vegetation reconstruction. Palaeogeography, Palaeoclimatology, Palaeoecology 300, 2945.Google Scholar
Powers, D.W. (1980). Geology of Mio-Pliocene sediments of the lower Kerio river valley, Kenya. PhD thesis, Department of Geological and Geophysical Sciences, Princeton University, Princeton.Google Scholar
Prat, S., Brugal, J.-P., Roche, H. and Texier, P.-J. (2003). Nouvelles découvertes de dents d’hominidés dans le membre Kaitio de la Formation de Nachukui (1,65–1,9 million d’années) à l’Ouest du Lac Turkana (Kenya). Comptes Rendus Palévol 2(8), 685693.Google Scholar
Prat, S, Brugal, J.-P., Tiercelin, J.-J., et al. (2005). First occurrence of early Homo in the Nachukui Formation (West Turkana, Kenya) at 2.3–2.4 Myr. Journal of Human Evolution, 49, 230240.Google Scholar
Prendergast, M., Luque, L., Domínguez-Rodrigo, M., et al. (2007). New excavations at Mumba Rockshelter, Tanzania. Journal of African Archaeology 5, 217243.Google Scholar
Presnyakova, D., Braun, D.R., Conard, N.J., et al. (2018). Site fragmentation, hominin mobility and LCT variability reflected in the early Acheulean record of the Okote Member, at Koobi Fora, Kenya. Journal of Human Evolution 125, 159180.Google Scholar
Profico, A., Di Vincenzo, F., Gagliardi, L., Piperno, M. and Manzi, G. (2016). Filling the gap. Human cranial remains from Gombore II (Melka Kunture, Ethiopia; ca. 850 ka) and the origin of Homo heidelbergensis. Journal of Anthropological Science 94, 124.Google Scholar
Pruetz, J.D. (2007). Evidence of cave use by savanna chimpanzees (Pan troglodytes verus) at Fongoli, Senegal: implications for thermoregulatory behavior. Primates 48, 316319.Google Scholar
Püschel, H.P., Bertrand, O.C., O’Reilly, J.E., Bobe, R. and Püschel, T. A. (2021). Divergence-time estimates for hominins provide insight into encephalization and body mass trends in human evolution. Nature Ecology & Evolution 5, 808819.Google Scholar
Quade, J. and Wynn, J.G. (Eds.) (2008). The Geology of Early Humans in the Horn of Africa. Geological Society of America Special Paper 446. Boulder: The Geological Society of America.Google Scholar
Quade, J., Levin, N., Semaw, S., et al. (2004). Paleoenvironments of the earliest stone toolmakers, Gona, Ethiopia. Geological Society of America Bulletin 116, 15291544.Google Scholar
Quade, J., Levin, N.E., Simpson, S.W., et al. (2008). The geology of Gona, Afar, Ethiopia. In: Quade, J. and Wynn, J.G. (Eds.). The Geology of Early Humans in the Horn of Africa. Geological Society of America Special Paper 446. Boulder: The Geological Society of America,pp. 131.Google Scholar
Quick, L. (2013). Late Quaternary palaeoenvironments of the southern Cape, South Africa palynological evidence from three coastal wetlands. Doctoral dissertation, University of Cape Town.Google Scholar
Quick, L.J., Carr, A.S., Meadows, M.E., et al. (2015). A late Pleistocene–Holocene multi‐proxy record of palaeoenvironmental change from Still Bay, southern Cape Coast, South Africa. Journal of Quaternary Science 30(8), 870885.Google Scholar
Quinn, R.L. and Lepre, C.J. (2020). Revisiting the pedogenic carbonate isotopes and paleoenvironmental interpretation of Kanapoi. Journal of Human Evolution 140, 102549.Google Scholar
Quinn, R.L., Lepre, C.J., Wright, J.D. and Feibel, C.S. (2007). Paleogeographic variations of pedogenic carbonate δ13C values from Koobi Fora, Kenya: implications for floral compositions of Plio-Pleistocene hominin environments. Journal of Human Evolution 53, 560573.Google Scholar
Quinn, R.L., Lepre, C.J., Feibel, C.S., et al. (2013). Pedogenic carbonate stable isotopic evidence for wooded habitat preference of early Pleistocene tool makers in the Turkana Basin. Journal of Human Evolution 65(1), 6578.Google Scholar
Quinn, R.L., Lewis, J., Brugal, J.P., et al. (2020). Pliocene environmental change, hominin dietary niche expansion, and the origins of stone tools. Palaeogeography, Palaeoclimatology, Palaeoecology 562, 110074.Google Scholar
R Development Core Team. (2011). R: A Language and Environment for Statistical Computing. Vienna: R Foundation for Statistical Computing.Google Scholar
R Development Core Team. (2019). R: A Language and Environment for Statistical Computing. Vienna: R Foundation for Statistical Computing.Google Scholar
Raaum, R.L., Sterner, K.N., Noviello, C.M., Stewart, C.B. and Disotell, T.R. (2005). Catarrhine primate divergence dates estimated from complete mitochondrial genomes: concordance with fossil and nuclear DNA evidence. Journal of Human Evolution 48(3), 237257.Google Scholar
Rage, J.C. and Bailon, S. (2011). Amphibia and Squamata. In: Harrison, T. (Ed.), Paleontology and Geology of Laetoli: Human Evolution in Context. Vol. 2: Fossil Hominins and the Associated Fauna. Dordrecht: Springer, pp. 467478.Google Scholar
Raia, P., Piras, P. and Kotsakis, T. (2005). Turnover pulse or Red Queen? Evidence from the large mammal communities during the Plio-Pleistocene of Italy. Palaeogeography, Palaeoclimatology, Palaeoecology 221, 293312.Google Scholar
Raichlen, D.A., Pontzer, H. and Sockol, M.D. (2008). The Laetoli footprints and early hominin locomotor kinematics. Journal Human Evolution 54, 112117.Google Scholar
Raichlen, D.A., Gordon, A.D., Harcourt-Smith, W.E.H., Foster, A.D. and Haas, W.R. (2010). Laetoli footprints preserve earliest direct evidence of human-like bipedal biomechanics. PLoS ONE 5(3), e9769.Google Scholar
Randall, R.M. (1972). The Hyaenidae of the Limeworks Site Makapansgat. Unpublished BSc dissertation, University of the Witwatersrand.Google Scholar
Randall, R.M. (1981). Fossil Hyaenidae from the Makapansgat Limeworks Deposit, South Africa. Palaeontologica Africana 24, 7585.Google Scholar
Rasmussen, C., Reichenbacher, B., Lenz, O., et al. (2017). Middle–late Miocene palaeoenvironments, palynological data and a fossil fish Lagerstätte from the Central Kenya Rift (East Africa). Geological Magazine 154, 2456.Google Scholar
Rasmussen, D.T., Pickford, M., Mein, P., Senut, B. and Conroy, G.C. (1996). Earliest known procaviid hyracoid from the Late Miocene of Namibia. Journal of Mammalogy 77(3), 745754.Google Scholar
Raup, D.M. (1977). Systematists follow the fossils. Paleobiology 3(3), 328329.Google Scholar
Ravelo, A.C., Andreasen, D.H., Lyle, M., Olivarez Lyle, A. and Wara, M.W. (2004). Regional climate shifts caused by gradual global cooling in the Pliocene epoch. Nature 429, 263267.Google Scholar
Ravelo, A.C., Dekens, P.S. and McCarthy, M. (2006). Evidence for El Niño-like conditions during the Pliocene. GSA Today 16, 411.Google Scholar
Raymo, M.E. and Huybers, P. (2008). Unlocking the mysteries of the ice ages. Nature 451, 284285.Google Scholar
Raynal, J.-P. and Kieffer, G. (2004). Lithology, dynamism and volcanic successions at Melka Kunture (Upper Awash, Ethiopia). In: Chavaillon, J. and Piperno, M. (Eds.), Studies on the Early Paleolithic site of Melka Kunture, Ethiopia. Florence: Istituto Italiano di Preistoria e Protostoria, pp. 111135.Google Scholar
Raynal, J.-P. and Texier, J.-P. (1989). Découverte d’Acheuléen ancien dans la carrière Thomas 1 à Casablanca, et problème de l’ancienneté de la présence humaine au Maroc. Comptes Rendus de l’Académie des Sciences 308, 17431749.Google Scholar
Raynal, J.-P., Texier, J.-P., Geraads, D. and Sbihi-Alaoui, F.-Z. (1990). Un nouveau gisement paléontologique plio–pléistocène Afrique du Nord: Ahl Al Oughlam (ancienne carrière Deprez) à Casablanca (Maroc). Comptes Rendus de l’Académie des Sciences Paris II 310, 315320.Google Scholar
Raynal, J.-P., Geraads, D., Magoga, L., et al. (1993). La grotte des Rhinocéros (Carrière Oulad Hamida 1, anciennement Thomas III, Casablanca), nouveau site acheuléen du Maroc atlantique. Comptes Rendu de l’Académie des Sciences Paris 316, 14771483.Google Scholar
Raynal, J.-P., Magoga, L., Sbihi-Alaoui, F.Z. and Geraads, D. (1995). The earliest occupation of Atlantic Morocco: the Casablanca evidence. In: Roebroeks, W., and van Kolfschoten, T. (Eds.), The Earliest Occupation of Europe. Leiden: University of Leiden, pp. 255262.Google Scholar
Raynal, J.-P., Lefèvre, D., Geraads, D. and El Graoui, M. (1999). Contribution du site paléontologique de Lissasfa (Casablanca, Maroc) à une nouvelle interprétation du Mio-Pliocène de la Meseta. Comptes Rendus de l’Académie des Sciences, Sciences de la Terre et des Planètes 329, 617622.Google Scholar
Raynal, J.-P., Sbihi-Alaoui, F.Z., Geraads, D., Magoga, L. and Mohib, A. (2001). The earliest occupation of North-Africa: the Moroccan perspective. Quaternary International 75, 6575.Google Scholar
Raynal, J.P., Sbihi-Alaoui, F.Z., Magoga, L., Mohib, A. and Zouak, M. (2002). Casablanca and the early occupation of north-atlantic Morocco. Quaternaire 13, 6577.Google Scholar
Raynal, J.-P., Kieffer, G. and Bardin, G. (2004). Garba IV and the Melka Kunture Formation. A preliminary lithostratigraphic approach. In: Chavaillon, J. and Piperno, M. (Eds.), Studies on the Early Paleolithic site of Melka Kunture, Ethiopia. Florence: Istituto Italiano di Preistoria e Protostoria, pp. 137166.Google Scholar
Raynal, J.-P., Amani, F., Geraads, D., et al. (2008). Felids Cave, a new Upper Pleistocene Palaeolithic site at Casablanca (Morocco). L’Anthropologie 112, 182200.Google Scholar
Raynal, J.P., Sbihi-Alaoui, F.Z., Mohib, A. and Geraads, D. (2009). Préhistoire ancienne au Maroc atlantique: bilan et perspectives régionales. Bulletin d’Archéologie marocaine 21, 954.Google Scholar
Raynal, J.-P., Sbihi-Alaoui, F.Z., Mohib, A., et al. (2010). Hominid cave at Thomas Quarry I (Casablanca, Morroco): recent findings and their context. Quaternary International 223–224, 369382.Google Scholar
Raynal, J.-P., Sbihi-Alaoui, F.Z., Mohib, A., et al. (2011). Contextes et âge des nouveaux restes dentaires humains du Pléistocène moyen de la carrière Thomas 1 à Casablanca (Maroc). Bulletin de la Société préhistorique Française 108, 645669.Google Scholar
Raynal, J.-P., Mohib, A. and Lefèvre, D. (2016a). Casablanca des origins. In: Raynal, J.-P. and Mohib, A. (Eds.), Préhistoire de Casablanca. 1 – La Grotte des Rhinocéros (fouilles 1991 et 1996). Villes et Sites archéologiques du Maroc, volume 6. Rabat: Ministère de la Culture, INSAP, pp. 1133.Google Scholar
Raynal, J.-P., Sbihi-Alaoui, F.Z., Mohib, A., et al. (2016b). Nouveaux restes dentaires humains dans la Grotte à Hominidés de la carrière Thomas I à Casablanca (Maroc): contexte et datation directe par ablation laser ICP-MS. In: Akerraz, A., Ettahiri, A.S., and Kbiri-Alaoui, M. (Eds.), Hommage à Joudia Hassar-Benslimane. Rabat: INSAP, Tome 1, pp. 5971.Google Scholar
Raynal, J.-P., Lefèvre, D., Geraads, D., El Graoui, M. and Rué, M. (2016c). La Grotte des Rhinocéros (Casablanca, Maroc): le remplissage et son âge. In: J.-P. Raynal and A. Mohib (Eds.), Préhistoire de Casablanca. 1 ‒ La Grotte des Rhinocéros (fouilles 1991 et 1996). Villes et Sites Archéologiques du Maroc, volume 6. Rabat: Ministère de la Culture, Rabat: INSAP, Tome 1, pp. 7986.Google Scholar
Raynal, J.-P., Gallotti, R., Mohib, A., Fernandes, P. and Lefèvre, D. (2017). The western quest, first and second regional Acheuleans at Thomas-Oulad Hamida quarries (Casablanca, Morocco). In: Wojtczak, D. et al. (Eds.), Vocation Préhistoire. Hommage à Jean-Marie Le Tensorer. Liège: ERAUL, pp. 309322.Google Scholar
Rayner, R.J., Moon, B. and Masters, J.C. (1993). The Makapansgat australopithecine environment. Journal of Human Evolution 24, 219231.Google Scholar
Reck, H. (1914). Erste vorläufige Mitteilung über den Fund eines fossilen Menschenskeletts aus Zentralafrika. Sitzungsberichte der Gesellschaft Naturforschender Freunde zu Berlin 3, 81e95.Google Scholar
Rector, A.L. and Reed, K.E. (2010). Middle and Late Pleistocene faunas of Pinnacle Point and their paleoecological implications. Journal of Human Evolution 59, 340357.Google Scholar
Rector, A.L. and Verrelli, B.C. (2010). Glacial cycling, large mammal community composition, and trophic adaptations in the Western Cape, South Africa. Journal of Human Evolution 58, 90102.Google Scholar
Rector, A.L., Reed, K.E., Meacham, S. and Steininger, C. (2016). The large mammal community from Cooper’s D and its significance for Paranthropus robustus ecology. American Journal of Physical Anthropology 159(S62), 264.Google Scholar
Reda, H.G., Lazagabaster, I.A. and Haile-Selassie, Y. (2019). Newly discovered crania of Nyanzachoerus jaegeri (Tetraconodontinae, Suidae, Mammalia) from the Woranso-Mille (Ethiopia) and reappraisal of its generic status. Journal of Mammalian Evolution, 26(2), 179199.Google Scholar
Reed, D. (2007). Serengeti micromammals and their implications for Olduvai paleoenvironments. In: Bobe, R., Alemseged, Z. and Behrensmeyer, A.K. (Eds.), Hominin Environments in the East African Pliocene: An Assessment of the Faunal Evidence. Dordrecht: Springer, pp. 217255.Google Scholar
Reed, D. (2011a). Serengeti micromammal communities and the paleoecology of Laetoli, Tanzania. In: Harrison, T. (Ed.), Paleontology and Geology of Laetoli: Human Evolution in Context. Vol. 1: Geology, Geochronology, Paleoecology, and Paleoenvironment. Dordrecht: Springer, pp. 253263.Google Scholar
Reed, D.N. (2011b). New murid (Mammalia, Rodentia) fossils from a late Pliocene (2.4 Ma) locality, Hadar AL 894, Afar Region, Ethiopia. Journal of Vertebrate Paleontology 31, 13261337.Google Scholar
Reed, D. and Denys, C. (2011). The taphonomy and paleoenvironmental implications of the Laetoli micromammals. In: Harrison, T. (Ed.), Paleontology and Geology of Laetoli: Human Evolution in Context. Vol. 1: Geology, Geochronology, Paleoecology, and Paleoenvironment. Dordrecht: Springer, pp. 265278.Google Scholar
Reed, D.N. and Geraads, D. (2012). Evidence for a late Pliocene faunal transition based on a new rodent assemblage from Oldowan locality Hadar AL 894, Afar Region, Ethiopia. Journal of Human Evolution 62, 328337.Google Scholar
Reed, D., Barr, W.A., McPherron, S.P., et al. (2015). Digital data collection in paleoanthropology. Evolutionary Anthropology 24, 238249.Google Scholar
Reed, K.E. (1996). The paleoecology of Makapansgat and other African Plio-Pleistocene hominid localities. PhD thesis, Anthropological Sciences. State University of New York, Stony Brook.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
Reed, K.E. (1998). Using large mammal communities to examine ecological and taxonomic structure and predict vegetation in extant and extinct assemblages. Paleobiology 24, 384408.Google Scholar
Reed, K.E. (2008). Paleoecological patterns at the Hadar hominin site, Afar Regional State, Ethiopia. Journal of Human Evolution 54, 743768.Google Scholar
Reed, K.E. (2013). Multiproxy paleoecology: reconstructing evolutionary context in paleoanthropology. In: Begun, D.R. (Ed.), A Companion to Paleoanthropology. Oxford: Wiley Blackwell, pp. 203225.Google Scholar
Reed, K.E. and Bibi, F. (2010). Fossil Tragelaphini (Artiodactyla: Bovidae) from the Late Pliocene Hadar Formation, Afar Regional State, Ethiopia. Journal of Mammalian Evolution 18, 5769.Google Scholar
Reed, K.E. and Bidner, L.R. (2004). Primate communities: past, present, and possible future. Yearbook of Physical Anthropology 47, 239.Google Scholar
Reed, K.E. and Fish, J.L. (2005). Tropical and temperate seasonal influences on human evolution. In: Brockman, D.K. and van Schaik, C.P. (Eds.), Seasonality in Primates: Studies of Living and Extinct Human and Non-Human Primates. Cambridge: Cambridge University Press, pp. 489518.Google Scholar
Reed, K.E. and Rector, A.L. (2006). African Pliocene paleoecology: hominin habitats, resources, and diets. In: Ungar, P. (Ed.), Evolution of the Human Diet: The Known, the Unknown, and the Unknowable. Oxford: Oxford University Press, pp. 262288.Google Scholar
Reed, K. and Russack, S.M. (2009). Tracking ecological change in relation to the emergence of Homo near the Plio-Pleistocene boundary. In: Grine, F.E., Fleagle, J.G. and Leakey, R.E. (Eds.), The First Humans: Origin and Early Evolution of the Genus Homo. Dordrecht: Springer, pp. 159171.Google Scholar
Rein, T. (2010). Locomotor function and phylogeny: implications for interpreting the hominoid fossil record. PhD thesis, New York University.Google Scholar
Remane, A. (1950). Die Zähne des Meganthropus africanus. Zeitschrift für Morphologie und Anthropologie 42, 311329.Google Scholar
Remane, A. (1954). Structure and relationships of Meganthropus africanus. American Journal of Physical Anthropology 12, 123126.Google Scholar
Rémy, J.A. (1976). Présence de Deinotherium sp. Kaup (Proboscidea, Mammalia) dans la faune miocène de Beni Mellal, Maroc. Géologie Méditerranéenne 3, 109114.Google Scholar
Renne, P., Walter, R., Verosub, K., Sweitzer, M. and Aronson, J. (1993). New data from Hadar (Ethiopia) support orbitally tuned time-scale to 3.3 Ma. Geophysical Research Letters 20, 10671070.Google Scholar
Renne, P.R., Swisher, C.C., Deino, A.L., et al. (1998). Intercalibration of standards, absolute ages and uncertainties in 40Ar/39Ar dating. Chemical Geology 145, 117152.Google Scholar
Renne, P.R., WoldeGabriel, G., Hart, W.K., Heiken, G. and White, T. D. (1999). Chronostratigraphy of the Miocene–Pliocene Sagantole Formation, Middle Awash Valley, Afar rift, Ethiopia. Geological Society of America Bulletin 111(6), 869885.Google Scholar
Retallack, G.J. (1991). Untangling the effects of burial alteration and ancient soil formation. Annual Review of Earth and Planetary Sciences 19(1), 183206.Google Scholar
Retallack, G.J. (2001). Soils of the Past: An Introduction to Paleopedology. Oxford: Blackwell Science.Google Scholar
Retallack, G.J., Bestland, E.A. and Dugas, D.P. (1995). Miocene paleosols and habitats of Proconsul on Rusinga Island, Kenya. Journal of Human Evolution 29(1), 5391.Google Scholar
Reti, J.S. (2016). Quantifying Oldowan stone tool production at Olduvai Gorge, Tanzania. PLoS ONE 11, e0147352.Google Scholar
Reynard, J.P. and Wurz, S. (2020). The palaeoecology of Klasies River, South Africa: an analysis of the large mammal remains from the 1984–1995 excavations of Cave 1 and 1A. Quaternary Science Reviews, 237, 106301.Google Scholar
Reynolds, S.C. (2006). Temporal changes in vegetation and mammalian communities during Oxygen Isotope Stage 3 at Sibudu Cave. Southern African Humanities 18(1), 301314.Google Scholar
Reynolds, S.C. (2007a). Temporal variation in Plio-Pleistocene Antidorcas (Mammalia: Bovidae) horncores: the case from Bolt’s Farm and why size matters. South African Journal of Science, 103(1–2), 4750.Google Scholar
Reynolds, S.C. (2007b). Mammalian body size changes and Plio-Pleistocene environmental shifts: implications for understanding hominin evolution in eastern and southern Africa. Journal of Human Evolution 53, 528548.Google Scholar
Reynolds, S.C. (2010). Where the wild things were: spatial and temporal distribution of carnivores in the Sterkfontein Valley in relation to the accumulation of mammalian assemblages. Journal of Taphonomy 8(2–3), 233257.Google Scholar
Reynolds, S. and Kibii, J. (2011). Sterkfontein at 75: review of palaeoenvironments, fauna and archaeology from the hominin site of Sterkfontein (Gauteng Province, South Africa). Palaeontologica Africana 46, 5988.Google Scholar
Reynolds, S.C., Vogel, J.C., Clarke, R.J. and Kuman, K.A. (2003). Preliminary results of excavations at Lincoln Cave, Sterkfontein. South African Journal of Science 99, 286288.Google Scholar
Reynolds, S.C., Clark, R.J. and Kuman, K.A. (2007). The view from Lincoln Cave: mid- to late Pleistocene fossil deposits from Sterkfontein hominid site, South Africa. Journal of Human Evolution 53, 260271.Google Scholar
Reynolds, S., Bailey, G. and King, G. (2011). Landscapes and their relation to hominin habitats: case studies from Australopithecus sites in eastern and southern Africa. Journal of Human Evolution 60, 281298.Google Scholar
Reynolds, S.C., Wilkinson, D.M., Marston, C.G. and O’Regan, H.J. (2015). The ‘mosaic habitat’ concept in human evolution: past and present. Transactions of the Royal Society of South Africa 70(1), 5769.Google Scholar
Reynolds, S.C., Marston, C.G., Hassani, H., King, G.C. and Bennett, M.R. (2016). Environmental hydro-refugia demonstrated by vegetation vigour in the Okavango Delta, Botswana. Scientific Reports, 6.Google Scholar
Rhodes, E.J., Raynal, J.-P., Geraads, D. and Sbihi-Alaoui, F.Z. (1994). Premières dates RPE pour l’Acheuléen du Maroc atlantique (Grotte des Rhinocéros, Casablanca). Comptes rendus de l’Académie de Sciences, Paris (II) 319, 11091115.Google Scholar
Rhodes, E.J., Singarayer, J.S., Raynal, J.-P., Westaway, K.E. and Sbihi-Alaoui, F.Z. (2006). New age estimations for the Palaeolithic assemblages and Pleistocene succession of Casablanca, Morocco. Quaternary Science Reviews 25, 25692585.Google Scholar
Rich, P.V. (1980). Preliminary report on the fossil avian remains from late Tertiary sediments at Langebaanweg (Cape Province), South Africa. South African Journal of Science 76(4), 166170.Google Scholar
Richard, M., Chazan, M. and Porat, N. (2022). Single grain TT-OSL ages for the Earlier Stone Age site of Bestwood 1 (Northern Cape Province, South Africa). Quaternary International 614, 16–22.Google Scholar
Richmond, B.G. and Jungers, W. L. (2008). Orrorin tugenensis femoral morphology and the evolution of hominin bipedalism. Science, 319(5870),16621665.Google Scholar
Richmond, B.G., Green, D.J., Lague, M.R., et al. (2020). The upper limb of Paranthropus boisei from Ileret, Kenya. Journal of Human Evolution 141, 102727.Google Scholar
Richter, D., Grün, R., Joannes-Boyau, R., et al. (2017). The age of the Homo sapiens fossils from Jebel Irhoud (Morocco) and the origins of the Middle Stone Age. Nature 546, 293296.Google Scholar
Rieux, A., Eriksson, A., Li, M., et al. (2014). Improved calibration of the human mitochondrial clock using ancient genomes. Molecular Biology and Evolution, 31(10), 27802792.Google Scholar
Riga, A., Mori, T., Pickering, T.R., Moggi‐Cecchi, J. and Menter, C.G. (2019). Ages‐at‐death distribution of the early Pleistocene hominin fossil assemblage from Drimolen (South Africa). American Journal of Physical Anthropology 168(3), 632636.Google Scholar
Rightmire, G.P. (1979). Cranial remains of Homo erectus from Beds II and IV, Olduvai Gorge, Tanzania. American Journal of Physical Anthropology 51, 99115.Google Scholar
Rightmire, G.P. (1998). Human evolution in the middle Pleistocene: the role of Homo heidelbergensis. Evolutionary Anthropology 6, 218227.Google Scholar
Rightmire, G.P. (2013). Homo erectus and Middle Pleistocene hominins: brain size, skull form, and species recognition. Journal of Human Evolution 65, 223252.Google Scholar
Rightmire, G.P. and Deacon, H.J. (1991). Comparative studies of late Pleistocene human remains from Klasies River mouth, South Africa. Journal of Human Evolution 20(2), 131156.Google Scholar
Ring, U. and Betzler, C. (1995). Geology of the Malawi Rift: kinematic and tectonosedimentary background to the Chiwondo Beds, northern Malawi. Journal of Human Evolution 28, 721.Google Scholar
Ring, U., Albrecht, C. and Schrenk, F. (2018). The East African Rift System: tectonics, climate, and biodiversity. In: Hoorn, C., Perrigo, A. and Antonelli, A. (Eds.), Mountains, Climate and Biodiversity. Hoboken: Wiley Blackwell.Google Scholar
Ritchie, M.E. (2008). Global environmental changes and their impact on the Serengeti. In: Sinclair, A.R.E., Packer, C., Mduma, S.A.R. and Fryxell, J.M. (Eds.), Serengeti III: Human Impacts on Ecosystem Dynamics. Chicago: University of Chicago Press, pp. 183208.Google Scholar
Rito, T., Richards, M.B., Fernandes, V., et al. (2013). The first modern human dispersals across Africa. PLoS ONE 8, e80031.Google Scholar
Robbins, L.M. (1987). Hominid footprints from Site G. In: Leakey, M.D. and Harris, J.M. (Eds.), Laetoli: A Pliocene Site in Northern Tanzania. Oxford: Clarendon Press, pp. 497502.Google Scholar
Roberts, D.L. (2003). Vertebrate trackways in Pleistocene coastal calcareous aeolianites South Africa. Abstract, XVI Inqua Congess. Reno, Nevada: Desert Research Institute, p. 96.Google Scholar
Roberts, D.L., Bateman, M.D., Murray-Wallace, C.V., Carr, A.S. and Holmes, P.J. (2008). Last interglacial fossil elephant trackways dated by OSL/AAR in coastal aeolianites, Still Bay, South Africa. Palaeogeography, Palaeoclimatology, Palaeoecology 257(3), 261279.Google Scholar
Roberts, D.L., Bateman, M.D., Murray-Wallace, C.V., Carr, A.S. and Holmes, P.J. (2009). West coast dune plumes: climate driven contrasts in dunefield morphogenesis along the western and southern South African coasts. Palaeogeography, Palaeoclimatology, Palaeoecology 271(1–2), 2438.Google Scholar
Roberts, D.L., Matthews, T., Herries, A.I., et al. (2011). Regional and global context of the Late Cenozoic Langebaanweg (LBW) palaeontological site: West Coast of South Africa. Earth-Science Reviews 106(3–4), 191214.Google Scholar
Robinson, C. (2011). Giraffidae. In: Harrison, T. (Ed.), Paleontology and Geology of Laetoli: Human Evolution in Context. Vol. 2: Fossil Hominins and the Associated Fauna. Dordrecht: Springer, pp. 339362.Google Scholar
Robinson, J.R. and Wadley, L. (2018). Stable isotope evidence for (mostly) stable local environments during the South African Middle Stone Age from Sibudu, KwaZulu-Natal. Journal of Archaeological Science 100, 3244.Google Scholar
Robinson, J.R., Rowan, J., Campisano, C.J., Wynn, J.G. and Reed, K.E. (2017). Late Pliocene environmental change during the transition from Australopithecus to Homo. Nature Ecology & Evolution 1, 0159.Google Scholar
Robinson, J.T. (1952). The australopithecine-bearing deposits of the Sterkfontein area. Annals of theTransvaal Museum 22(1), 119.Google Scholar
Robinson, J.T. (1953a). Meganthropus, australopithecines and hominids. American Journal of Physical Anthropology 11, 138.Google Scholar
Robinson, J.T. (1953b). Telanthropus and its phylogenetic significance. American Journal of Physical Anthropology 11, 445501.Google Scholar
Robinson, J.T. (1954). The genera and species of the Australopithecinae. American Journal of Physical Anthropology 12, 181200.Google Scholar
Robinson, J.T. (1955). Further remarks on the relationship between “Meganthropus” and australopithecines. American Journal of Physical Anthropology 13, 429445.Google Scholar
Robinson, J.T. (1956). The Dentition of the Australopithecinae. Transvaal Museum Memoir No. 9. Pretoria: Transvaal Museum.Google Scholar
Robinson, J.T. (1961). The australopithecines and their bearing on the origin of man and tool-making. South African Journal of Science 57, 313.Google Scholar
Robinson, J.T. (1962). Sterkfontein stratigraphy and the significance of the Extension Site. South African Archaeological Bulletin 17, 87107.Google Scholar
Robinson, J.T. (1963). Adaptive radiation in the australopithecines and the origin of man. In: Howell, F.C. and Bourlière, F. (Eds.), African Ecology and Human Evolution. Chicago: Aldine, pp.385416.Google Scholar
Robinson, P. (1972). Pachytragus solignaci, a new species of caprine bovid from the late Miocene Beglia Formation of Tunisia. Notes du Service géologique de Tunisie 37, 7394.Google Scholar
Robinson, P. (1986) Very hypsodont antelopes from the Beglia Formation (central Tunisia), with a discussion of the Rupicaprinae. Rocky Mountain Geology 24(special paper 3), 305315.Google Scholar
Robinson, P. and Black, C.C. (1974). Vertebrate faunas from the Neogene of Tunisia. Annals of the Geological Survey of Egypt 4, 319332.Google Scholar
Robinson, P., Black, C. and Craig, C. (1969). Note préliminaire sur les vertébrés fossiles du Vindobonien (formation Beglia) du Bled Douarah, Gouvernorat de Gafsa, Tunisie. Notes du Service géologique de Tunisie 31, 6770.Google Scholar
Roche, C. (2004). ‘The Springbok … Drink the Rain’s Blood’: indigenous knowledge and its use in environmental history – the case of the /Xam and an understanding of Springbok Treks. South African Historical Journal 53, 122.Google Scholar
Roche, H. (2005). From simple flaking to shaping: stone knapping evolution among early Hominids. In: Roux, V. and Bril, B. (Eds.), Stone Knapping: The Necessary Conditions for a Uniquely Hominid Behaviour. Cambridge: Mc Donald Institute Monograph Series, pp. 3548.Google Scholar
Roche, H. (2011). The archaeology of human origins: the contribution of West Turkana, Kenya. In: Sept, J. and Pilbeam, D. (Eds.), Casting the Net Wide, Papers in Honor of Glynn Isaac. American School of Prehistoric Research Monograph series. Cambridge, MA: Oxbow Books, pp. 7592.Google Scholar
Roche, H. and Kibunjia, M. (1996). Contribution of the West Turkana plio-pleistocene to the archaeology of the Lower Omo/Turkana Basin. Kaupia 6, 2730.Google Scholar
Roche, H., Delagnes, A., Brugal, J.-P., et al. (1999). Evidence for early hominids lithic production and technical skill at 2.3 Myr, West Turkana, Kenya. Nature 399, 5760.Google Scholar
Roche, H., Brugal, J.-P., Delagnes, A., et al. (2003). Les sites archéologiques plio-pléistocènes de la Formation de Nachukui (Ouest Turkana, Kenya): bilan préliminaire 1996–2000. Comptes Rendus Palévol 2(8), 663673.Google Scholar
Roche, D., Ségalen, L., Senut, B. and Pickford, M. (2013). Stable isotope analyses of tooth enamel carbonate of large herbivores from the Tugen Hills deposits: palaeoenvironmental context of the earliest Kenyan hominids. Earth and Planetary Science Letters 381, 3951.Google Scholar
Rodriguez, J. (1999). Use of cenograms in mammalian palaeoecology. A critical review. Lethaia 32(4), 331347.Google Scholar
Rogers, M.J. and Semaw, S. (2009). From nothing to something: the appearance and context of the earliest archaeological record. In: Camps, M. and Chauhan, P. (Eds.), Sourcebook of Paleolithic Transitions. New York: Springer, pp. 155171.Google Scholar
Rogers, M.J., Harris, J.W.K. and Feibel, C.S. (1994). Changing patterns of land use by Plio-Pleistocene hominids in the Lake Turkana Basin. Journal of Human Evolution 27, 139158.Google Scholar
Roman, D., Campisano, C., Quade, J., et al. (2008). Composite tephrostratigraphy of the Dikika, Gona, Hadar, and Ledi-Geraru project areas, northern Awash, Ethiopia. Geological Society of America Special Papers 446, 119134.Google Scholar
Rook, L. (2008). The discovery of the Sahabi Site: Ardito Desio or Carlo Petrocchi? Garyounis Scientific Bulletin 2008, 1321.Google Scholar
Rook, L., Ghinassi, M., Libsekal, Y., et al. (2010). Stratigraphic context and taxonomic assessment of the large cercopithecoid (Primates, Mammalia) from the late Early Pleistocene palaeoanthropological site of Buia (Eritrea). Journal of Human Evolution 59, 692697.Google Scholar
Rook, L., Ghinassi, M., Carnevale, G., et al. (2013). Stratigraphic context and paleoenvironmental significance of minor taxa (Pisces, Reptilia, Aves, Rodentia) from the late Early Pleistocene paleoanthropological site of Buia (Eritrea). Journal of Human Evolution 64, 8392.Google Scholar
Rook, L., Abbate, E., Bondioli, L., et al. (2014). The Homo-bearing late Early Pleistocene site of Mulhuli-Amo, northern Danakil Depression of Eritrea. Geochronological setting, paleoenvironmental reconstruction and paleoanthropological findings. International Symposium on The African Human Fossil Record, pp. 1718 (abstract).Google Scholar
Rosenzweig, M.L. (1995). Species Diversity in Space and Time. Cambridge: Cambridge University Press.Google Scholar
Rossie, J.B. and Hill, A. (2018). A new species of Simiolus from the Middle Miocene of the Tugen Hills, Kenya. Journal of Human Evolution 125, 5058.Google Scholar
Rossignol-Strick, M. (1985). Mediterranean Quaternary sapropels, an immediate response of the African monsoon to variation of insolation. Paleogeography, Paleoclimatology, Paleoecology 49, 237263.Google Scholar
Rossouw, L. (2009). The application of fossil grass-phytolith analysis in the reconstruction of Cainozoic environments in the South African interior. Doctoral dissertation, University of the Free State.Google Scholar
Rossouw, L. (2016). An Early Pleistocene phytolith record from Wonderwerk Cave, Northern Cape, South Africa. African Archaeological Review 33, 251263.Google Scholar
Rossouw, L. and Scott, L. (2011). Phytoliths and pollen, the microscopic plant remains in Pliocene volcanic sediments around Laetoli, Tanzania. In: Harrison, T. (Ed.), Paleontology and Geology of Laetoli, Tanzania: Human Evolution in Context. Vol. 1: Geology, Geochronology, Paleoecology and Paleoenvironment. Dordrecht: Springer, pp. 201215.Google Scholar
Rossouw, L., Stynder, D.D. and Haarhof, P. (2009). Evidence for opal phytolith preservation in the Langebaanweg ‘E’ Quarry Varswater Formation and its potential for palaeohabitat reconstruction. South African Journal of Science 105(5–6), 223227.Google Scholar
Rowan, J., Faith, J.T., Gebru, Y. and Fleagle, J.G. (2015). Taxonomy and paleoecology of fossil Bovidae (Mammalia, Artiodactyla) from the Kibish Formation, southern Ethiopia: implications for dietary change, biogeography, and the structure of living bovid faunas of East Africa. Palaeogeography, Palaeoclimatology, Palaeoecology 420, 210222.Google Scholar
Rowan, J., Kamilar, J.M., Beaudrot, L. and Reed, K.E. (2016). Strong influence of palaeoclimate on the structure of modern African mammal communities. Proceedings of the Royal Society B: Biological Sciences 283(1840), 20161207.Google Scholar
Rowan, J., Locke, E.M., Robinson, J.R., et al. (2017). Fossil Giraffidae (Mammalia, Artiodactyla) from Lee Adoyta, Ledi-Geraru, and Late Pliocene dietary evolution in giraffids from the Lower Awash Valley, Ethiopia. Journal of Mammalian Evolution 24, 359371.Google Scholar
Rovinsky, D.S., Herries, A.I., Menter, C.G. and Adams, J.W. (2015). First description of in situ primate and faunal remains from the Plio-Pleistocene Drimolen Makondo palaeocave infill, Gauteng, South Africa. Palaeontologia Electronica 18(2), 121.Google Scholar
Rubenstein, D., Low Mackey, B., Davidson, Z.D., Kebede, F. and King, S.R.B. (2016). Equus grevyi. The IUCN Red List of Threatened Species 2016, e.T7950A89624491Google Scholar
Ruth, A.A., Raghanti, M.A., Meindl, R.S. and Lovejoy, C.O. (2016). Locomotor pattern fails to predict foramen magnum angle in rodents, strepsirrhine primates, and marsupials. Journal of Human Evolution 94, 4552.Google Scholar
Rüther, H., Chazan, M., Schroeder, R., et al. (2009). Laser scanning for conservation and research of African cultural heritage sites: the case study of Wonderwerk Cave, South Africa. Journal of Archaeological Science 36, 18471856.Google Scholar
Rutherford, M.C. and Westfall, R.H. (1986). Biomes of southern Africa – an objective categorization. Memoires of the Botanical Survey of South Africa 54, 198.Google Scholar
Sabatier, M. (1978). Un nouveau Tachyoryctes (Mammalia, Rodentia) du basin Pliocène de Hadar (Éthiopie). Géobios 11, 9599.Google Scholar
Sabatier, M. (1979). Les rongeurs des sites à Hominidés de Hadar et Melka-Kunture (Ethiopie). Thèse USTL, Montpellier, France.Google Scholar
Sabatier, M. (1982). Les rongeurs du site Pliocène à hominidés de Hadar (Ethiopie). Paleovertebrata 12, 156.Google Scholar
Saegusa, H. and Haile-Selassie, Y. (2009). Proboscidea. In: Haile-Selassie, Y. and WoldeGabriel, G. (Eds.), Ardipithecus kadabba: Late Miocene Evidence from the Middle Awash, Ethiopia. Berkeley: University of California Press, pp. 469516.Google Scholar
Sahle, Y., El Zaatari, S. and White, T.D. (2017). Hominid butchers and biting crocodiles in the African Plio–Pleistocene. Proceedings of the National Academy of Sciences 114(50), 1316413169.Google Scholar
Sahnouni, M. (2006). Les plus vieilles traces d’occupation humaine en Afrique du Nord: perspective de l’Ain Hanech, Algérie. Comptes Rendus Palevol 5, 243254.Google Scholar
Sahnouni, M., Hadjouis, D., van der Made, J., et al. (2002). Further research at the Oldowan site of Ain Hanech, North-Eastern Algeria. Journal of Human Evolution 43, 925937.Google Scholar
Sahnouni, M., Rosell, J., Van der Made, J., et al. (2013). The first evidence of cut marks and usewear traces from the Plio-Pleistocene locality of El-Kherba (Ain Hanech), Algeria: implications for early hominin subsistence activities circa 1.8 Ma. Journal of Human Evolution 64, 137150.Google Scholar
Sahnouni, M., Parés, J.M., Duval, M., et al. (2018). 1.9-million- and 2.4-million-year-old artifacts and stone tool–cutmarked bones from Ain Boucherit, Algeria. Science 362, 12971301. 10.1126/science.aau0008Google Scholar
Sakai, T., Saneyoshi, M., Tanaka, S., et al. (2010). Climate shift recorded at around 10 Ma in Miocene succession of Samburu Hills, northern Kenya Rift, and its significance. Geological Society, London, Special Publications 342(1), 109127.Google Scholar
Salzburger, W., Van Bocxlaer, B. and Cohen, A.S. (2014). Ecology and evolution of the African Great Lakes and their faunas. Annual Review of Ecology, Evolution, and Systematics 45, 519545.Google Scholar
Salzmann, U., Williams, M., Haywood, A.M., et al. (2011). Climate and environment of a Pliocene warm world. Palaeogeography, Palaeoclimatology, Palaeoecology 309, 18.Google Scholar
Sampson, C.G. (2003). Amphibians from the Acheulean site at Duinefontein 2 (Western Cape, South Africa). Journal of Archaeological Science 30(5), 547557.Google Scholar
Sanders, H.L. (1968). Marine benthic diversity: a comparative study. The American Naturalist 102(925), 243282.Google Scholar
Sanders, W.J. (1990). Fossil Proboscidea from the Pliocene Lusso Beds of the Western Rift, Zaire. In: Boaz, N.T. (Ed.), Evolution of Environments and Hominidae in the African Western Rift Valley. Martinsville: Virginia Museum of Natural History.Google Scholar
Sanders, W.J. (1997). Fossil Proboscidea from the Wembere–Manonga Formation, Manonga Valley, Tanzania. In: Harrison, T. (Ed.), Neogene Paleontology of the Manonga Valley, Tanzania. New York: Plenum Press, pp. 265310.Google Scholar
Sanders, W.J. (2007). Taxonomic review of fossil Proboscidea (Mammalia) from Langebaanweg, South Africa. Transactions of the Royal Society of South Africa 62, 116.Google Scholar
Sanders, W.J. (2011). Proboscidea. In: Harrison, T. (Ed.), Paleontology and Geology of Laetoli: Human Evolution in Context. Dordrecht: Springer, pp. 233262.Google Scholar
Sanders, W.J. (2020). Proboscidea from Kanapoi, Kenya. Journal of Human Evolution 140, 102547.Google Scholar
Sanders, W. and Haile-Selassie, Y. (2012). A new assemblage of mid-Pliocene proboscideans from the Woranso-Mille area, Afar Region, Ethiopia: taxonomic, evolutionary, and paleoecological considerations. Journal of Mammalian Evolution 19, 105128.Google Scholar
Sanders, W.J. and Miller, E.R. (2002). New proboscideans from the Early Miocene of Wadi Moghra, Egypt. Journal of Vertebrate Paleontology 22, 388404.Google Scholar
Sanders, W.J., Trapani, J. and Mitani, J.C. (2003). Taphonomic aspects of crowned hawk-eagle predation on monkeys. Journal of Human Evolution 44, 87105.Google Scholar
Sanders, W.J., Gheerbrant, E., Harris, J.M., Saegusa, H. and Delmer, C. (2010). Proboscidea. In: Werdelin, L. and Sanders, W.J. (Eds.), Cenozoic Mammals of Africa. Berkeley: University of California Press, pp. 161251.Google Scholar
Sandrock, O., Kullmer, O., Schrenk, F., Yuwayeyi, Y.M. and Bromage, T.G. (2007). Fauna, taphonomy and ecology of the Plio-Pleistocene Chiwondo Beds, Northern Malawi. In: Bobe, R., Alemseged, Z. and Behrensmeyer, A.K. (Eds.), Hominin Environments in the East African Pliocene: an Assessment of the Faunal Evidence. Dordrecht: Springer, pp. 315332.Google Scholar
Sands, W.A. (1987). Ichnocoenoses of probable termite origin from Laetoli. In: Leakey, M.D. and Harris, J.M. (Eds.), Laetoli: A Pliocene Site in Northern Tanzania. Oxford: Clarendon Press, pp. 409433.Google Scholar
Saneyoshi, M., Nakayama, K., Sakai, T., Sawada, Y. and Ishida, H. (2006). Half graben filling processes in the early phase of continental rifting: the Miocene Namurungule Formation of the Kenya Rift. Sedimentary Geology 186, 111131.Google Scholar
Sankaran, M., Hanan, N.P., Scholes, R.J., et al. (2005). Determinants of woody cover in African savannas. Nature 438, 846849.Google Scholar
Sarna-Wojcicki, A.M., Meyer, C.E., Roth, P.H. and Brown, F.H. (1985). Ages of tuff beds at East African early hominid sites and sediments in the Gulf of Aden. Nature 313, 306308.Google Scholar
Sarnthein, M., Thiede, J., Pflaumann, U., et al. (1982). Atmospheric and oceanic circulation patterns off Northwest Africa during the past 25 million years. In Rad, U., Hinz, K., Sarnthein, M. and Seibold, E. (Eds.), Geology of the Northwest African Continental Margin. New York: Springer Science & Business Media, pp. 545604.Google Scholar
Sausse, F. (1975). La mandibule atlanthropienne de la carrière Thomas I (Casablanca). L’Anthropologie 79, 81112.Google Scholar
Savage, R.J.G. and Hamilton, W.R. (1973). Introduction to the Miocene mammal faunas of Gebel Zelten, Libya. Bulletin of the British Museum (Natural History) Geology 22, 513527.Google Scholar
Sawada, Y., Pickford, M., Itaya, T., et al. (1998). K–Ar ages of Miocene Hominoidea (Kenyapithecus and Samburupithecus) from Samburu Hills, northern Kenya. Comptes Rendus de l’Académie des Sciences – Series IIA – Earth and Planetary Science 326, 445451.Google Scholar
Sawada, Y., Pickford, M., Senut, B., et al. (2002). The age of Orrorin tugenensis, an early hominid from the Tugen Hills, Kenya. Comptes Rendus Palevol 1, 293303.Google Scholar
Sawada, Y., Saneyoshi, M., Nakayama, K., et al. (2006). The ages and geological backgrounds of Miocene hominoids Nacholapithecus, Samburupithecus, and Orrorin from Kenya. In: Ishida, H., Tuttle, R., Pickford, M., Ogihara, N. and Nakatsukasa, M. (Eds.), Human Origins and Environmental Backgrounds. Boston, MA: Springer US, pp. 7196.Google Scholar
Saylor, B.Z., Angelini, J., Deino, A., et al. (2016). Tephrostratigraphy of the Waki-Mille area of the Woranso-Mille paleoanthropological research project, Afar, Ethiopia. Journal of Human Evolution 93, 2545.Google Scholar
Schefuß, E., Schouten, S., Jansen, J.F. and Damsté, J.S.S. (2003). African vegetation controlled by tropical sea surface temperatures in the mid-Pleistocene period. Nature 422, 418421.Google Scholar
Schick, K.D. (1987). Modeling the formation of Early Stone Age artifact concentrations. Journal of Human Evolution 16, 789807.Google Scholar
Schmid, P. (2004). Functional interpretation of the Laetoli footprints. In: Meldrum, D.J. and Hilton, C.E. (Eds.), From Biped to Strider: The Emergence of Modern Human Walking, Running, and Resource Transport. New York: Kluwer Academic, pp. 4962.Google Scholar
Schmid, P. and Berger, L.R. (1997). Middle Pleistocene hominid carpal proximal phalanx from the Gladysvale site, South Africa. South African Journal of Science 93(10), 430431.Google Scholar
Schmitt, J.-J., Wempler, J.-M., Chavaillon, J. and Andrews, M.C. (1977). Initial K/Ar and paleomagnetic results of the Melka-Kunturé early-man sites, Ethiopia. In: Proceedings 8th Panafrican Congress of Prehistory and Quaternary Studies. Nairobi: International Louis Leakey Memorial Institute for African Prehistory.Google Scholar
Schmitt, T.J. and Nairn, A.E.M. (1984). Interpretations of the magnetostratigraphy of the Hadar hominid site, Ethiopia. Nature 309, 704706.Google Scholar
Schneider, S., Hornung, J. and Hinderer, M. (2017). Evolution of the northern Albertine Rift reflected in the provenance of synrift sediments (Nkondo-Kaiso area, Uganda). Journal of African Earth Sciences 131, 183197.Google Scholar
Schoeninger, M.J., Reeser, H. and Hallin, K. (2003). Paleoenvironment of Australopithecus anamensis at Allia Bay, East Turkana, Kenya: evidence from mammalian herbivore enamel stable isotopes. Journal of Anthropological Archaeology 22(3), 200207.Google Scholar
Scholz, C.A., Cohen, A.S., Johnson, T.C., et al. (2011). Scientific drilling in the Great Rift Valley: the 2005 Lake Malawi Scientific Drilling Project – an overview of the past 145,000 years of climate variability in Southern Hemisphere East Africa. Palaeogeography, Palaeoclimatology, Palaeoecology 303(1–4), 319.Google Scholar
Schrenk, F., Bromage, T.G., Betzler, C.G., Ring, U. and Juwayeyi, Y.M. (1993). Oldest Homo and Pliocene biogeography of the Malawi Rift. Nature 365, 833836.Google Scholar
Schrenk, F., Bromage, T.G., Gorthner, A. and Sandrock, O. (1995). Paleoecology of the Malawi Rift: vertebrate and invertebrate faunal contexts of the Chiwondo Beds, northern Malawi. Journal of Human Evolution 28, 5970.Google Scholar
Schroeder, L., Scott, J.E., Garvin, H.M., et al. (2017). Skull diversity in the Homo lineage and the relative position of Homo naledi. Journal of Human Evolution 104, 124135.Google Scholar
Schubert, B.W. (2007). Dental mesowear and the palaeodiets of bovids from Makapansgat Limeworks Cave, South Africa. Palaeontologica Africana 42, 4350.Google Scholar
Schubert, B.W., Ungar, P.S., Sponheimer, M. and Reed, K.E. (2006). Microwear evidence for Plio-Pleistocene bovid diets from Makapansgat Limeworks Cave, South Africa. Palaeogeography, Palaeoclimatology, Palaeoecology 241, 301319.Google Scholar
Schulze, R.E. (1997). South African Atlas of Agrohydrology and Climatology. Report TT82/96 ACRU Report 46. Pretoria: South African Water Research Commission, pp. 276.2.Google Scholar
Schuster, M., Duringer, P., Ghienne, J.-F., et al. (2006). The age of the Sahara desert. Science, 311(5762), 821821.Google Scholar
Schwarcz, H.P., Grün, R. and Tobias, P.V. (1994). ESR dating studies of the australopithecine site of Sterkfontein, South Africa. Journal of Human Evolution 26, 175181.Google Scholar
Schwartz, J.H. and Tattersall, I. (2003). The Human Fossil Record, Vol. 2. The Craniodental Morphology of Genus Homo (African and Asia). New York: Wiley-Liss.Google Scholar
Schweitzer, F.R. and Wilson, M.L. (1982). Byneskranskop 1, a late Quaternary living site in the southern Cape Province, South Africa. Annals of the South African Museum 88, 1203.Google Scholar
Scott, J.R. (2012). Dental microwear texture analysis of Pliocene bovids from four early hominin fossil sites in eastern Africa: implications for paleoenvironmental dynamics and human evolution. PhD dissertation, University of Arkansas.Google Scholar
Scott, L. (1987). Pollen analysis of hyena coprolites and sediments from Equus Cave, Taung, Southern Kalahari (South Africa). Quaternary Research 28(1), 144156.Google Scholar
Scott, L. and Neumann, F.H. (2018). Pollen-interpreted palaeoenvironments associated with the Middle and Late Pleistocene peopling of Southern Africa. Quaternary International 495, 169184.Google Scholar
Scott, L. and Nyakale, M. (2002). Pollen indications of Holocene palaeoenvironments at Florisbad spring in the central Free State, South Africa. The Holocene 12(4), 497503.Google Scholar
Scott, L. and Rossouw, L. (2005). Reassessment of botanical evidence for palaeoenvironments at Florisbad, South Africa. South African Archaeological Bulletin 60(182), 96.Google Scholar
Scott, L. and Thackeray, J.F. (2015). Palynology of Holocene deposits in Excavation 1 at Wonderwerk Cave, Northern Cape (South Africa). African Archaeological Review 32, 839855.Google Scholar
Scott, L., Fernández-Jalvo, Y., Carrión, J. and Brink, J. (2003). Preservation and interpretation of pollen in hyaena coprolites: taphonomic observations from Spain and southern Africa. Palaeontologia Africana 39, 8391.Google Scholar
Scott, L., Neumann, F.H., Brook, G.A., et al. (2012). Terrestrial fossil-pollen evidence of climate change during the last 26 thousand years in Southern Africa. Quaternary Science Reviews 32, 100118.Google Scholar
Scott, R.S., Ungar, P.S., Bergstrom, T.S., et al. (2006). Dental microwear texture analysis: technical considerations. Journal of Human Evolution 51(4), 339349.Google Scholar
Sealy, J., Naidoo, N., Hare, V.J., Brunton, S. and Faith, J.T. (2020). Climate and ecology of the palaeo-Agulhas Plain from stable carbon and oxygen isotopes in bovid tooth enamel from Nelson Bay Cave, South Africa. Quaternary Science Reviews 235, 105974.Google Scholar
Seiffert, E.R., Perry, J.M., Simons, E.L. and Boyer, D.M. (2009). Convergent evolution of anthropoid-like adaptations in Eocene adapiform primates. Nature 461(7267), 11181121.Google Scholar
Sellers, W.I., Cain, G.M., Wang, W. and Crompton, R.H. (2005). Stride lengths, speed and energy costs in walking of Australopithecus afarensis: using evolutionary robotics to predict locomotion of early human ancestors. Journal of the Royal Society Interface 2, 431441.Google Scholar
Selvaggio, M.M. (1998). Evidence for a three-stage sequence of hominid and carnivore involvement with long bones at FLK Zinjanthropus, Olduvai Gorge, Tanzania. Journal of Archaeological Science 25, 191202.Google Scholar
Selvaggio, M.M. and Wilder, J. (2001). Identifying the involvement of multiple carnivore taxa with archaeological bone assemblages. Journal of Archaeological Science 28, 465470.Google Scholar
Semaw, S., Renne, P., Harris, J.W.K., et al. (1997). 2.5-million-year-old stone tools from Gona, Ethiopia. Nature 385, 333336.Google Scholar
Semaw, S., Schick, K., Toth, N., et al. (2001). Further 2.5–2.6 million year old artifacts, new Plio-Pleistocene archaeological sites and hominid discoveries of 1999 from Gona Ethiopia. Journal of Human Evolution 40, A20.Google Scholar
Semaw, S., Rogers, M.J., Quade, J., et al. (2003). 2.6-Million-year-old stone tools and associated bones from OGS-6 and OGS-7, Gona, Afar, Ethiopia. Journal of Human Evolution 45, 169177.Google Scholar
Semaw, S., Simpson, S.W., Quade, J., et al. (2005). Early Pliocene hominids from Gona, Ethiopia. Nature 433, 301305.Google Scholar
Semaw, S., Rogers, M.J. and Stout, D. (2009). The Oldowan–Acheulian transition: is there a “developed Oldowan” artifact tradition? In: Camps, M. and Chauhan, P. (Eds.), Sourcebook of Paleolithic Transitions. New York, Springer, pp. 173193.Google Scholar
Semaw, S., Rogers, M.J., Simpson, S.W., et al. (2020). Co-occurrence of Acheulian and Oldowan artifacts with Homo erectus cranial fossils from Gona, Afar, Ethiopia. Science Advances 6, eaaw4694.Google Scholar
Semprebon, G.M. and Rivals, F. (2007). Was grass more prevalent in the pronghorn past? An assessment of the dietary adaptations of Miocene to recent Antilocapridae (Mammalia: Artiodactyla). Palaeogeography, Palaeoclimatology, Palaeoecology 253(3–4), 332347.Google Scholar
Semprebon, G.M. and Rivals, F. (2010). Trends in the paleodietary habits of fossil camels from the Tertiary and Quaternary of North America. Palaeogeography, Palaeoclimatology, Palaeoecology 295(1–2), 131145.Google Scholar
Sen, S. (1990). Hipparion datum and its chronologic evidence in the Mediterranean area. In: Lindsay, E.H., Fahlbusch, V. and Mein, P. (Eds.), European Neogene Mammal Chronology. New York: Springer Science & Business Media, pp. 495505.Google Scholar
Sénégas, F., Thackeray, J.F., Gommery, D. and Braga, J. (2002). Palaeontological sites on ‘Bolt’s Farm’, Sterkfontein Valley, South Africa. Annals of the Transvaal Museum 39(1), 6567.Google Scholar
Sénégas, F., Paradis, E. and Michaux, J. (2005). Homogeneity of fossil assemblages extracted from mine dumps: an analysis of Plio‐Pleistocene fauna from South African caves. Lethaia 38(4), 315322.Google Scholar
Senut, B. and Gommery, D. (1997). Squelette postcrânien d’Otavipithecus, Hominoidea du Miocène moyen de Namibie. Annales de Paléontologie 83, 267284.Google Scholar
Senut, B. and Pickford, M. (Eds.) (1994). Geology and Palaeobiology of the Albertine Rift Valley, Uganda-Zaire, Vol. II, Palaeobiology. Publication Occasionelle. Orléans: Centre International pour la Formation et les Echanges Géologiques – CIFEG, p. 423.Google Scholar
Senut, B., Pickford, M., Gommery, D., et al. (2001). First hominid from the Miocene (Lukeino Formation, Kenya). Comptes Rendus de l’Academie des Sciences, Serie II: Sciences de la Terre et des Planetes 332, 137144.Google Scholar
Senut, B., Pickford, M. and Ségalen, L. (2009). Neogene desertification of Africa. Comptes Rendus Geoscience 341(8–9), 591602.Google Scholar
Şenyürek, M. (1955). A note on the teeth of Meganthropus africanus Weinert from Tanganyika Territory. Belleten 19, 155.Google Scholar
SepkoskiJr, J.J. (1988). Alpha, beta, or gamma: where does all the diversity go? Paleobiology 14, 221234.Google Scholar
Sepulchre, P., Ramstein, G., Fluteau, F., et al. (2006). Tectonic uplift and eastern African aridification. Science 313, 14191423.Google Scholar
Sewell, L. (2019). Using a multiproxy analysis of springbok fossils to track two million years of vegetation changes as experienced by our hominin ancestors. Unpublished PhD thesis. Department of Archaeology, Anthropology and Forensic Science, Bournemouth University.Google Scholar
Sewell, L., Merceron, G., Hopley, P.J., Zipfel, B. and Reynolds, S.C. (2019). Using springbok (Antidorcas) dietary proxies to reconstruct inferred palaeovegetational changes over 2 million years in Southern Africa. Journal of Archaeological Science: Reports 23, 10141028.Google Scholar
Shaar, R., Matmon, A., Horwitz, L.K., et al. (2021). Magnetostratigraphy and cosmogenic dating of Wonderwerk Cave: new constraints for the chronology of the South African Earlier Stone Age. Quaternary Science Reviews 259, 106907.Google Scholar
Shackleton, N.J. (2000). The 100,000-year ice-age cycle identified and found to lag temperature, carbon dioxide, and orbital eccentricity. Science 289, 18971902.Google Scholar
Shackleton, N.J. and Kennett, J.P. (1975). Paleotemperature history of the Cenozoic and the initiation of Antarctic glaciation: oxygen and carbon isotope analyses in DSDP Sites 277, 279, and 281. Initial Reports of the Deep Sea Drilling Project 29, 743755.Google Scholar
Shackleton, N.J., Backman, J., Zimmerman, H., et al. (1984). Oxygen isotope calibration of the onset of ice-rafting and history of glaciation in the North Atlantic region. Nature 307, 620623.Google Scholar
Shackleton, N.J., Berger, A. and Peltier, W.R. (1990). An alternative astronomical calibration of the lower Pleistocene timescale based on ODP Site 677. Earth and Environmental Science Transactions of the Royal Society of Edinburgh 81(4), 251261.Google Scholar
Shackleton, R.M. (1978). Geological map of the Olorgesailie area. In: Bishop, W.W. (Ed.), Geological Background to Fossil Man. Edinburgh: Scottish Academic Press, pp. 171172.Google Scholar
Shaffer, B.S. (1992). Quarter-inch screening: understanding biases in recovery of vertebrate faunal remains. American Antiquity 57, 129136.Google Scholar
Shaffer, B.S. and Sanchez, J.L.J. (1994). Comparison of 1/8˝- and 1/4˝-mesh recovery of controlled samples of small-to-medium sized mammals. American Antiquity 59, 525530.Google Scholar
Shaw, J.C.M. (1939). Further remains of a Sterkfontein ape. Nature 143, 117.Google Scholar
Shaw, J.C.M. (1940). Concerning some remains of a new Sterkfontein primate. Annals of the Transvaal Museum 20, 145156.Google Scholar
Sheehan, P.M., Fastovsky, D.E., Hoffman, R.G., Berghaus, C.B. and Gabriel, D.L. (1991). Sudden extinction of the dinosaurs: latest Cretaceous, Upper Great Plains, U.S.A. Science 254, 835839.Google Scholar
Shipman, P. (1986). Studies of hominid–faunal interactions at Olduvai Gorge. Journal of Human Evolution 15, 691706.Google Scholar
Shipman, P. and Harris, J. (1988). Habitat preference and paleoecology of Australopithecus boisei in Eastern Africa. In: Grine, F. (Ed.), Evolutionary History of the Robust Australopithecines. New York: De Gruyter, pp. 343381.Google Scholar
Shipman, , Bosler, W., Davis, K.L., et al. (1981). Butchering of giant geladas at an Acheulian site. Current Anthropology 22, 257268.Google Scholar
Shipman, P., Potts, R. and Pickford, M. (1983). Lainyamok: a new middle Pleistocene hominid site. Nature 306, 365368.Google Scholar
Shorrocks, B. (2007). The Biology of African Savannahs. Oxford: Oxford University Press.Google Scholar
Shultz, S. and Maslin, M.A. (2013). Early human speciation, brain expansion and dispersal influenced by African climate pulses. PLoS ONE 8, e76750.Google Scholar
Sievers, C. (2006). Seeds from the middle stone age layers at Sibudu Cave. Southern African Humanities 18(1), 203222.Google Scholar
Sievers, C. (2011). Sedges from Sibudu, South Africa: evidence for their use. In: Fahmy, A.G. et al. (Eds.), Windows on the African Past: Current Approaches to African Archaeobotany. Reports in African Archaeology 3. Frankfut am Main: Africa Magna Verlag, pp. 918.Google Scholar
Signor, P.W. and Lipps, J.H. (1982). Sampling bias, gradual extinction patterns, and catastrophes in the fossil record. In: Silver, L.T. and Schultz, P.H. (Eds.), Geological Implications of Impacts of Large Asteroids and Comets on the Earth. Boulder: Geological Society of America, vol. 190, pp. 291296.Google Scholar
Sikes, N. (1994). Early hominin habitat preferences in East Africa: paleosol carbon isotopic evidence. Journal of Human Evolution 27, 2545.Google Scholar
Sikes, N.E., Potts, R. and Behrensmeyer, A.K. (1999). Early Pleistocene habitat in Member 1 Olorgesailie based on paleosol stable isotopes. Journal of Human Evolution 37, 721746.Google Scholar
Sillen, A. and Hoering, T. (1993). Chemical characterization of burnt bones from Swartkrans. In: Brain, C.K. (Ed.), Swartkrans: A Cave’s Chronicle of Early Man. Transvaal Museum Monograph No. 8. Pretoria: Transvaal Museum, pp. 243249.Google Scholar
Sillen, A., Hall, G., Richardson, S. and Armstrong, R. (1998). 87Sr/86Sr ratios in modern and fossil food-webs of the Sterkfontein Valley: implications for early hominid habitat preference. Geochimica et Cosmochimica Acta 62(14), 24632473.Google Scholar
Simberloff, D. (1972). Properties of the rarefaction diversity measurement. The American Naturalist 106, 414418.Google Scholar
Simpson, E.H. (1949). Measurement of diversity. Nature 163, 688.Google Scholar
Simpson, G.G. (1964). Species diversity of North American recent mammals. Systematic Zoology 13, 5773.Google Scholar
Simpson, G.G. (1965). Family: Galagidae. In: Leakey, L.S.B. (Ed.), Olduvai Gorge 1951–61: Volume 1. A Preliminary Report on the Geology and Fauna. Cambridge: Cambridge University Press, pp. 1516.Google Scholar
Simpson, S.W., Quade, J., Levin, N.E., et al. (2008). A female Homo erectus pelvis from Gona, Ethiopia. Science 322, 10891092.Google Scholar
Simpson, S.W., Kleinsasser, L., Quade, J., et al. (2015). Late Miocene hominin teeth from the Gona Paleoanthropological Research Project area, Afar, Ethiopia. Journal of Human Evolution 81, 6882.Google Scholar
Simpson, S.W., Levin, N.E., Quade, J., Rogers, M.J. and Semaw, S. (2019). Ardipithecus ramidus postcrania from the Gona Project area, Afar Regional State, Ethiopia. Journal of Human Evolution 129, 145.Google Scholar
Sinclair, A.R.E. (1979a). The eruption of the ruminants. In: Sinclair, A.R.E. and Norton-Griffiths, M. (Eds.), Serengeti, Dynamics of an Ecosystem. Chicago: University of Chicago Press, pp. 82103.Google Scholar
Sinclair, A.R.E. (1979b). Serengeti, dynamics of an ecosystem. In: Sinclair, A.R.E. and Norton-Griffiths, M. (Eds.), Serengeti, Dynamics of an Ecosystem. Chicago: University of Chicago Press, pp., 82103.Google Scholar
Sinclair, A.R.E., Mduma, S.A.R. and Arcese, P. (2000). What determines phenology and synchrony of ungulate breeding in the Serengeti? Ecology 81(8), 21002111.Google Scholar
Singer, B.S. (2014). A Quaternary geomagnetic instability time scale. Quaternary Geochronology 21(c), 2952.Google Scholar
Singer, R. and Wymer, J. (1968). Archaeological investigations at the Saldanha skull site in South Africa. The South African Archaeological Bulletin 23(91), 6374.Google Scholar
Singer, R. and Wymer, J. (1982). The Middle Stone Age at Klasies River Mouth in South Africa. Chicago: University of Chicago Press.Google Scholar
Skinner, J.D. and Amithers, R.H.N. (1990). The Mammals of the Southern African Subregion (2nd ed.). Pretoria: University of Pretoria.Google Scholar
Skinner, M.M., Kivell, T.L., Potze, S. and Hublin, J.-J. (2013). Microtomographic archive of fossil hominin specimens from Kromdraai B, South Africa. Journal of Human Evolution 64, 434447.Google Scholar
Smit, H.A., Robinson, T.J., Watson, J.E.M. and Jansen van Vuren, B. (2008). A new species of elephant-shrew (Afrotheria: Macroscelidea: Elephantulus) from South Africa. Journal of Mammalogy 89, 12571269.Google Scholar
Smith, A.B. (1994). Systematics and the Fossil Record. Oxford: Blackwell Scientific.Google Scholar
Smith, C.C., Morgan, M.E. and Pilbeam, D. (2010). Isotopic ecology and dietary profiles of Liberian chimpanzees. Journal of Human Evolution 58, 4355.Google Scholar
Smith, G.M., Ruebens, K., Gaudzinski-Windheuser, S. and Steele, T.E. (2019). Subsistence strategies throughout the African Middle Pleistocene: faunal evidence for behavioral change and continuity across the Earlier to Middle Stone Age transition. Journal of Human Evolution 127, 120.Google Scholar
Smith, P., Nshimirimana, R., de Beer, F.C., et al. (2012). Canteen Kopje: a new look at an old skull. South African Journal of Science 108(1/2), 19.Google Scholar
Smith, R.J. (2016). Explanations for adaptations, just-so stories, and limitations on evidence in evolutionary biology. Evolutionary Anthropology 25(6), 276287.Google Scholar
Smith, S.M., Sprain, C.J., Clemens, W.A., et al. (2018). Early mammalian recovery after the end-Cretaceous mass extinction: a high-resolution view from McGuire Creek area, Montana, USA. GSA Bulletin 130, 20002014.Google Scholar
Smithers, R.H.N. (1983). The Mammals of the Southern African Subregion. Pretoria: University of Pretoria.Google Scholar
Soares, P., Alshamali, F., Pereira, J.B., et al. (2012). The expansion of mtDNA haplogroup L3 within and out of Africa. Molecular Biology and Evolution 29, 915927.Google Scholar
Soligo, C. and Andrews, P. (2005). Taphonomic bias, taxonomic bias and historical non-equivalence of faunal structure in early hominin localities. Journal of Human Evolution 49, 206229.Google Scholar
Solounias, N. and Semprebon, G.M. (2002). Advances in the reconstruction of ungulate ecomorphology with application to early fossil equids. American Museum Novitates 3366, 149.Google Scholar
Solounias, N., McGraw, W.S., Hayek, L.A.C. and Werdelin, L. (2000). The paleodiet of the Giraffidae. In Vrba, E.S. and Schaller, G.B. (Eds.), Antelopes, Deer, and Relatives. New Haven: Yale University Press, pp. 8495.Google Scholar
Souron, A. (2012). Histoire évolutive du genre Kolpochoerus (Cetartiodactyla: Suidae) au Plio-Pléistocène en Afrique Orientale. Thesis, University of Poitiers, France.Google Scholar
Souron, A. (2017). Diet and ecology of extant and fossil wild pigs. In: Melletti, M. and Meijaard, E. (Eds.), Ecology, Conservation and Management of Wild Pigs and Peccaries. Cambridge: Cambridge University Press, pp. 2938.Google Scholar
Souron, A., Boisserie, J.R. and White, T.D. (2013). A new species of the suid genus Kolpochoerus from Ethiopia. Acta Palaeontologica Polonica 60(1), 7996.Google Scholar
Souron, A., Boisserie, J.-R. and White, T.D. (2015). A new species of Kolpochoerus from Ethiopia. Acta Paleontologica Polonica 60, 7996.Google Scholar
South African Rain Atlas, online resource http://134.76.173.220/rainfall/index.html, accessed 20/08/2013.Google Scholar
Spencer, L. (1995a). Antelopes and grasslands: reconstructing African hominid environments. PhD thesis, Anthropological Sciences. State University of New York, Stony Brook.Google Scholar
Spencer, L. (1995b). Morphological correlates of dietary resource partitioning in the African bovidae. Journal of Mammology 76, 448471.Google Scholar
Sponheimer, M. and Lee-Thorp, J.A. (1999). Isotopic evidence for the diet of an early hominid, Australopithecus africanus. Science 283, 368370.Google Scholar
Sponheimer, M. and Lee-Thorp, J.A. (2003). Using carbon isotope data of fossil bovid communities for palaeoenvironmental reconstruction. South African Journal of Science 99(5–6), 273275.Google Scholar
Sponheimer, M. and Lee-Thorp, J.A. (2009). Biogeochemical evidence for the environments of early Homo in South Africa. In: Grine, F.E.., Fleagle, J.G. and Leakey, R.E. (Eds.), The First Humans: Origin and Early Evolution of the Genus Homo. Vertebrate Paleobiology and Paleoanthropology. New York: Springer Science, pp. 185194.Google Scholar
Sponheimer, M., Reed, K.E. and Lee-Thorp, J. (1999). Combining isotopic and ecomorphological data to refine bovid paleodietary reconstruction: a case study from the Makapansgat Limeworks hominin locality. Journal of Human Evolution 36, 705718.Google Scholar
Sponheimer, M., Reed, K. and Lee-Thorp, J.A. (2001). Isotopic palaeoecology of Makapansgat Limeworks perissodactyla. South African Journal of Science 97(7–8), 327329.Google Scholar
Sponheimer, M., Lee-Thorp, J.A., de Ruiter, D.J., et al. (2005a). Hominins, sedges, and termites: new carbon isotope data from the Sterkfontein Valley and Kruger National Park. Journal of Human Evolution 418, 301312.Google Scholar
Sponheimer, M., de Ruiter, D.J., Lee-Thorp, J. and Späth, A. (2005b). Sr/Ca and early hominin diets revisited: new data from modern and fossil enamel. Journal of Human Evolution 48, 147156.Google Scholar
Sponheimer, M., Loudon, J. E., Codron, D., et al. (2006a). Do “savanna” chimpanzees consume C4 resources? Journal of Human Evolution 51, 128133.Google Scholar
Sponheimer, M., Passey, B.H., de Ruiter, D.J., et al. (2006b). Isotope evidence for dietary variability in the early hominin Paranthropus robustus. Science 314, 980982.Google Scholar
Sponheimer, M., Lee-Thorp, J.A. and de Ruiter, D.J. (2007). Icarus, isotopes, and australopith diets. In: Ungar, P.S. (Ed.), Evolution of the Human Diet: The Known, The Unknown, and The Unknowable. Oxford: Oxford University Press, pp. 132149.Google Scholar
Sponheimer, M., Alemseged, Z., Cerling, T. E., et al. (2013). Isotopic evidence of early hominin diets. Proceedings of the National Academy of Sciences of the United States of America 110, 1051310518.Google Scholar
Spoor, F., Leakey, M.G., Gathogo, P.N., et al. (2007). Implications of new early Homo fossils from Ileret, east of Lake Turkana, Kenya. Nature 448, 688691.Google Scholar
Spoor, F., Leakey, M.G. and Leakey, L.N. (2010). Hominin diversity in the Middle Pliocene of eastern Africa: the maxilla of KNM-WT 40000. Philosophical Transactions of the Royal Society B: Biological Sciences 365, 33773388.Google Scholar
Stammers, R.C., Caruana, M.V. and Herries, A.I. (2018). The first bone tools from Kromdraai and stone tools from Drimolen, and the place of bone tools in the South African Earlier Stone Age. Quaternary International 495, 87101.Google Scholar
Stanistreet, I.G. (2012). Fine resolution of early hominin time, Beds I and II, Olduvai Gorge, Tanzania. Journal of Human Evolution 63, 300308.Google Scholar
Stanistreet, I.G., Stollhofen, H., Njau, J.K., et al. (2018). Lahar inundated, modified and preserved 1.88 Ma early hominin (OH24 and OH56) Olduvai DK site. Journal of Human Evolution 116, 2742.Google Scholar
Stankiewicz, J. and de Wit, M.J. (2006). A proposed drainage evolution model for Central Africa – did the Congo flow east? Journal of African Earth Sciences 44, 7584.Google Scholar
Stauffer, R.L., Walker, A., Ryder, O., Lyons-Weiler, M. and Hedges, S.B. (2001). Human and ape molecular clocks and constraints on paleontological hypotheses. Journal of Heredity 92(6), 469474.Google Scholar
Steele, T.E. (2012). Late Pleistocene human subsistence in Northern Africa: the state of our knowledge and placement in a continental context. In: Hublin, J.-J. and McPherron, S.P. (Eds.), Modern Origins: A North African Perspective. Dordrecht: Springer, pp. 107125.Google Scholar
Steele, T.E. and Klein, R.G. (2013). The Middle and later Stone Age faunal remains from Diepkloof Rock Shelter, Western Cape, South Africa. Journal of Archaeological Science 40, 34533462.Google Scholar
Stearns, C.E. (1978). Pliocene–Pleistocene emergence of the Moroccan Meseta. Geological Society of America Bulletin 89, 16301644.Google Scholar
Stein, R. and Sarnthein, M. (1984). Late Neogene oxygen isotope stratigraphy and terrigenous flux rates at Site 544B off Morocco. Initial Reports DSDP 79, 385394.Google Scholar
Steininger, C. (2011). The dietary behaviour of early Pleistocene bovids from Cooper’s Cave and Swartkrans, South Africa. Unpublished PhD thesis, University of the Witwatersrand, Johannesburg.Google Scholar
Steininger, C., Berger, L.R. and Kuhn, B.F. (2008). A partial skull of Paranthropus robustus from Cooper’s Cave, South Africa. South African Journal of Science 104, 14.Google Scholar
Steininger, F.F., Rabeder, G. and Rögl, F. (1985). Land mammal distribution in the Mediterranean Neogene: a consequence of geokinematic and climatic events. In: Stanley, D.J. and Wezel, F.C. (Eds.), Geological Evolution of the Mediterranean Basin. New York: Springer, pp. 559571).Google Scholar
Steiper, M. and Young, N. (2006). Primate molecular divergence dates. Molecular Phylogenetics and Evolution 41(2), 384394.Google Scholar
Stern, J.T. (2000). Climbing to the top: a personal memoir of Australopithecus afarensis. Evolutionary Anthropology 9, 113133.Google Scholar
Stern, J.T. and Susman, R.L. (1983). The locomotor anatomy of Australopithecus afarensis. American Journal of Physical Anthropology 60, 279317.Google Scholar
Stern, N. (1994). The implications of time-averaging for reconstructing the land-use patterns of early tool-using hominids. Journal of Human Evolution 27(1–3), 89105.Google Scholar
Stern, N., Porch, N. and McDougall, I. (2002). FxJj43: a window into a 1.5 million-year-old palaeolandscape in the Okote Member of the Koobi Fora Formation, Northern Kenya. Geoarchaeology 17, 349392.Google Scholar
Stewart, B.A. and Mitchell, P.J. (2018). Late Quaternary palaeoclimates and human-environment dynamics of the Maloti–Drakensberg region, southern Africa. Quaternary Science Reviews 196, 120.Google Scholar
Stewart, K.M. (1990). Fossil fish from the upper Semliki. In: Boaz, N.T. (Ed.), Evolution of Environments and Hominidae in the African Western Rift Valley. Martinsville: Virginia Museum of Natural History, pp. 141163.Google Scholar
Stewart, K.M. (1997). Fossil fish from the Manonga Valley, Tanzania: description, paleoecology, and biogeographic relationships. In: Harrison, T. (Ed.), Neogene Paleontology of the Manonga Valley, Tanzania: A Window into the Evolutionary History of East Africa. New York: Springer, pp. 333349.Google Scholar
Stewart, K.M. (2001). The freshwater fish of Neogene Africa (Miocene–Pleistocene): systematics and biogeography. Fish and Fisheries 2, 177230.Google Scholar
Stewart, K. (2003). Fossil fish remains from Mio-Pliocene deposits at Lothagam, Kenya. In: Leakey, M.G. and Harris, J.M. (Eds.), Lothagam: The Dawn of Humanity in Eastern Africa. New York: Columbia University Press, pp. 75111.Google Scholar
Stewart, K.M. and Murray, A.M. (2013). Earliest fish remains from the Lake Malawi Basin, Malawi, and biogeographical implications. Journal of Vertebrate Paleontology 33, 532539.Google Scholar
Stewart, K.M. and Rufolo, S.J. (2020). Kanapoi revisited: paleoecological and biogeographical inferences from the fossil fish. Journal of Human Evolution 140, 102452.Google Scholar
Stidham, T.A. (2007). Preliminary assessment of the late Miocene avifauna from Lemudong’o, Kenya. Kirtlandia 56, 173176.Google Scholar
Stidham, T.A. (2009). A lovebird (Psittaciformes: Agapornis) from the Plio-Pleistocene Kromdraai B locality, South Africa. South African Journal of Science 105(3–4), 155157.Google Scholar
Stidham, T.A. (2010). A small Pleistocene lovebird (Psittacidae: Agapornis) from Plovers Lake, South Africa. Neues Jahrbuch für Geologie und Paläontologie-Abhandlungen 256(1), 123128.Google Scholar
Stoetzel, E., Marion, L., Nespoulet, R., El Hajraoui, M.A. and Denys, C. (2011). Taphonomy and palaeoecology of the late Pleistocene to middle Holocene small mammal succession of El Harhoura 2 cave (Rabat – Témara, Morocco). Journal of Human Evolution 60, 133.Google Scholar
Stollhofen, H. and Stanistreet, I.G. (2012). Plio-Pleistocene synsedimentary fault compartments, foundation for the eastern Olduvai Basin palaeoenvironmental mosaic, Tanzania. Journal of Human Evolution 63, 309327.Google Scholar
Stout, D., Quade, J., Semaw, S., Rogers, M.J. and Levin, N.E. (2005). Raw material selectivity of the earliest stone toolmakers at Gona, Afar, Ethiopia. Journal of Human Evolution 48, 365380.Google Scholar
Stout, D., Semaw, S., Rogers, M.J. and Cauche, D. (2010). Technological variation in the earliest Oldowan from Gona, Afar, Ethiopia. Journal of Human Evolution 58, 474491.Google Scholar
Stratford, D.J. (2011). The underground central deposits of the Sterkfontein caves, South Africa. Unpublished PhD thesis, University of the Witwatersrand, Johannesburg.Google Scholar
Stratford, D. (2015). The sterkfontein caves: geomorphology and hominin-bearing deposits. In: Grab, S. and Knight, J. (Eds.), Landscapes and Landforms of South Africa. World Geomorphological Landscapes. Cham: Springer International Publishing, pp. 147153.Google Scholar
Stratford, D. (2017). A review of the geomorphological context and stratigraphy of the Sterkfontein caves, South Africa. In: Klimchouk, A.N., Palmer, A., De Waele, J.S., Auler, A. and Audra, P. (Eds.), Hypogene Karst Regions and Caves of the World Cave and Karst Systems of the World. Cham: Springer, pp. 879891.Google Scholar
Stratford, D.J. (2018). The Sterkfontein Caves after eighty years of paleoanthropological research: the journey continues. American Anthropologist 120(1), 3954.Google Scholar
Stratford, D.J., Bruxelles, L., Clarke, R.J. and Kuman, K. (2012). New stratigraphic interpretations of the fossil and artefact-bearing deposits of the Name Chamber, Sterkfontein. The South African Archaeological Bulletin 67(196), 159167.Google Scholar
Stratford, D., Grab, S. and Pickering, T.R. (2014). The stratigraphy and formation history of fossil- and artefact-bearing sediments in the Milner Hall, Sterkfontein Cave, South Africa: new interpretations and implications for palaeoanthropology and archaeology. Journal of African Earth Sciences 96, 155167.Google Scholar
Stratford, D., Bruxelles, L., Thackeray, J.F., Pickering, T.R. and Verheyden, S. (2020). Comments on ‘U–Pb dated flowstones restrict South African early hominin record to dry climate phases’ (Pickering et al. Nature 2018; 565: 226–229). South African Journal of Science 116(3–4), 12.Google Scholar
Strauss, D. and Sadler, P.M. (1989). Classical confidence intervals and Bayesian probability estimates for ends of local taxon ranges. Mathematical Geology 21(4), 411427.Google Scholar
Street-Perrot, F.A. and Perrott, R. (1993). A Holocene vegetation, lake levels and climate of Africa. In: Wright, H.E. Jr., Kutz bach, J.E., Webb, T. III, et al. (Eds.), Global Climates since the Last Glacial Maximum. Minneapolis: University of Minnesota Press, pp. 318356.Google Scholar
Stringer, C. (2012). The status of Homo heidelbergensis (Schoetensack, 1908). Evolutionary Anthropology 21, 101107.Google Scholar
Stromer, E. (1902). Wirbeltierreste aus dem mittlerem Pliozän des Natrontales und einige subfossile und rezenten Säugetierestes aus Agypten. Zeitschrift der Deutschen Geologischen Gesellschaft 54, 108115.Google Scholar
Struhsaker, T.T. (1967). Auditory communication among vervet monkeys (Cercopithecus aethiops). In: Altmann, S.A. (Ed.), Social Communication Among Primates. Chicago: University of Chicago Press, pp. 281324.Google Scholar
Stuut, J.-B.W., Crosta, X., van der Borg, K. and Schneider, R.R. (2004). On the relationship between Antactic sea ice and southwestern African climate during the late Quaternary. Geology 32, 909912.Google Scholar
Stynder, D.D. (1997). The use of faunal evidence to reconstruct site history and Hoedjiespunt 1 (HDP1), Western Cape. Doctoral dissertation, University of Cape Town.Google Scholar
Stynder, D.D. (2009). The diets of ungulates from the hominid fossil-bearing site of Elandsfontein, Western Cape, South Africa. Quaternary Research 71(1), 6270.Google Scholar
Stynder, D.D. (2011). Fossil bovid diets indicate a scarcity of grass in the Langebaanweg E Quarry (South Africa) late Miocene/early Pliocene environment. Paleobiology 37(1), 126139.Google Scholar
Su, D. (2005). The Paleoecology of Laetoli, Tanzania: evidence from the mammalian fauna of the Upper Laetolil Beds. PhD dissertation, New York University.Google Scholar
Su, D.F. (2011). Large mammal evidence for the paleoenvironment of the Upper Laetolil and Upper Ndolanya Beds of Laetoli, Tanzania. In: Harrison, T. (Ed.), Paleontology and Geology of Laetoli: Human Evolution in Context. Vol. 1: Geology, Geochronology, Paleoecology, and Paleoenvironment. Dordrecht: Springer, pp. 381392.Google Scholar
Su, D.E. and Harrison, T. (2007). The paleoecology of the Upper Laetolil Beds at Laetoli – A reconsideration of the large mammal evidence. In: Bobe, R., Alemseged, Z. and Behrensmeyer, A.K. (Eds.), Hominin Environments in the East African Pliocene: An Assessment of the Faunal Evidence. Berlin: Springer, pp. 279313.Google Scholar
Su, D. and Harrison, T. (2008). Ecological implications of the relative rarity of fossil hominins at Laetoli. Journal of Human Evolution 55, 672681.Google Scholar
Su, D.F. and Harrison, T. (2015). The paleoecology of the Upper Laetolil Beds, Laetoli, Tanzania. A review and synthesis. Journal of African Earth Sciences 101, 405419.Google Scholar
Su, D.F., Ambrose, S.H., DeGusta, D. and Haile-Selassie, Y. (2009). Paleoenvironment. In Haile-Selassie, Y. and WoldeGabriel, G. (Eds.), Ardipithecus kadabba: Late Miocene Evidence from the Middle Awash. Berkeley: University of California Press, pp. 521547.Google Scholar
Suc, J.-P., Popescu, S.-M., Fauquette, S., et al. (2018). Reconstruction of Mediterranean flora, vegetation and climate for the last 23 million years based on an extensive pollen dataset. Ecologia Mediterranea 44, 5385.Google Scholar
Sudre, J. and Hartenberger, J.L. (1992). Oued Mya 1, nouveau gisement demammifères du Miocène supérieur dans le sud Algérien. Geobios 25(4), 553565.Google Scholar
Surovell, T.A., Finley, J.B., Smith, G.M., Brantingham, P.J. and Kelly, R. (2009). Correcting temporal frequency distributions for taphonomic bias. Journal of Archaeological Science 36, 17151724.Google Scholar
Susman, R.L. and Stern, J.T. (1991). Locomotor behavior of early hominids: epistemology and fossil evidence. In: Coppens, Y. and Senut, B. (Eds.), Origine(s) de la Bipédie chez les Hominidés. Paris: Cahiers de Paléoanthropologie, CNRS, pp. 121132.Google Scholar
Susman, R.L., Stern, J.T. and Jungers, W.L. (1984). Arboreality and bipedality in the Hadar hominids. Folia Primatologia 43, 113156.Google Scholar
Susman, R.L., Stern, J.T. and Jungers, W.L. (1985). Locomotor adaptations in the Hadar hominids. In: Delson, E. (Ed.), Ancestors: The Hard Evidence. New York: Alan R. Liss, pp. 184192.Google Scholar
Sutton, M.B., Pickering, T.R., Pickering, R., et al. (2009). Newly discovered fossil- and artifact-bearing deposits, uranium-series ages, and Plio-Pleistocene hominids at Swartkrans Cave, South Africa. Journal of Human Evolution 57, 688696.Google Scholar
Sutton, M.B., Kuman, K. and Steininger, C. (2017). Early Pleistocene stone artefacts from Cooper’s Cave, South Africa. The South African Archaeological Bulletin 72(206), 156161.Google Scholar
Suwa, G. (1990). A comparative analysis of hominid dental remains from the Shungura and Usno Formations, Omo Valley, Ethiopia. Unpublished PhD dissertation, University of California, Berkeley.Google Scholar
Suwa, G. and Ambrose, S.H. (2014). Reply to Cerling et al. Current Anthropology 55, 473474.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
Suwa, G., Asfaw, B., Beyene, Y., et al. (1997). The first skull of Australopithecus boisei. Nature 389, 489492.Google Scholar
Suwa, G., Nakaya, H., Asfaw, B., et al. (2003). Plio-Pleistocene terrestrial mammal assemblage from Konso, Southern Ethiopia. Journal of Vertebrate Paleontology 23, 901916.Google Scholar
Suwa, G., Asfaw, B., Haile-Selassie, Y., et al. (2007a). Early Pleistocene Homo erectus fossils from Konso, southern Ethiopia. Anthropological Science 115, 133151.Google Scholar
Suwa, G., Kono, R.T., Katoh, S., Asfaw, B. and Beyene, Y. (2007b). A new species of great ape from the late Miocene epoch in Ethiopia. Nature 448(7156), 921924.Google Scholar
Suwa, G., Nakaya, H. and Asfaw, B. (2014a). The Konso Formation paleontological assemblages: collecting and documentation methodologies. In: Suwa, G., Beyene, Y. and Asfaw, B. (Eds.), The Konso-Gardula Research Project Volume 1. Paleonotological Collections: Background and Fossil Aves, Cercopithecidae, and Suidae. Tokyo: The University Museum, The University of Tokyo, Bulletin, 47, 59.Google Scholar
Suwa, G., Souron, A. and Asfaw, B. (2014b). Fossil Suidae of the Konso Formation. In: Suwa, G., Beyene, Y. and Asfaw, B. (Eds.), The Konso-Gardula Research Project Volume 1. Paleonotological Collections: Background and Fossil Aves, Cercopithecidae, and Suidae, Tokyo: The University Museum, The University of Tokyo, Bulletin, 47, 7388.Google Scholar
Suwa, G., Beyene, Y., Nakaya, H., et al. (2015). Newly discovered cercopithecid, equid and other mammalian fossils from the Chorora Formation, Ethiopia. Anthropological Science 123, 1939.Google Scholar
Szalay, F.S. and Delson, E. (1979). Evolutionary History of the Primates. New York: Academic Press.Google Scholar
Taieb, M. (1971). Les dépôts quaternaires sédimentaires de la vallée de l’Aouache et leurs relations avec la néotectonique cassante du rift. Quaternaria 15, 351365.Google Scholar
Taieb, M. (1974). Evolution Quaternarie du Bassin de l’Awash (Rift Éthiopien et Afar). PhD dissertation, Universite Paris VI.Google Scholar
Taieb, M. and Fritz, B. (Eds.) (1987). Lake Natron: géologie, géochimie et paléontologie d’un bassin évaporatique du rift est-africain. Sciences Géologiques Bulletin 40.Google Scholar
Taieb, M. and Tiercelin, J.J. (1979). Sédimentation Pliocène et paléoenvironments de rift: Exemple de la formation à Hominidés d’Hadar (Afar, Éthiopie). Bulletin du Société Géologiques de France 21, 243253.Google Scholar
Taieb, M. and Tiercelin, J.J. (1980). La stratigraphie et paléoenvironnements sédimentaires de la formation d’Hadar, depression de l’Afar, Éthiopie. In: Leakey, R.E. and Ogot, B.A. (Eds.), Proceedings of the 8th Panafrican Congress on Prehistory and Quaternary Studies, Nairobi, 1977. Nairobi: TILLMIAP, pp. 109114.Google Scholar
Taieb, M., Coppens, Y., Johanson, D.C., Kalb, J.E. (1972). Dépôts Sédimentaires et faunes du Plio-Pléistocène de la Basse Vallée de l’Awash (Afar Central, Éthiopie). CComptes Rendu de l’Académie des Sciences Paris, Série D 275, 819822.Google Scholar
Taieb, M., Johanson, D.C., Coppens, Y., Bonnefille, R. and Kalb, J.E. (1974). Découverte d’Hominidés dans les séries Plio-pléistocènes d’Hadar (Bassin de l’Awash; Afar, Ethiopie). Comptes Rendus de l’Académie des Sciences, Série D 279, 735738.Google Scholar
Taieb, M., Johanson, D.C., Coppens, Y. and Aronson, J.L. (1976). Geological and palaeontological background of Hadar hominid site, Afar, Ethiopia. Nature 260, 289293.Google Scholar
Takemoto, H., Kawamoto, Y. and Furuichi, T. (2015). How did bonobos come to range south of the congo river? Reconsideration of the divergence of Pan paniscus from other Pan populations. Evolutionary Anthropology: Issues, News, and Reviews 24, 170184.Google Scholar
Tamrat, E., Thouveny, N., Taieb, M. and Opdyke, N.D. (1995). Revised magnetostratigraphy of the Plio-Pleistocene sedimentary sequence of the Olduvai Formation (Tanzania). Palaeogeography, Palaeoclimatology, Palaeoecology 114, 273283.Google Scholar
Tamrat, E., Thouveny, N., Taieb, M. and Brugal, J.P. (2014). Magnetostratigraphic study of the Melka Kunture archaeological site (Ethiopia) and its chronological implications. Quaternary International 343, 516.Google Scholar
Tarver, J.E., Donoghue, P.C.J. and Benton, M.J. (2011). Is evolutionary history repeatedly rewritten in light of new fossil discoveries? Proceedings of the Royal Society B: Biological Sciences 278, 599604.Google Scholar
Tassy, P. (1994). Les proboscideans fossiles (Mammalia) du Rift Occidental, Ouganda. In: Senut, B. and Pickford, M. (Eds.), Geology and Palaeobiology of the Albertine Rift Valley, Uganda-Zaire, Vol. II, Palaeobiology. Publication Occasionelle. Orléans: Centre International pour la Formation et les Echanges Géologiques – CIFEG, pp. 217257.Google Scholar
Tattersfield, P. (2011). Terrestrial Mollusca. In: Harrison, T. (Ed.), Paleontology and Geology of Laetoli: Human Evolution in Context. Vol. 2: Fossil Hominins and the Associated Fauna. Dordrecht: Springer, pp. 567587.Google Scholar
Taub, D.R. (2000). Climate and the U.S. distribution of C4 grass subfamilies and decarboxylation variants of C4 photosynthesis. American Journal of Botany 87, 12111125.Google Scholar
Tauxe, L., Deino, A., Behrensmeyer, A.K. and Potts, R. (1992). Pinning down the Brunhes/Matuyama and upper Jaramillo boundaries: a reconciliation of orbital and isotopic time scales. Earth and Planetary Science Letters 109, 561572.Google Scholar
Tchernov, E., Horwitz, L.K., Ronen, A. and Lister, A. (1994). The faunal remains from Evron Quarry in relation to other Lower Paleolithic hominid sites in the southern Levant. Quaternary Research 42(3), 328339.Google Scholar
Teaford, M.F. (1994). Dental microwear and dental function. Evolutionary Anthropology 3, 1730.Google Scholar
Teaford, M.F. and Ungar, P.S. (2000). Diet and the evolution of the earliest human ancestors. Proceedings of the National Academy of Sciences 97(25), 1350613511.Google Scholar
Teague, R.L. (2009). The ecological context of the Early Pleistocene hominin dispersal to Asia. PhD dissertation, The George Washington University.Google Scholar
Tennant, J.P., Mannion, P.D. and Upchurch, P. (2016). Sea level regulated tetrapod diversity dynamics through the Jurassic/Cretaceous interval. Nature Communications 7, 12737.Google Scholar
Texier, J.-P., Lefèvre, D. and Raynal, J.-P. (1994). Contribution pour un nouveau cadre stratigraphique des formations littorales quaternaires de la région de Casablanca. Comptes rendus de l’Académie de Sciences Paris (II) 318, 12471253.Google Scholar
Texier, J.-P., Lefèvre, D., Raynal, J.-P. and El Graoui, M. (2002). Lithostratigraphy of the littoral deposits of the last one million years in Casablanca region (Maroc). Quaternaire 13, 2341.Google Scholar
Texier, P.J., Roche, H. and Harmand, S. (2006). Kokiselei 5, Formation de Nachukui (Kenya): un témoignage de la variabilité ou de l’évolution des comportements techniques au Pléistocène ancien? BAR International Series 1522, 1122.Google Scholar
Texier, P.J., Porraz, G., Parkington, J., et al. (2010). A Howiesons Poort tradition of engraving ostrich eggshell containers dated to 60,000 years ago at Diepkloof Rock Shelter, South Africa. Proceedings of the National Academy of Sciences 107(14), 61806185.Google Scholar
Tfwala, C.M., van Rensburg, L.D., Schall, R., Mosia, S.M. and Dlamini, P. (2017). Precipitation intensity–duration–frequency curves and their uncertainties for Ghaap plateau. Climate Risk Management 16, 19.Google Scholar
Thackeray, A.I., Thackeray, J.F., Beaumont, P.B. and Vogel, J.C. (1981). Dated rock engravings from Wonderwerk Cave, South Africa. Science 214, 6467.Google Scholar
Thackeray, J.F. (1979). An analysis of faunal remains from archaeological sites in southern south west Africa (Namibia). South African Archaeological Bulletin 34, 1833.Google Scholar
Thackeray, J.F. (1983). Man, animals and extinctions: the analysis of Holocene faunal remains from Wonderwerk Cave, South Africa. PhD thesis, Yale University.Google Scholar
Thackeray, J.F. (1984). Climate change and mammalian fauna from the Holocene deposits at Wonderwerk Cave, northern Cape. In: Vogel, J.C. (Ed.), Late Cainozoic Paleoclimates of the Southern Hemisphere. Rotterdam: A.A. Balkema, pp. 371374.Google Scholar
Thackeray, J.F. (2015). Analysis of faunal remains from Holocene deposits, Excavation 1, Wonderwerk Cave, South Africa. African Archaeological Review 32, 729750.Google Scholar
Thackeray, J.F. (2016). A history of research on human evolution in South Africa from 1924 to 2016. Revue de primatologie 7(7).Google Scholar
Thackeray, J.F. and Brink, J.S. (2004). Damaliscus niro horns from Wonderwerk Cave and other Pleistocene sites: morphological and chronological considerations. Palaeontologica Africana 40, 8993.Google Scholar
Thackeray, J.F. and Lee-Thorp, J.A. (1992). Isotopic analysis of equid teeth from Wonderwerk Cave, northern Cape Province, South Africa. Palaeogeography, Palaeoclimatology, Palaeoecology 99(1–2), 141150.Google Scholar
Thackeray, J.F. and Watson, V. (1994). A preliminary account of faunal remains from Plovers Lake. South African Journal of Science 90, 231232.Google Scholar
Thackeray, J.F., Van de Venter, A., Van Heerden, J., Chimimba, C.T. and Lee-Thorp, J.A. (1999). Stable carbon isotope ratios in four subspecies of modern Aethomys namaquensis: spatial variability as a possible analogue of temporal variability. In: Lee-Thorp, J.A. and Clift, H. (Eds.), The Environmental Background to Hominid Evolution in Africa. Abstracts of the XV International Union for Quaternary Research Congress: 177. Cape Town: University of Cape Town.Google Scholar
Thackeray, J.F., de Ruiter, D.J., Berger, L.R. and van der Merwe, N.J. (2001). Hominid fossils from Kromdraai: a revised list of specimens discovered since 1938. Annals of the Transvaail Museum 38, 4356.Google Scholar
Thackeray, J.F., Kirschvink, J.L. and Raub, T.D. (2002). Palaeomagnetic analyses of calcified deposits from the Plio-Pleistocene hominid site of Kromdraai, South Africa. South African Journal of Science 98, 537540.Google Scholar
Thackeray, J.F., McBride, V.A., Segonyane, S.P. and Franklyn, C.B. (2003). Trace element analysis of breccias associated with the type specimen of Australopithecus (Paranthropus) robustus from Kromdraai. Annals of the Transvaal Museum 40, 147150.Google Scholar
Thackeray, J.F., Braga, J., Sénégas, F., et al. (2005). Discovery of a humerus shaft from Kromdraai B: part of the skeleton of the type specimen of Paranthropus robustus (Broom, 1938)? Annals of the Transvaal Museum 42, 9293.Google Scholar
Thackeray, J.F., Gommery, D., Sénégas, F., et al. (2008). A survey of past and present work on Plio-Pleistocene deposits on Bolt’s Farm, Cradle of Humankind, South Africa. Annals of the Transvaal Museum 45(1), 8389.Google Scholar
Thomas, D.S.G. and Shaw, P.A. (1991). The Kalahari Environment. Cambridge: Cambridge University Press.Google Scholar
Thomas, H. (1981a). Le fauna de la Grotte à Néandertaliens du Jebel Irhoud (Maroc). Quaternaria 23, 191217.Google Scholar
Thomas, H. (1981b). Les Bovidés miocènes de la formation de Ngorora du bassin de Baringo (Rift valley, Kenya). Proceedings of the Koninklijke Nederlandse Akademie van Wetenschappen 84, 335410.Google Scholar
Thomas, H. and Petter, G. (1986). Révision de la faune de mammifères du Miocène supérieurde Menacer (ex-Marceau), Algérie: Discussion sur l’âge du gisement. Geobios 19(3), 357373.Google Scholar
Thomas, P. (1884). Recherches stratigraphiques et paléontologiques sur quelques formations d’eau douce de l’Algérie. Mémoires de la Société Géologique de France, series 3, 3, 151.Google Scholar
Thompson, J.C. (2010). Taphonomic analysis of the middle stone age faunal assemblage from Pinnacle Point Cave 13B, Western Cape, South Africa. Journal of Human Evolution 59(3–4), 321339.Google Scholar
Thompson, J.C. and Henshilwood, C.S. (2011). Taphonomic analysis of the Middle Stone Age larger mammal faunal assemblage from Blombos Cave, southern Cape, South Africa. Journal of Human Evolution 60(6), 746767.Google Scholar
Thompson, J.C. and Henshilwood, C.S. (2014). Tortoise taphonomy and tortoise butchery patterns at Blombos Cave, South Africa. Journal of Archaeological Science 41, 214229.Google Scholar
Thompson, J.C., McPherron, S.P., Bobe, R., et al. (2015). Taphonomy of fossils from the hominin-bearing deposits at Dikika, Ethiopia. Journal of Human Evolution 86, 112135.Google Scholar
Thompson, J.C., Kimbel, W.H., Hovers, E. and Marean, C.W. (2016). New approaches to taphonomy and field survey of fossisl across the Hadar paleo-landscape at 3.3 Ma. PaleoAnthropology PAS 2016 abstracts, A31.Google Scholar
Thouveny, N. and Taieb, M. (1986). Preliminary magnetostratigraphic record of Pleistocene deposits, Lake Natron Basin, Tanzania. In: Frostick, L.E., Renaut, R.W., Reid, I. and Tiercelin, J.-J. (Eds.), Sedimentation in the African Rifts. Geological Society Special Publication 25. Washington, DC: Geological Society, pp. 331336.Google Scholar
Thouveny, N. and Taieb, M. (1987). Étude paléomagnétique des formations du Plio-Pléistocène de la région de la Peninj (Ouest du Lac Natron, Tanzanie). Limites de l´interprétation magnétostratigraphique. Strasbourg: Sciences et Géologie Bulletin 40, 5770.Google Scholar
Tiedemann, R., Sarnthein, M. and Shackleton, N.J. (1994). Astronomic timescale for the Pliocene Atlantic delta18O and dust flux records of Ocean Drilling Program site 659. Paleoceanography 9, 619638.Google Scholar
Tiercelin, J.J. (1986). The Pliocene Hadar Formation, Afar depression of Ethiopia. In: Frostick, L.E., Renaut, R.W., Reid, I. and Tiercelin, J.J. (Eds.), Sedimentation in the African Rifts. Oxford: Blackwell Scientific, pp. 221240.Google Scholar
Tiercelin, J.-J., Schuster, M., Roche, H., et al. (2010). New considerations on the stratigraphy and environmental context of the oldest (2.34 Ma) Lokalalei archaeological site complex of the Nachukui Formation, West Turkana, northern Kenya Rift. Journal of African Earth Sciences 58(2), 157184.Google Scholar
Timmermann, A. and Friedrich, T. (2016). Late Pleistocene climate drivers of early human migration. Nature 538, 9295.Google Scholar
Tipper, J.C. (1979). Rarefaction and rarefiction – the use and abuse of a method in paleoecology. Paleobiology 5(4), 423434.Google Scholar
Tipple, B.J. (2013). Capturing climate variability during our ancestors’ earliest days. Proceedings of the National Academy of Sciences 110(4), 11441145.Google Scholar
Tobias, P.V. (1967a). Olduvai Gorge. Volume 2, The cranium of Australopithecus (Zinjanthropus) boisei. Cambridge: Cambridge University Press.Google Scholar
Tobias, P.V. (1967b). Pleistocene deposits and new fossil localities in Kenya. Nature 215(5100), 476.Google Scholar
Tobias, P.V. (1971). Human skeletal remains from the Cave of Hearths, Makapansgat, Northern Transvaal. American Journal of Physical Anthropology 34, 335368.Google Scholar
Tobias, P.V. (1979). Men, minds and hands: cultural awakenings over two million years of humanity. The South African Archaeological Bulletin 34(130), 8592.Google Scholar
Tobias, P.V. (Ed). (1985). The former Taung cave system in light of contemporary reports and its bearing on the skull’s provenance: early deterrents to the acceptance of Australopithecus. In: Hominid Evolution: Past, Present and Future Proceedings of the Taung Diamond Jubilee, Johannesburg and Mmabatho, South Africa, pp. 2540.Google Scholar
Tobias, P.V. (1988). Numerous apparently synapomorphic features in Australopithecus robustus, Australopithecus boisei and Homo habilis: support for the Skelton–McHenry–Drawhorn hypothesis. In: Grine, FE (Ed.), Evolutionary History of the “Robust” Australopithecines. pp. 293308. Adline de Gruyter: New York.Google Scholar
Tobias, P.V. (1991). Olduvai Gorge. Volume 4, The Skulls, Endocasts and Teeth of Homo habilis. Cambridge: Cambridge University Press.Google Scholar
Tobias, P.V. (2000). The fossil hominids. In: Partridge, T.C., Maud, R.R. (Eds.), The Cenozoic of Southern Africa. Oxford Monograph on Geology and Geophysics. Oxford: Oxford University Press, pp. 252276.Google Scholar
Tobias, P.V. and Hughes, A.R. (1969). The new Witwatersrand University excavation at Sterkfontein: progress report, some problems and first results. The South African Archaeological Bulletin 24(95/96), 158169.Google Scholar
Tobias, P.V. and von Koenigswald, G.H.R. (1964). A comparison between the Olduvai hominines and those of Java and some implications for hominid phylogeny. Nature 204, 515518.Google Scholar
Tobias, P.V., Vogel, J.C., Oschadleus, H.-D., Partridge, T.C. and McKee, J.K. (1993). New isotopic and sedimentological measurements of the Thabaseek deposits (South Africa) and the dating of the Taung hominid. Quaternary Research 40, 360367.Google Scholar
Todd, N.E. (2006). Trends in proboscidean diversity in the African Cenozoic. Journal of Mammalian Evolution 13(1), 110.Google Scholar
Todd, N.E. (2010). New phylogenetic analysis of the family Elephantidae based on cranial–dental morphology. Anatomical Record 293, 7490.Google Scholar
Toerien, M.J. (1952). The fossil hyaenids of the Makapansgat Valley. South African Journal of Science 489, 293300.Google Scholar
Toerien, M.J. (1955). A sabre-tooth cat from the Makapangat Valley. Palaeontologica Africana 3, 4346.Google Scholar
Toffolo, M.B., Brink, J.S. and Berna, F. (2015). Bone diagenesis at the Florisbad spring site, Free State Province (South Africa): implications for the taphonomy the Middle and Late Pleistocene faunal assemblages. Journal of Archaeological Science: Reports 4, 152163.Google Scholar
Toffolo, M.B., Brink, J.S. and Berna, F. (2019). Microstratigraphic reconstruction of formation processes and paleoenvironments at the Early Pleistocene Cornelia-Uitzoek hominin site, Free State Province, South Africa. Journal of Archaeological Science: Reports 25, 2539.Google Scholar
Tommasini, P. (1886). La sablière de Ternifine. Bulletin de la Société Géographique de Oran 6, 5152.Google Scholar
Tong, H. (1986). The Gerbillinae (Rodentia) from Tighennif (Pleistocene of Algeria) and their significance. Modern Geology 10, 197214.Google Scholar
Tong, H. (1989). Origine et evolution des Gerbillidae (Mammalia, Rodentia) en Afrique du Nord. Mémoires de la Société géologique de France NS 155, 1118.Google Scholar
Tooth, S., McCarthy, T.S., Brandt, D., Hancox, P.J. and Morris, R. (2002). Geological controls on the formation of alluvial meanders and floodplain wetlands: the example of the Klip River, eastern Free State, South Africa. Earth Surface Processes and Landforms 27, 797815.Google Scholar
Tooth, S., Brandt, D., Hancox, P.J. and McCarthy, T.S. (2004). Geological controls on alluvial river behaviour: a comparative study of three rivers on the South African Highveld. Journal of African Earth Sciences 38(1), 7997.Google Scholar
Toots, H. (1965). Sequence of disarticulation in mammalian skeletons. Rocky Mountain Geology 4, 3739.Google Scholar
Trapani, J., Sanders, W.J., Mitcani, J.C. and Heard, A. (2006). Precision and consistency of the taphonomic signature by crowned hawk-eagles (Stephanoaetus coronatus) in Kibale National Park, Uganda. Palaios 21, 114131.Google Scholar
Trauth, M.H., Deino, A.L., Bergner, A.G.N. and Strecker, M.R. (2003). East African climate change and orbital forcing during the last 175 kyr BP. Earth and Planetary Science Letters 206, 297313.Google Scholar
Trauth, M.H., Maslin, M.A., Deino, A. and Strecker, M.R. (2005). Late Cenozoic moisture history of East Africa. Science 309(5743), 20512053.Google Scholar
Trauth, M.H., Maslin, M.A., Deino, A.L., et al. (2007). High- and low-latitude forcing of Plio-Pleistocene East African climate and human evolution. Journal of Human Evolution 53, 475486.Google Scholar
Trauth, M.H., Larrasoaña, J.C. and Mudelsee, M. (2009). Trends, rhythms and events in Plio-Pleistocene African climate. Quaternary Science Reviews 28, 399411.Google Scholar
Trauth, M.H., Maslin, M.A., Deino, A.L., et al. (2010). Human evolution in a variable environment: the amplifier lakes of Eastern Africa. Quaternary Science Reviews 29, 29812988.Google Scholar
Treydte, A.C., Bernasconi, S.M., Kreuzer, M. and Edwards, P.J. (2006). Diet of the common warthog (Phacochoerus africanus) on former cattle grounds in a Tanzanian savanna. Journal of Mammalogy 87, 889898.Google Scholar
Tribolo, C., Mercier, N., Douville, E., et al. (2013). OSL and TL dating of the Middle Stone Age sequence at Diepkloof Rock Shelter (South Africa): a clarification. Journal of Archaeological Science 40(9), 34013411.Google Scholar
Trueman, C.N. and Martill, D.M. (2002). The long-term survival of bone: the role of bioerosion. Archaeometry 44, 371382.Google Scholar
Tryon, C.A. and Faith, J.T. (2013). Variability in the Middle Stone Age of Eastern Africa. Current Anthropology 54, S234S254.Google Scholar
Tryon, C.A. and McBrearty, S. (2002). Tephrostratigraphy and the Acheulian to Middle Stone Age transition in the Kapthurin Formation, Kenya. Journal of Human Evolution 42, 211235.Google Scholar
Tryon, C.A., Faith, J.T., Peppe, D.J., et al. (2010). The Pleistocene archaeology and environments of the Wasiriya Beds, Rusinga Island, Kenya. Journal of Human Evolution 59, 657671.Google Scholar
Tryon, C.A., Peppe, D.J., Faith, J.T., et al. (2012). Late Pleistocene artefacts and fauna from Rusinga and Mfangano islands, Lake Victoria, Kenya. Azania: Archaeological Research in Africa 47, 1438.Google Scholar
Tryon, C.A., Crevecoeur, I., Faith, J.T., et al. (2015). Late Pleistocene age and archaeological context for the hominin calvaria from GvJm-22 (Lukenya Hill, Kenya). Proceedings of the National Academy of Sciences of the USA 112, 26822687.Google Scholar
Tsubamoto, T., Yutaka, K., Nakaya, H., et al. (2015). New specimens of a primitive hippopotamus, Kenyapotamus coryndonae, from the Upper Miocene Nakali Formation, Kenya. Journal of the Geological Society of Japan 121, 153159.Google Scholar
Tsubamoto, T., Kunimatsu, Y., Sakai, T., et al. (2017). Listriodontine suid and tragulid artiodactyls (Mammalia) from the Upper Miocene Nakali Formation, Kenya. Paleontological Research 21, 347357.Google Scholar
Tsubamoto, T., Kunimatsu, Y., Sakai, T., et al. (2020). A new species of Nyanzachoerus (Mammalia, Artiodactyla, Suidae, Tetraconodontinae) from the Upper Miocene Nakali Formation, Kenya. Paleontological Research 24, 4163.Google Scholar
Tsujikawa, H. (2005a). The palaeoenvironment of Samburupithecus kiptalami based on its associated fauna. African Study Monographs 32, 5162.Google Scholar
Tsujikawa, H. (2005b). The updated Late Miocene large mammal fauna from Samburu Hills, northern Kenya. African Study Monographs 32, 150.Google Scholar
Tuomisto, H. (2012). An updated consumer’s guide to evenness and related indices. Oikos, 121(8), 12031218.Google Scholar
Turner, A. (1984). Hominids and fellow travellers: human migration into high latitudes as part of a large mammal community. In: Foley, R.A. (Ed.), Hominid Evolution and Community Ecology. London: Academic Press, pp. 193217.Google Scholar
Turner, A. (1989). Sample selection, schlepp effects and scavenging: the implications of partial recovery for interpretations of the terrestrial mammal assemblage from Klasies River Mouth. Journal of Archaeological Science 16(1), 111.Google Scholar
Turner, A. (1990). The evolution of the guild of larger terrestrial carnivores during the Plio-Pleistocene in Africa. Geobios 23(3), 349368.Google Scholar
Turner, A. and Wood, B. (1993). Taxonomic and geographic diversity in robust australopithecines and other Plio-Pleistocene larger mammals. Journal of Human Evolution 24, 147168.Google Scholar
Turner, A. and O’Regan, H. (2015). Zoogeography: primate and early hominin distribution and migration patterns. In: Henke, W. and Tattersall, I. (Eds.), Handbook of Paleoanthropology. Berlin: Springer-Verlag, pp. 623642).Google Scholar
Turner, B.R. (1980). Sedimentological characteristics of the “red muds” at Makapansgat limeworks. Palaeontologia Africana 23, 5158.Google Scholar
Tuttle, R.H. (1985). Ape footprints and Laetoli impressions: a response to the SUNY claims. In: Tobias, P.V. (Ed.), Hominid Evolution: Past, Present and Future. New York: Alan R. Liss, pp. 129133.Google Scholar
Tuttle, R.H. (1987). Kinesiological inference and evolutionary implications from Laetoli bipedal trails G-1, G-2/3 and A. In: Leakey, M.D. and Harris, J.M. (Eds.), Laetoli: A Pliocene Site in Northern Tanzania. Oxford: Clarendon Press, pp. 503523.Google Scholar
Tuttle, R.H. (1990). The pitted pattern of Laetoli feet. Natural History 90, 6065.Google Scholar
Tuttle, R.H. (1994). Up from electromyography: primate energetics and the evolution of human bipedalism. In: Corruccini, R.S. and Ciochon, R.L. (Eds.), Integrative Paths to the Past. Engelwood Cliffs: Prentice Hall, pp. 251268.Google Scholar
Tuttle, R.H. (2008). Footprint clues in hominid evolution and forensics: lessons and limitations. Ichnos 15, 158165.Google Scholar
Tuttle, R.H., Webb, D.M., Weidl, E. and Baksh, M. (1990). Further progress on the Laetoli trails. Journal of Archaeological Science 17, 347362.Google Scholar
Tuttle, R.H., Webb, D. and Tuttle, N. (1991). Laetoli footprint trails and the evolution of hominid bipedalism. In: Coppens, Y. and Senut, B. (Eds.), Origine(s) de la Bipédie chez les Hominidés. Paris: Cahiers de Paléoanthropologie, CNRS, pp. 203218.Google Scholar
Tuttle, R.H., Webb, D.M., Tuttle, N.I. and Baksh, M. (1992). Footprints and gaits of bipedal apes, bears, and barefoot people: perspective on Pliocene tracks. In: Matano, S., Tuttle, R.H., Ishida, H. and Goodman, M. (Eds.) Topics in Primatology, Vol. 3: Evolutionary Biology, Reproductive Endocrinology, and Virology. Tokyo: University of Tokyo Press, pp. 221242.Google Scholar
Twiss, P.C., Suess, E. and Smith, R.M. (1969). Morphological classification of grass phytoliths. Soil Science Society of America Journal 33(1), 109115.Google Scholar
Ullrich, H. (2001). Garusi Hominid 4 – Ein Australopithecinenzahn aus der Sammlung Kohl-Larsen. In: Schultz, M., Atzwanger, K., Bräuer, G., et al. (Eds.), Homo – Unsere Herkunft und Zukunft. Proceedings 4. Kongress der Gesellschaft für Anthropologie e. V. Göttingen: Cuvillier Verlag.Google Scholar
Ungar, P.S. (1994). Incisor microwear of Sumatran anthropoid primates. American Journal of Physical Anthropology 94, 339363.Google Scholar
Ungar, P.S. (2005). Dental evidence for the diets of fossil primates from Rudabánya, Northeastern Hungary with comments on extant primate analogs and “noncompetitive” sympatry. Palaeontologia Italia 90, 97112.Google Scholar
Ungar, P.S. (Ed.) (2007). Evolution of the Human Diet. Oxford: Oxford University Press.Google Scholar
Ungar, P. and Sponheimer, M. (2011). The diets of early hominins. Science 334, 190193.Google Scholar
Ungar, P.S., Grine, F.E. and Teaford, M.F. (2006). Diet in early Homo: a review of the evidence and a new model of adaptive versatility. Annual Review of Anthropology, 35, 209228.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(3), 163181.Google Scholar
Ungar, P.S., Scott, R.S., Scott, J.R. and Teaford, M. (2008). Dental microwear analysis: historical perspectives and new approaches. In: Irish, J.D. and Nelson, G.C. (Eds.), Technique and Application in Dental Anthropology. Cambridge: Cambridge University Press, pp. 389425.Google Scholar
Uno, K.T., Cerling, T.E., Harris, J.M., et al. (2011). Late Miocene to Pliocene carbon isotope record of differential diet change among East African herbivores. Proceedings of the National Academy of Sciences of the United States of America 108, 65096514.Google Scholar
Uno, K.T., Polissar, P.J., Kahle, E., et al. (2016a). A Pleistocene palaeovegetation record from plant wax biomarkers from the Nachukui Formation, West Turkana, Kenya. Philosophical Transactions of the Royal Society of London B: Biological Sciences 371(1698).Google Scholar
Uno, K.T., Polissar, P.J., Jackson, K.E. and deMenocal, P.B. (2016b). Neogene biomarker record of vegetation change in eastern Africa. Proceedings of the National Academy of Sciences 113(23), 63556363.Google Scholar
Uno, K.T., Rivals, F., Bibi, F., et al. (2018). Large mammal diets and paleoecology across the Oldowan–Acheulean transition at Olduvai Gorge, Tanzania from stable isotope and tooth wear analyses. Journal of Human Evolution 120, 7691.Google Scholar
Upchurch, P., Mannion, P.D., Benson, R.B.J., Butler, R.J. and Carrano, M.T. (2011). Geological and anthropogenic controls on the sampling of the terrestrial fossil record: a case study from the Dinosauria. In: McGowan, A.J. and Smith, A.B. (Eds.), Comparing the Geological and Fossil Records: Implications for Biodiversity Studies. Geological Society of London Special Publications, 358. London: Geological Society, pp. 209240.Google Scholar
Urbanek, C., Faupl, P., Hujer, W., et al. (2005). Geology, paleontology and paleoanthropology of the Mount Galili Formation in the southern Afar Depression, Ethiopia – preliminary results. Joannea Geologie und Paläontologie 6, 2943.Google Scholar
Vaks, A., Bar-Matthews, M., Ayalon, A., et al. (2007). Desert speleothems reveal climatic window for African exodus of early modern humans. Geology 35, 831834.Google Scholar
Val, A. (2019). New data on avifaunal remains associated with the Middle Stone Age layers from Diepkloof Rock Shelter, Western Cape, South Africa. Journal of Archaeological Science: Reports 26, 101880.Google Scholar
Val, A. and Stratford, D.J. (2015). The macrovertebrate fossil assemblage from the Name Chamber, Sterkfontein: taxonomy, taphonomy and implications for site formation processes. Palaeontologia Africana 50, 117.Google Scholar
Val, A., Carlson, K.J., Steininger, C., et al. (2011). 3D techniques and fossil identification: an elephant shrew hemi-mandible from the Malapa site. South African Journal of Science 107(11–12), 15.Google Scholar
Val, A., Taru, P. and Steininger, C. (2014). New taphonomic analysis of large-bodied primate assemblage from Cooper’s D, Bloubank Valley, South Africa. South African Archeology Bulletin 69, 4958.Google Scholar
Val, A., de la Peña, P. and Wadley, L. (2016). Direct evidence for human exploitation of birds in the Middle Stone Age of South Africa: the example of Sibudu Cave, KwaZulu-Natal. Journal of Human Evolution 99, 107123.Google Scholar
Vallé, F., Dupont, L.M., Leroy, S.A.G., Schefuß, E. and Wefer, G. (2014). Pliocene environmental change in West Africa and the onset of strong NE trade winds (ODP Sites 659 and 658). Palaeogeography, Palaeoclimatology, Palaeoecology 414, 403414.Google Scholar
Van Bocxlaer, B. (2020). Paleoecological insights from fossil freshwater mollusks of the Kanapoi Formation (Omo-Turkana Basin, Kenya). Journal of Human Evolution 140, 102341.Google Scholar
Van Bruggen, A.C. (1978). Land molluscs. In: Werger, M.J.A. (Ed.), Biogeography and Ecology of Southern Africa. The Hague: W. Junk, pp. 877923.Google Scholar
Van Couvering, J.A.H. (1980). Community evolution in East Africa during the late Cenozoic. In: Behrensmeyer, A.K. and Hill, A.P. (Eds.), Fossils in the Making. Chicago: University of Chicago Press, pp.272298.Google Scholar
van Dam, J.A., Abdul Aziz, H., Sierra, M.A.A., et al. (2006). Long-period astronomical forcing of mammal turnover. Nature 443, 687691.Google Scholar
Van Damme, D. and Pickford, M. (1999). The late Cenozoic Viviparidae (Mollusca, Gastropoda) of the Albertine Rift Valley (Uganda–Congo). Hydrobiologia 390, 171217.Google Scholar
Van Damme, D. and Pickford, M. (2003). The late Cenozoic Thiaridae (Mollusca, Gastropoda, Cerithioidea) of the Albertine Rift Valley (Uganda–Congo) and their bearing on the origin and evolution of the Tanganyikan thalassoid malacofauna. Hydrobiologia 498, 183.Google Scholar
Van Damme, D. and Van Bocxlaer, B. (2009). Freshwater molluscs of the Nile Basin, past and present. In Dumont, H.J. (Ed.), The Nile. Monographiae Biologicae, 89. Dordrecht: Springer, pp. 585629.Google Scholar
van der Laan, E., Hilgen, F.J., Lourens, L.J., et al. (2012). Astronomical forcing of Northwest African climate and glacial history during the late Messinian (6.5–5.5 Ma). Palaeogeography, Palaeoclimatology, Palaeoecology 313–314, 107126.Google Scholar
van der Made, J. (1998) Biometrical trends in the Tetraconodontinae, a subfamily of pigs. Earth and Environmental Science Transactions of the Royal Society of Edinburgh 89(3), 199225.Google Scholar
Van der Made, J., Morales, J. and Montoya, P. (2006). Late Miocene turnover in the Spanish mammal record in relation to palaeoclimate and the Messinian Salinity Crisis. Palaeogeography, Palaeoclimatology, Palaeoecology 238(1), 228246.Google Scholar
van der Merwe, N.J. (2013). Isotopic ecology of fossil fauna from Olduvai Gorge at ca 1.8 Ma, compared with modern fauna. South African Journal of Science 109(11–12), 114.Google Scholar
van der Merwe, N., Thackeray, F., Lee-Thorp, J.A. and Luyt, J. (2003). The carbon isotope ecology and diet of Australopithecus africanus at Sterkfontein, South Africa. Journal of Human Evolution 44, 581597.Google Scholar
van der Merwe, N.J., Masao, F.T. and Bamford, M.K. (2008). Isotopic evidence for contrasting diets of early hominins Homo habilis and Australopithecus boisei of Tanzania. South African Journal of Science 104, 153155.Google Scholar
Van der Meulen, A.J. and Daams, R. (1992). Evolution of early–middle Miocene rodent faunas in relation to long-term palaeoenvironmental changes. Palaeogeography, Palaeoclimatology, Palaeoecology 93, 227253.Google Scholar
Van Dijk, D.E. (2003). Pliocene frogs from Langebaanweg, Western Cape Province, South Africa: news & views. South African Journal of Science 99(3), 123124.Google Scholar
Van Dijk, D.E. (2006a). Langebaanweg anuran bones and associated biology. African Natural History 2, 184.Google Scholar
Van Dijk, D.E. (2006b). Anura from the Late Pleistocene at Klasies River main site, South Africa. African Natural History 2(1), 167171.Google Scholar
Van Hoepen, E.C.N. (1930). Fossiele perde van Cornelia, O.V.S. Paleontology Navorsinge van die Nasionale Museum, Bloemfontein 2, 124.Google Scholar
Van Hoepen, E.C.N. (1932a). Die stamlyn van die sebras. Paleontology Navorsinge van die Nasionale Museum, Bloemfontein 2, 2537.Google Scholar
Van Hoepen, E.C.N. (1932b). Voorlopige beskrywing van Vrystaatse soogdiere. Paleontology Navorsinge van die Nasionale Museum, Bloemfontein 2, 6365.Google Scholar
Van Hoepen, E.C.N. (1947). A preliminary description of new Pleistocene mammals of South Africa. Paleontology Navorsinge van die Nasionale Museum, Bloemfontein 2, 103106.Google Scholar
Van Neer, W. and Uerpmann, H.-P. (1989). Paleoecological significance of the Holocene faunal remains of the B.O.S. missions. In: Kuper, R. (Ed.), Forschungen zur Umweltgeschichte der Ostsahara. Köln: Heinrich-Barth-Institut, pp. 307341.Google Scholar
Van Pletzen, L. (2000). The large mammal fauna from Klasies River. Doctoral dissertation, Stellenbosch, Stellenbosch University..Google Scholar
Van Riet Lowe, C. (1938). The Makapan caves. An archaeological note. South African Journal of Science 35(12), 371381.Google Scholar
van Riet Lowe, C. (1954). The Cave of Hearths. The South African Archaeological Bulletin 9, 2529.Google Scholar
Van Valen, L. (1973). A new evolutionary law. Evolutionary Theory 1, 130.Google Scholar
Van Valkenburgh, B. (1988). Trophic diversity in past and present guilds of large predatory mammals. Paleobiology 14, 155173.Google Scholar
van Zinderen Bakker, E.M. (1976). The evolution of late Quaternary paleoclimates of southern Africa. Palaeoecology of Africa 9, 160202.Google Scholar
Van Zinderen Bakker, E.M. (1982). Pollen analytical studies of theWonderwerk Cave, South Africa. Pollen et Spores 24, 235250.Google Scholar
Van Zyl, W., Brink, J.S. and Badenhorst, S. (2016). Pleistocene Bovidae from X Cave on Bolt’s Farm in the Cradle of Humankind in South Africa. Annals of the Ditsong National Museum of Natural History 6(7), 3973.Google Scholar
Vainer, S., Erel, Y. and Matmon, A. (2018). Provenance and depositional environments of Quaternary sediments in the southern Kalahari Basin. Chemical Geology 476, 352369.Google Scholar
Varnham, G.L., Mannion, P.D. and Kammerer, C.F. (2021). Spatiotemporal variation in completeness of the early cynodont fossil record and its implications for mammalian evolutionary history. Palaeontology 64(2), 307333.Google Scholar
Vercammen, P. and Mason, D.R. (1993). The warthogs (Phacochoerus africanus and P. aethiopicus). In: Oliver, W.L.R. (Ed.), Pigs, Peccaries and Hippos: Status Survey and Action Plan. Gland: World Conservation Union/Species Survival Commission, pp. 7584.Google Scholar
Verdcourt, B. (1963). The Miocene non-marine Mollusca of Rusinga Island, Lake Victoria and other localities in Kenya. Palaeontographica 121(A), 137.Google Scholar
Verdcourt, B. (1987). Mollusca from the Laetolil and Upper Ndolanya Beds. In: Leakey, M.D. and Harris, J.M. (Eds.), Laetoli: A Pliocene Site in Northern Tanzania. Oxford: Clarendon Press, pp. 438450.Google Scholar
Verna, C., Texier, P.J., Rigaud, J.P., Poggenpoel, C. and Parkington, J. (2013). The middle stone age human remains from Diepkloof rock shelter (Western cape, South Africa). Journal of Archaeological Science 40(9), 35323541.Google Scholar
Verheyen, E., Salzburger, W., Snoeks, J. and Meyer, A. (2003). Origin of the superflock of cichlid fishes from Lake Victoria, East Africa. Science 300(5617), 325329.Google Scholar
Vermeersch, P.M. (2000). Palaeolithic Living Sites in Upper and Middle Egypt. Leuven: Leuven University Press.Google Scholar
Verniers, J. (1997). Detailed stratigraphy of the Neogene sediments at Tinde and other localities in the central Manonga Basin. In: Harrison, T. (Ed.), Neogene Paleontology of the Manonga Valley, Tanzania: A Window into the Evolutionary History of East Africa. New York: Springer, pp. 3365.Google Scholar
Verschuren, D., Sinninghe Damsté, J.S., Moernaut, J., et al. (2009). Half-precessional dynamics of monsoon rainfall near the East African equator. Nature 462, 637641.Google Scholar
Viehl, K. (2003). Untersuchungen zur Nahrungsökologie des Afrikanischen Riesenwaldschweins (Hylochoerus meinertzhageni Thomas) im Queen Elizabeth National Park, Uganda. Doctoral dissertation, Universität Hannover.Google Scholar
Vilakazi, N., Gommery, D. and Kgasi, L. (2018). First fossil record of the spitting Elapidae in the cradle of humankind, South Africa. South African Archaeological Bulletin 73(207),35.Google Scholar
Villa, P. and Mahieu, E. (1991). Breakage patterns of human long bones. Journal of Human Evolution 21, 2748.Google Scholar
Villa, P., Soriano, S., Tsanova, T., et al. (2012). Border Cave and the beginning of the Later Stone Age in South Africa. Proceedings of the National Academy of Sciences of the USA 109, 1320813213.Google Scholar
Villaseñor, A. (2017). Diversity in the middle Pliocene: a multi-proxy analysis of climate’s effect on vegetation and mammalian community structure with implications for human evolution. Thesis, George Washington University, p. 135.Google Scholar
Vignaud, P., Duringer, P., Mackaye, H.T., et al. (2002). Geology and palaeontology of the upper Miocene Toros-Menalla hominid locality, Chad. Nature 418, 152155.Google Scholar
von den Driesch, A. (2004). The Middle Stone Age fish fauna from the Klasies river main site, South Africa. Anthropozoologica 39(2), 3359.Google Scholar
Villmoare, B., Kimbel, W.H., Seyoum, C., et al. (2015). Early Homo at 2.8 Ma from Ledi-Geraru, Afar, Ethiopia. Science, 347(6228), 13521355.Google Scholar
Vincens, A. and Casanova, J. (1987). Modern background of Natron–Magadi basin (Tanzania–Kenya): physiography, climate, hydrology and vegetation. Sciences Géologique Bulletin 40, 921.Google Scholar
Viola, B., Kullmer, O., Sandrock, O., Hujer, W. and Seidler, H. (2008). An early australopithecine femur from Galili, Ethiopia. Journal of Vertebrate Paleontology 28(3), 156157.Google Scholar
Viriot, L., Vignaud, P., Boisserie, J.-R., et al. (2008). Late Miocene to Early Pliocene Chadian vertebrate faunas: palaeoenvironmental and palaeobiogeographical implications. In: Salem, M.J., El-Arnauti, A. and El Soher Sale, A. (Eds.), Geology of East Libya: Third Symposium on the Sedimentary Basins of Libya. Benghazi: Earth Science Society of Libya, pp. 303308.Google Scholar
Vogel, J.C. (1985). Further attempts at dating the Taung tufas. In: Tobias, P.V. (Ed.), Hominid Evolution: Past, Present and Future. New York: Alan R. Liss, pp. 189194.Google Scholar
Vogel, J.C. and Partridge, T.C. (1983). Preliminary radiometric ages for the Taung tufas. In Vogel, J.C. (Ed.), Late Cainozoic Palaeoclimates of the Southern Hemisphere. Balkema: Rotterdam, pp. 507514.Google Scholar
Vogel, J.C., Fuls, A. and Ellis, R. (1978). The geographical distribution of Kranz grasses in South Africa. South African Journal of Science 74, 209219.Google Scholar
Von Endt, D.W. and Ortner, D.J. (1984). Experimental effects of bone size and temperature on bone diagensis. Journal of Archaeological Science 11, 247253.Google Scholar
Voorhies, M.R. (1969). Taphonomy and population dynamics of an early Pliocene vertebrate fauna, Knox County, Nebraska. Contributions to Geology, University of Wyoming Special Paper 1, 169.Google Scholar
Vrba, E.S. (1974). Chronological and ecological implications of the fossil Bovidae at the Sterkfontein Australopithecine site. Nature 250, 1923.Google Scholar
Vrba, E. (1975). Some evidence of the chronology and paleoecology of Sterkfontein, Swartkrans and Kromdraai from the fossil Bovidae. Natur, e 254, 301304.Google Scholar
Vrba, E. (1976). The Fossil Bovidae of Sterkfontein, Swartkrans and Kromdraai. Transvaal Museum Memoir no. 21. Pretoria: Heer.Google Scholar
Vrba, E.S. (1980). The significance of bovid remains as indicators of environment and predation patterns. In: Behrensmeyer, A.K. and Hill, A.P. (Eds.), Fossils in the Making. Chicago: University of Chicago Press, pp. 247271.Google Scholar
Vrba, E.S. (1981). The Kromdraai australopithecine site revisited in 1980: recent investigations and results. Annals of the Transvaal Museum 33, 1760.Google Scholar
Vrba, E. (1982a). Progress Report: Gondolin Site, Broederstroom Permit 3/102. Pretoria: Transvaal Museum, p. 1.Google Scholar
Vrba, E.S. (1982b). Biostratigraphy and chronology, based particularly on Bovidae, of southern hominid-associated assemblages: Makapansgat, Sterkfontein, Taung, Kromdraai, Swartkrans; also Elandsfontein (Saldanha), Broken Hill (now Kabwe) and Cave of Hearths. Premier Congres Int de Paleontologie Humaine, 707e752.Google Scholar
Vrba, E. (1985a). Ecological and adaptive changes assoicated with early hominid evolution. In: Delson, E. (Ed.), Ancestors: The Hard Evidence. New York: Alan R. Liss, pp. 6371.Google Scholar
Vrba, E. (1985b). Environment and evolution: alternative causes of the temporal distribution of evolutionary events. South African Journal of Science 81, 229236.Google Scholar
Vrba, E.S. (1985c). Early hominids in southern Africa: updated observations on chronological and ecological background. In: Tobias, P.V. (Ed.), Hominid Evolution: Past, Present and Future. New York: Alan R Liss, pp. 195200.Google Scholar
Vrba, E.S. (1987a). A revision of the Bovini Bovidae and a preliminary revised checklist of Bovidae from Makapansgat. Palaeontologica Africana 26, 3346.Google Scholar
Vrba, E.S. (1987b). New species and a new genus of Hippotragini Bovidae from Makapansgat Limeworks. Palaeontologica Africana 26, 4758.Google Scholar
Vrba, E.S. (1988). Late Pliocene climatic events and hominid evolution. In: Grine, F.E. (Ed.), Evolutionary History of the “Robust” Australopiths. New York: Aldine de Gruyter, pp. 405426.Google Scholar
Vrba, E.S. (1992). Mammals as a key to evolutionary theory. Journal of Mammalogy 73, 128.Google Scholar
Vrba, E.S. (1993). Turnover-pulses, the Red Queen, and related topics. American Journal of Science 293A, 418452.Google Scholar
Vrba, E.S. (1995a). On the connections between paleoclimate and evolution. In: Vrba, E.S., Denton, G.H., Partridge, T.C. and Burckle, L.H. (Eds.), Paleoclimate and Evolution, with Emphasis on Human Origins. New Haven: Yale University Press, pp. 2445.Google Scholar
Vrba, E. (1995b). The fossil record of African antelopes (Mammalia, Bovidae) in relation to human evolution and paleoclimate. In: Vrba, E., Denton, G., Partridge, T.C. and Burckle, L. (Eds.), Paleoclimate and Evolution with Emphasis on Human Origins. New Haven: Yale University Press, pp. 385424.Google Scholar
Vrba, E.S. (1997). New fossils of Alcelaphini and Caprinae (Bovidae, Mammalia) from Awash, Ethiopia, and phylogenetic analysis of Alcelaphini. Paleontologica Africana 34, 127198.Google Scholar
Vrba, E.S. (2015). Role of environmental stimuli in hominid origins. In: Henke, W. and Tattersall, I. (Eds.), Handbook of Paleoanthropology (1837–1886). Berlin: Springer-Verlag.Google Scholar
Vrba, E. and Gatesy, J. (1994). New antelope fossils from Awash, Ethiopia, and phylogenetic analysis of Hippotragini (Bovidae, Mammalia). Palaeontologia Africana 31, 5572.Google Scholar
Vrba, E. and Panagos, D.C. (1982). New perspectives on taphonomy, palaeoecology and chronology of the Kromdraai apeman. Palaeontologia Africana 15, 1326.Google Scholar
Wadley, L. (1987). Later Stone Age Hunters and Gatherers of the Southern Transvaal: Social and Ecological Interpretation. British Archaeological Reports Vol. 25. Oxford: B.A.R. International Series.Google Scholar
Wadley, L. (1997). Rose Cottage Cave: archaeological work 1987 to 1997. South African Journal of Science 93(10), 439444.Google Scholar
Wadley, L. (2004). Vegetation changes between 61 500 and 26 000 years ago: the evidence from seeds in Sibudu Cave, KwaZulu-Natal: Sibudu Cave. South African Journal of Science 100(3), 167173.Google Scholar
Wadley, L. (2006). Partners in grime: results of multi-disciplinary archaeology at Sibudu Cave. Southern African Humanities 18(1), 315341.Google Scholar
Wadley, L. (2010). Were snares and traps used in the Middle Stone Age and does it matter? A review and case study from Sibudu, South Africa. Journal of Human Evolution 58, 179192.Google Scholar
Wadley, L. and Jacobs, Z. (2006). Sibudu Cave: background to the excavations, stratigraphy and dating. Southern African Humanities 18(1), 126.Google Scholar
Wadley, L., Hodgskiss, T. and Grant, M. (2009). Implications for complex cognition from the hafting of tools with compound adhesives in the Middle Stone Age, South Africa. Proceedings of the National Academy of Sciences of the USA 106, 95909594.Google Scholar
Walker, A. (1981). Dietary hypotheses and human evolution. Philosophical Transactions of the Royal Society, London 292, 5764.Google Scholar
Walker, A. and Leakey, R.E. (1993). The Nariokotome Homo erectus Skeleton. Cambridge, MA: Harvard University Press, p. 457.Google Scholar
Walker, A. and Teaford, M. (1988). The hunt for Proconsul. Scientific American 260, 7682.Google Scholar
Walker, A., Leakey, R.E., Harris, J.M. and Brown, F.H. (1986). 2.5-Myr Australopithecus boisei from west of Lake Turkana, Kenya. Nature 322, 517522.Google Scholar
Walker, J., Cliff, R.A. and Latham, A.G. (2006). U–Pb isotopic age of the StW 573 hominid from Sterkfontein, South Africa. Science 314, 15921594.Google Scholar
Walker, S.J., Lukich, V. and Chazan, M. (2014). Kathu Townlands: a high density Earlier Stone Age locality in the interior of South Africa. PLoS One 9(7), e103436.Google Scholar
Walsh, J. (1969). Geology of the Eldama Ravine – Kabarnet area. Report of the Geological Survey Kenya, 83.Google Scholar
Walter, R.C. (1994). The age of Lucy and the first family: single-crystal 40Ar/39Ar dating of the Denen Dora and lower Kada Hadar members of the Hadar Formation, Ethiopia. Geology 22, 610.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
Walter, R.C., Manega, P.C., Hay, R.L., Drake, R.E. and Curtis, G.H. (1991). Laser-fusion 40Ar/39Ar dating of Bed I, Olduvai Gorge, Tanzania. Nature 354, 145149.Google Scholar
Wara, M.W., Ravelo, A.C. and Delaney, M.L. (2005). Permanent El Nino-like conditions during the Pliocene Warm Period. Science 309, 758761.Google Scholar
Ward, C.V. (2002). Interpreting the posture and locomotion of Australopithecus afarensis: where do we stand? Yearbook of Physical Anthropology 45, 185215.Google Scholar
Ward, C.V. (2014). Taxonomic affinity of the Pliocene hominin fossils from Fejej, Ethiopia. Journal of Human Evolution 73, 98102.Google Scholar
Ward, S. and Hill, A. (1987). Pliocene hominid partial mandible from Tabarin, Baringo, Kenya. American Journal of Physical Anthropology 72, 2137.Google Scholar
Ward, C., Leakey, M. and Walker, A. (1999a). The new hominid species Australopithecus anamensis. Evolutionary Anthropology 7, 197205.Google Scholar
Ward, C.V., Leakey, M.G., Brown, F.H., Harris, J.M. and Walker, A. (1999b). South Turkwel: a new Pliocene hominid site in Kenya. Journal of Human Evolution 36, 6995.Google Scholar
Ward, C.V., Leakey, M.G. and Walker, A. (2001). Morphology of Australopithecus anamensis from Kanapoi and Allia Bay, Kenya. Journal of Human Evolution 41, 255368.Google Scholar
Ward, C.V., Plavcan, J.M. and Manthi, F.K. (2010). Anterior dental evolution in the Australopithecus anamensis-afarensis lineage. Philosophical Transactions of the Royal Society B 365, 33333344.Google Scholar
Ward, C.V., Plavcan, J.M. and Manthi, F.K. (2020). New fossils of Australopithecus anamensis from Kanapoi, West Turkana, Kenya (2012–2015). Journal of Human Evolution 140, 102368.Google Scholar
Watson, V. (1991). Form, function and fibres: a preliminary study of the Swartkrans fossil birds. Koedoe 34, 2329.Google Scholar
Watson, V. (1993a). Composition of the Swartkrans bone accumulations, in terms of skeletal parts and animals represented. In: Brain, C. (Ed.), Swartkrans: A Cave’s Chronicle of Early Man. Pretoria: Transvaal Museum, pp. 3574.Google Scholar
Watson, V. (1993b). Glimpses from Gondolin: a faunal analysis of a fossil site near Broederstroom, Transvaal, South Africa. Palaeontologica Africana 30, 3542.Google Scholar
Weigelt, E., Dupont, L. and Uenzelmann-Neben, G. (2008). Late Pliocene climate changes documented in seismic and palynology data at the southwest African Margin. Global and Planetary Change 63(1), 3139.Google Scholar
Weigelt, J. (1927). Rezente Wirbeltierleichen und ihre palaobio-logische Bedeutung. Leipzig: Max Weg Verlag.Google Scholar
Weigelt, J. (1989). Recent Vertebrate Carcasses and Their Paleobiological Implications (translated by Schaefer, J.). Chicago: The University of Chicago Press.Google Scholar
Weinert, H. (1950). Über die neuen Vor- und Frühmenschenfunde aus Afrika, Java, China und Frankreich. Zeitschrift für Morphologie und Anthropologie 42, 113145.Google Scholar
Weiss, R.E. and Marshall, C.R. (1999). The uncertainty in the true end point of a fossil’s stratigraphic range when stratigraphic sections are sampled discretely. Mathematical Geology 31(4), 435453.Google Scholar
Wells, L.H. (1967). Antelopes in the Pleistocene of southern Africa. In: Bishop, W.W. and Clark, J.D. (Eds.), Background to Evolution in Africa. Chicago: Chicago University Press, pp 99107.Google Scholar
Wells, L.H. (1969). Faunal subdivision of the Quaternary in southern Africa. South African Archaeological Bulletin 24, 9395.Google Scholar
Wells, L.H. and Cooke, H.S.B. (1956). Fossil Bovidae from the Limeworks quarry, Makapansgat, Potgietersrust. Palaeontologica Africana 4, 155.Google Scholar
Wells, L.H., Cooke, H.B.S. and Malan, B.D. (1942). The associated fauna and culture of the Vlakkraal thermal springs, O.F.S. Transactions of the Royal Society of South Africa 29, 203233.Google Scholar
Wendorf, F. and Schild, R. (1976). Prehistory of the Nile Valley. New York: Academic Press.Google Scholar
Wendorf, F., Schild, R. and Close, A.E. (1993). Egypt During the Last Interglacial: The Middle Plaeolithic of Bir Tarfawi and Bir Sahara East. New York: Plenum Press.Google Scholar
Wengler, L., Weisrock, A., Brochier, J.-E., et al. (2002). Enregistrement fluviatile et paléoenvironnements au Pléistocène supérieur sur la bordure méridionale atlantique de l’Anti-Atlas (Oued Assaka, SO Marocain). Quaternaire 13, 179192.Google Scholar
Werdelin, L. and Barthelme, J. (1997). Brown hyena (Parahyaena brunnea) from the Pleistocene of Kenya. Journal of Vertebrate Paleontology 17, 758761.Google Scholar
Werdelin, L. and Dehghani, R. (2011). Carnivora. In: Harrison, T. (Ed.), Paleontology and Geology of Laetoli: Human Evolution in Context. Vol. 2: Fossil Hominins and the Associated Fauna. Dordrecht: Springer, pp. 189232.Google Scholar
Werdelin, L. and Lewis, M.E. (2001). A revision of the genus Dinofelis (Mammalia, Felidae). Zoological Journal of the Linnean Society 132, 147258.Google Scholar
Werdelin, L. and Lewis, M.E. (2013a). Temporal change in functional richness and eveness in the eastern African Plio-Pleistocene carnivoran guild. PLoS ONE 8(8), 111.Google Scholar
Werdelin, L. and Lewis, M.E. (2013b). Koobi Fora Research Project Volume 7, The: Carnivora. In: Leakey, R.E. (Ed.), Koobi Fora: Researches into Geology, Paleontology, and Human Origins. San Francisco: California Academy of Sciences, p. 333.Google Scholar
Werdelin, L. and Peigné, S. (2010). Carnivora. In: Werdelin, L. and Sanders, W.J. (Eds.), Cenozoic Mammals of Africa. Berkeley: University of California Press, pp. 603657.Google Scholar
Werdelin, L. and Sanders, W.J. (2010). Cenozoic Mammals of Africa. Berkeley: University of California Press.Google Scholar
Werdelin, L. and Simpson, S.W. (2009). The last amphicyonid (Mammalia, Carnivora) in Africa. Geodiversitas 31, 775787.Google Scholar
Werdelin, L., Lewis, M. and Haile-Selassie, Y. (2014). Mid-Pliocene Carnivora from the Woranso-Mille Area, Afar Region, Ethiopia. Journal of Mammalian Evolution 21, 331347.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. In: Vrba, E.S., Denton, G.H., Partridge, T.C. and Burckle, L.H. (Eds.), Paleoclimate and Evolution with Emphasis on Human Origins. New Haven: Yale University Press, pp. 356368.Google Scholar
Wesselman, H.B., Black, M.T. and Asnake, M. (2009). Small mammals. In: Haile-Selassie, Y. and WoldeGabriel, G. (Eds.), Ardipithecus kadabba: Late Miocene Evidence from the Middle Awash, Ethiopia. Berkeley: University of California Press, pp. 105133.Google Scholar
Wessels, W., Fejfar, O., Peláez-Campomanes, P., Van der Meulen, A., and De Bruijn, H. (2003). Miocene small mammals from Jebel Zelten, Libya. Coloquios de Paleontología, Volumen Extraordinario 1, 699715.Google Scholar
Western, D. (1980). Linking the ecology of past and present mammal communities. In: Behrensmeyer, A.K. and Hill, A.P. (Eds.), Fossils in the Making: Vertebrate Taphonomy and Paleoecology. Chicago: The University of Chicago Press, pp. 4154.Google Scholar
Western, D. (1989). The ecological role of elephants in Africa. Pachyderm 12, 4245.Google Scholar
Western, D. (1997). In the Dust of Kilimanjaro. Washington, DC: Island Press.Google Scholar
Western, D. and Behrensmeyer, A.K. (2009). Bone assemblages track animal community structure over 40 years in an African savanna ecosystem. Science 324, 10611064.Google Scholar
Weston, E.M. (2003). Fossil Hippopotamidae from Lothagam. In: Leakey, M.G. and Harris, J.M. (Eds.), Lothagam: The Dawn of Humanity in Eastern Africa. New York: Columbia University Press, pp. 441483.Google Scholar
Weston, E.M. and Boisserie, J.R. (2010). Hippopotamidae. In: Werdelin, L. and Sanders, W. (Eds.), Cenozoic Mammals of Africa. Berkeley: University of California Press, pp. 853871.Google Scholar
Wheeler, P.E. (1991). The thermoregulatory advantages of hominid bipedalism in open equatorial environments: the contribution of increased convective heat loss and cutaneous evaporative cooling. Journal of Human Evolution 21(2), 107115.Google Scholar
White, E.M. and Hannus, L.A. (1983). Chemical weathering of bone in archaeological soils. American Antiquity 48, 316322.Google Scholar
White, F. (1983). The Vegetation of Africa, a Descriptive Memoir to Accompany the UNESCO/AETFAT/UNSO Vegetation Map of Africa (3 Plates, Northwestern Africa, Northeastern Africa, and Southern Africa, 1: 5,000,000). Paris: UNESCO.Google Scholar
White, S.E. (1915). The Rediscovered Country. New York: Doubleday.Google Scholar
White, T.D. (1977). New fossil hominids from Laetolil, Tanzania. American Journal of Physical Anthropology 46, 197230.Google Scholar
White, T.D. (1980a). Evolutionary implications of Pliocene hominid footprints. Science 208, 175176.Google Scholar
White, T.D. (1980b). Additional fossil hominids from Laetoli, Tanzania: 1976–1979 specimens. American Journal of Physical Anthropology 53, 487504.Google Scholar
White, T.D. (1981). Primitive hominid canine from Tanzania. Science 213, 348349.Google Scholar
White, T.D. (1984). Pliocene hominids from the Middle Awash, Ethiopia. Courier Forschungsinstitut Senckenberg 69, 5768.Google Scholar
White, T.D. (1985). The hominids of Hadar and Laetoli: An element-by-element comparison of the dental samples. In: Delson, E. (Ed.), Ancestors: The Hard Evidence. New York: Alan R. Liss, pp. 138152.Google Scholar
White, T.D. (1986). Cut marks on the Bodo cranium: a case of prehistoric defleshing. American Journal of Physical Anthropology 69(4), 503509.Google Scholar
White, T.D. (1988). The comparative biology of “Robust” Australopithecus: clues from context. In: Grine, F.E. (Ed.), The Evolutionary History of the “Robust” Australopithecines. New York: Aldine de Gruyter, pp. 449483.Google Scholar
White, T.D. (1995). African omnivores: global climatic change and Plio-Pleistocene hominids and suids. In Vrba, E.S., Denton, G.H., Partridge, T.C. and Burckle, L.H. (Eds.), Paleoclimate and Evolution, with Emphasis on Human Origins. New Haven: Yale University Press, pp. 369384.Google Scholar
White, T.D. (2002). Earliest hominids. In: Hartwig, W.C. (Ed.), The Primate Fossil Record. Cambridge: Cambridge University Press, pp. 407417.Google Scholar
White, T.D. (2004). Managing paleoanthropology’s nonrenewable resources: a view from Afar. Comptes Rendus Palevol 3, 341351.Google Scholar
White, T.D. and Harris, J.M. (1977). Suid evolution and correlation of African hominid localities. Science 198, 1321.Google Scholar
White, T.D. and Suwa, G. (1987). Hominid footprints at Laetoli: facts and interpretations. American Journal of Physical Anthropology 72, 485514.Google Scholar
White, T.D. and Suwa, G. (2004). A new species of Notochoerus (Artiodactyla, Suidae) from the Pliocene of Ethiopia. Journal of Vertebrate Paleontology 24, 474480.Google Scholar
White, T.D., Moore, R.V. and Suwa, G. (1984). Hadar biostratigraphy and hominid evolution. Journal of Vertebrate Paleontology 4, 575583.Google Scholar
White, T.D., Suwa, G., Hart, W.K., et al. (1993). New discoveries of Australopithecus at Maka in Ethiopia. Nature 366, 261265.Google Scholar
White, T.D., Suwa, G. and Asfaw, B. (1994). Australopithecus ramidus, a new species of early hominid from Aramis, Ethiopia. Nature 371, 306312.Google Scholar
White, T.D., Suwa, G. and Asfaw, B. (1996). Ardipithecus ramidus, a root species for Australopithecus. In: Facchini, F. (Ed.), The First Humans and Their Cultural Manifestations. Proceedings of the Colloquia of the XIII International Congress of Prehistoric and Protohistoric Sciences, Forli, pp. 15–23.Google Scholar
White, T.D., Suwa, G., Simpson, S. and Asfaw, B. (2000). Jaws and teeth of Australopithecus afarensis from Maka, Middle Awash, Ethiopia. American Journal of Physical Anthropology 111, 4568.Google Scholar
White, T.D., Asfaw, B., DeGusta, D., et al. (2003). Pleistocene Homo sapiens from Middle Awash, Ethiopia. Nature 423, 742747.Google Scholar
White, T.D., Asfaw, B. and Suwa, G. (2005). Pliocene hominid fossils from Gamedah, Middle Awash, Ethiopia. Transactions of the Royal Society of South Africa 60, 7983.Google Scholar
White, T.D., WoldeGabriel, G., Asfaw, B., et al. (2006). Asa Issie, Aramis and the origin of Australopithecus. Nature, 440, 883889.Google Scholar
White, T.D., Ambrose, S.H., Suwa, G., et al. (2009a). Macrovertebrate paleontology and the Pliocene habitat of Ardipithecus ramidus. Science, 326, 67.Google Scholar
White, T.D., Asfaw, B., Beyene, Y., et al. (2009b). Ardipithecus ramidus and the paleobiology of early hominids. Science, 326(5949), 6486.Google Scholar
White, T.D., Ambrose, S.H., Suwa, G. and WoldeGabriel, G. (2010). Response to comment on the paleoenvironment of Ardipithecus ramidus. Science 328, 1105.Google Scholar
Whittaker, R.H. (1960). Vegetation of the Siskiyou mountains, Oregon and California. Ecological Monographs 30(3), 279338.Google Scholar
Whittaker, R.H. (1972). Evolution and measurement of species diversity. Taxon 21(2–3), 213251.Google Scholar
Wildman, D.E., Uddin, M., Liu, G., Grossman, L.I. and Goodman, M. (2003). Implications of natural selection in shaping 99.4% nonsynonymous DNA identity between humans and chimpanzees: enlarging genus Homo. Proceedings of the National Academy of Sciences 100(12), 71817188.Google Scholar
Wilkins, J. and Chazan, M. (2012). Blade production ~500 thousand years ago at Kathu Pan 1, South Africa: support for a multiple origins hypothesis for early Middle Pleistocene blade technologies. Journal of Archaeological Science 39(6), 18831900.Google Scholar
Wilkins, J., Pollarolo, L. and Kuman, K. (2010). Prepared core reduction at the site of Kudu Koppie in northern South Africa: temporal patterns across the earlier and Middle Stone Age boundary. Journal of Archaeological Science 37, 12791292.Google Scholar
Wilkins, J., Schoville, B.J., Brown, K.S. and Chazan, M. (2012). Evidence for early hafted hunting technology. Science 338(2012), 942946.Google Scholar
Will, M., Krapp, M., Stock, J.T. and Manica, A. (2021). Different environmental variables predict body and brain size evolution in Homo. Nature Communications 12(1), 112.Google Scholar
Williams, F.L. (2013). Dietary reconstruction of Pliocene Parapapio whitei from Makapansgat, South Africa, using dental microwear texture analysis. Folia Primatologia 85, 2137.Google Scholar
Williams, F.L.E. and Geissler, E. (2014). Reconstructing the diet and paleoecology of Plio-Pleistocene Cercopithecoides williamsi from Sterkfontein, South Africa diet and paleoecology in C. williamsi. Palaios 29(9), 483494.Google Scholar
Williams, F.L. and Patterson, J.W. (2010). Reconstructing the paleoecology of Taung, South Africa from low magnification of dental microwear features in fossil primates. Palaios 25, 439448.Google Scholar
Williams, L.A.J. and Chapman, G.R. (1986). Relationships between major structures, salic volcanism and sedimentation in the Kenya Rift from the equator northwards to Lake Turkana. In: Frostick, L.E., Renaut, R.W., Rein, I. and Tiercelin, J.J. (Eds.), Sedimentation in the African Rifts. Geological Society Special Publication No. 25. London: Geological Society of London, pp. 5974.Google Scholar
Williams, M.A., Williams, F.H., Gasse, F., Curtis, G.H. and Adamson, D.A. (1979). Pleistocene environments at Gadeb prehistoric site, Ethiopia. Nature 282, 2933.Google Scholar
Williamson, P.G. (1990). Late Cenozoic mollusc faunas from the north west African rift (Uganda–Zaire). In: Boaz, N.T. (Ed.), Evolution of Environments and Hominidae in the African Western Rift Valley. Martinsville: Virginia Museum of Natural History, pp. 125139.Google Scholar
Willig, M.R. and Scheiner, S.M. (2011). The state of theory in ecology. In: Scheiner, S.M. and Willig, M.R. (Eds.), The Theory of Ecology. Chicago: University of Chicago Press, pp. 333347.Google Scholar
Willig, M.R., Kaufman, D.M. and Stevens, R.D. (2003). Latitudinal gradients of biodiversity: pattern, process, scale, and synthesis. Annual Review of Eecology, Evolution, and Systematics 34(1), 273309.Google Scholar
Wilkinson, M.J., Schoville, B.J., Brown, K.S. and Chazan, M. (2012). Evidence for early hafted hunting technology. Science 338, 942946.Google Scholar
Wilkinson, R.D., Steiper, M.E., Soligo, C., et al. (2011). Dating primate divergences through an integrated analysis of palaeontological and molecular data. Systematic Biology 60, 1631.Google Scholar
Winkler, A.J. (2002). Neogene paleobiogeography and East African paleoenvironments: contributions from the Tugen Hills rodents and lagomorphs. Journal of Human Evolution 42(1–2), 237–56.Google Scholar
Winkler, A.J., Denys, C. and Avery, D.M. (2010). Rodentia. In: Werdelin, L. and Sanders, W.J. (Eds.), Cenozoic Mammals of Africa. Berkeley: University of California Press, pp. 263304.Google Scholar
WoldeGabriel, G., White, T., Suwa, G., et al. (1992). Kesem-Kebena: a newly discovered paleoanthropological research area in Ethiopia. Journal of Field Archaeology 19, 471493.Google Scholar
WoldeGabriel, G., White, T.D., Suwa, G., et al. (1994). Ecological and temporal placement of early Pliocene hominids at Aramis, Ethiopia. Nature 371, 330333.Google Scholar
WoldeGabriel, G., Haile-Selassie, Y., Renne, P.R., et al. (2001). Geology and paleontology of the Late Miocene Middle Awash Valley, Afar rift, Ethiopia. Nature 412, 175178.Google Scholar
WoldeGabriel, G., Gilbert, W.H., Hart, W.K., Renne, P.R. and Ambrose, S. (2008). Geology and geochronology. In: Gilbert, W.H. and Asfaw, B. (Eds.), Homo erectus: Pleistocene Evidence from the Middle Awash, Ethiopia. Berkeley: University of California Press, pp. 1343.Google Scholar
WoldeGabriel, G., Ambrose, S. H., Barboni, D., et al. (2009a). The geological, isotopic, botanical, invertebrate, and lower vertebrate surroundings of Ardipithecus ramidus. Science 326, 65655.Google Scholar
WoldeGabriel, G., Hart, W.K., Renne, P.R., Haile-Selassie, Y. and White, T. (2009b). Stratigraphy of the Adu-Asa Formation. In: Haile-Selassie, Y. and WoldeGabriel, G. (Eds.), Ardipithecus kadabba: Late Miocene Evidence from the Middle Awash, Ethiopia. Berkeley: University of California Press, pp. 2762.Google Scholar
WoldeGabriel, G., Endale, T., White, T.D., et al. (2013). The role of tephra studies in African paleoanthropology as exemplified by the Sidi Hakoma Tuff. Journal of African Earth Sciences 77, 4158.Google Scholar
Wolfenden, E., Ebinger, C., Yirgu, G., Renne, P.R. and Kelley, S.P. (2005). Evolution of a volcanic rifted margin: Southern Red Sea, Ethiopia. Geological Society of America Bulletin 117, 846864.Google Scholar
Wolff, R.G. (1975). Sampling and sample size in ecological analyses of fossil mammals. Paleobiology 1, 195204.Google Scholar
Wolpoff, M.H., Senut, B., Pickford, M. and Hawks, J. (2002). Sahelanthropus or ‘Sahelpithecus? Nature 419, 581582.Google Scholar
Wolverton, S. and Lyman, R.L. (2012). Conservation Biology and Applied Zooarchaeology. Tucson: University of Arizona Press.Google Scholar
Wood, B.A. (1991). Koobi Fora Research Project Volume 4: Hominid Cranial Remains. Oxford: Clarendon Press.Google Scholar
Wood, B. and Boyle, E.K. (2016). Hominin taxic diversity: fact or fantasy? American Journal of Physical Anthropology 159, 3778Google Scholar
Wood, B. and Constantino, P. (2007). Paranthropus boisei: fifty years of evidence and analysis. Yearbook of Physical Anthropology 50, 106132.Google Scholar
Wood, B. and Leakey, M. (2011). The Omo-Turkana Basin fossil hominins and their contribution to our understanding of human evolution in Africa. Evolutionary Anthropology 20, 264292.Google Scholar
Wood, B. and Richmond, B.G. (2000). Human evolution: taxonomy and paleobiology. Journal of Anatomy 196, 1960.Google Scholar
Wood, P.C. (2003). Fossil turtles from Lothagam. In: Leakey, M. G. and Harris, J. M. (Eds.), Lothagam: The Dawn of Humanity in Eastern Africa. New York: Columbia University Press, pp. 115136.Google Scholar
Wood, R.C. (1973). A possible correlation between the ecology of living African pelomedusid turtles and their relative abundance in the fossil record. Copeia 1973, 627629.Google Scholar
Woodborne, S., Hall, G., Robertson, I., et al. (2015). A 1000-year carbon isotope rainfall proxy record from South African baobab trees (Adansonia digitate L.). PLoS ONE 10, e0124202.Google Scholar
Woodhead, J., Hellstrom, J., Maas, R., et al. (2006). U–Pb geochronology of speleothems by MC-ICPMS. Quaternary Geochronology 1(3), 208221.Google Scholar
Woodward, A.S. (1921). A new cave man from Rhodesia, South Africa. Nature 108, 371372.Google Scholar
Wrangham, R.W. (1980). Bipedal locomotion as a feeding adaptation in gelada baboons, and its implications for hominid evolution. Journal of Human Evolution 9, 329331.Google Scholar
Wronski, T. and Hausdorf, B. (2008). Distribution patterns of land snails in Ugandan rain forests support the extistence of Pleistocene forest refugia. Journal of Biogeography 35, 17591768.Google Scholar
Wurz, S. (1999). The Howiesons Poort backed artefacts from Klasies River: an argument for symbolic behaviour. The South African Archaeological Bulletin 54(169), 3850.Google Scholar
Wurz, S. (2002). Variability in the middle stone age lithic sequence, 115,000–60,000 years ago at Klasies river, South Africa. Journal of Archaeological Science 29(9), 10011015.Google Scholar
Wurz, S., Bentsen, S.E., Reynard, J., et al. (2018). Connections, culture and environments around 100 000 years ago at Klasies River main site. Quaternary International 495, 102115.Google Scholar
Wynn, J.G. (2000). Paleosols, stable carbon isotopes, and paleoenvironmental interpretation of Kanapoi, Northern Kenya. Journal of Human Evolution 39, 411432.Google Scholar
Wynn, J.G. (2004). Influence of Plio-Pleistocene aridification on human evolution: evidence from paleosols of the Turkana Basin, Kenya. American Journal of Physical Anthropology 123, 106118.Google Scholar
Wynn, J.G. and Bedaso, Z.K. (2010). Is the Pliocene Ethiopian monsoon extinct? A comment on Aronson et al. (2008). Journal of Human Evolution 59, 133138.Google Scholar
Wynn, J.G., Alemseged, Z., Bobe, R., et al. (2006). Geological and palaeontological context of a Pliocene juvenile hominin at Dikika, Ethiopia. Nature 443, 332336.Google Scholar
Wynn, J.G., Roman, D.C., Alemseged, Z., et al. (2008). Stratigraphy, depositional environments, and basin strcuture of the Hadar and Busidima Formations at Dikika, Ethiopia. In: Quade, J. and Wynn, J.G. (Eds.), The Geology of Early Humans in the Horn of Africa. Boulder: Geological Society of America, pp. 87118.Google Scholar
Wynn, J.G., Sponheimer, M., Kimbel, W.H., et al. (2013). Diet of Australopithecus afarensis from the Pliocene Hadar Formation, Ethiopia. Proceedings of the National Academy of Sciences 110, 1049510500.Google Scholar
Wynn, J.G., Reed, K.E., Sponheimer, M., et al. (2016). Dietary flexibility of Australopithecus afarensis in the face of paleoecological change during the middle Pliocene: faunal evidence from Hadar, Ethiopia. Journal of Human Evolution 99, 93106.Google Scholar
Wynn, J.G., Alemseged, Z., Bobe, R., et al. (2020). Isotopic evidence for the timing of the dietary shift toward C4 foods in eastern African Paranthropus. Proceedings of the National Academy of Sciences 117(36), 2197821984.Google Scholar
Yalden, D.W., Largen, M.J., Kock, D. and Hillman, J.C. (1996). Catalogue of the mammals of Ethiopia and Eritrea. 7. Revised checklist, zoogeography and conservation. Tropical Zoology 9, 73164.Google Scholar
Yasui, K., Kunimatsu, Y., Kuga, N., Bajope, B. and Ishida, H. (1992). Fossil mammals from the Neogene strata in the Sinda Basin, eastern Zaire. African Study Monographs 17, 87107.Google Scholar
Yemane, T. (1997). Stratigraphy and sedimentology of the Hadar Formation. PhD dissertation, Iowa State University.Google Scholar
Yoder, A.D. and Yang, Z. (2000). Estimation of primate speciation dates using local molecular clocks. Molecular Biology and Evolution 17(7), 10811090.Google Scholar
Yost, C.L., Ivory, S.J., Deino, A.L., et al. (2021). Phytoliths, pollen, and microcharcoal from the Baringo Basin, Kenya reveal savanna dynamics during the Plio-Pleistocene transition. Palaeogeography, Palaeoclimatology, Palaeoecology 570, 109779.Google Scholar
Young, R.B. (1925). The calcareous tufa deposits of the Campbell Rand, from Boetsap to Taungs Native Reserve. Transvaal Geological Society of South Africa 28, 5567.Google Scholar
Zachos, J., Pagani, M., Sloan, L., Thomas, E. and Billups, K. (2001). Trends, rhythms, and aberrations in global climate 65 Ma to present. Science 292, 686693.Google Scholar
Zaitsev, A.N., Wenzel, T., Spratt, J., et al. (2011). Was Sadiman volcano a source for the Laetoli Footprint Tuff? Journal of Human Evolution 61, 121124.Google Scholar
Zaitsev, A.N., Marks, M.A.W., Wenzel, T., et al. (2012). Mineralogy, geochemistry and petrology of the phonolitic to nephelinitic Sadiman volcano, Crater Highlands, Tanzania. Lithos 152, 6683.Google Scholar
Zaitsev, A.N., Spratt, J., Sharygin, V.V., et al. (2015). Mineralogy of the Laetoli Footprint Tuff: a comparison with possible volcanic sources from the Crater Highlands and Gregory Rift. Journal of African Earth Sciences 111, 214221.Google Scholar
Zanolli, C., Bondioli, L., Candilio, F., et al. (2013). Endostructural morphology of the late Early Pleistocene human dental remains from Uadi Aalad and Mulhuli-Amo, Danakil (Afar) depression of Eritrea. American Journal of Physical Anthropology 150(S56), 298 (abstract).Google Scholar
Zanolli, C., Bondioli, L., Coppa, A., et al. (2014). The late Early Pleistocene human dental remains from Uadi Aalad and Mulhuli-Amo (Buia), Eritrean Danakil: macromorphology and microstructure. Journal of Human Evolution 74, 96113.Google Scholar
Zavada, M.S. and Cadman, A. (1993). Palynological investigations at the Makapansgat Limeworks: an australopithecine site. Journal of Human Evolution 25, 337350.Google Scholar
Zazzo, A., Bocherens, H., Brunet, M., et al. (2000). Herbivore paleodiet and paleoenvironmental changes in Chad during the Pliocene using stable isotope ratios of tooth enamel carbonate. Paleobiology 26(2), 294309.Google Scholar
Zhang, Z., Ramstein, G., Schuster, M., et al. (2014). Aridification of the Sahara desert caused by Tethys Sea shrinkage during the Late Miocene. Nature 513(7518), 401404.Google Scholar
Ziegler, M., Simon, M.H., Hall, I.R., et al. (2013). Development of Middle Stone Age innovation linked to rapid climate change. Nature Communications 4, 1905.Google Scholar
Zilberman, U., Smith, P., Piperno, M. and Condemi, S. (2004). Evidence of amelogenesis imperfecta in an early African Homo erectus. Journal of Human Evolution 46(6), 647653.Google Scholar
Zinke, J., Dullo, W.-C., Heiss, G.A. and Eisenhauer, A. (2004). ENSO and Indian Ocean subtropical dipole variability is recorded in a coral record off southwest Madagascar for the period 1659 to 1995. Earth and Planetary Science Letters 228, 177194.Google Scholar
Zollikofer, C.P.E., Ponce, de León, M.S., Lieberman, D.E., et al. (2005). Virtual cranial reconstruction of Sahelanthropus tchadensis. Nature 434, 755759.Google Scholar
Zouhri, S., Geraads, D., El Boughabi, S. and El Harfi, A. (2012). Discovery of an Upper Miocene vertebrate fauna near Tizi N’Tadderht, Skoura, Ouarzazate Basin (Central High Atlas, Morocco). Comptes Rendus Palevol 11, 455461.Google Scholar
Zouhri, S., Benammi, M., Geraads, D. and El Boughabi, S. (2017). Mammifères du Néogène continental du Maroc: Faunes, biochronologie et paléobiogéographie. Mémoires de la Société Géologique de France 180, 527588.Google Scholar
Zuberbühler, K. (2001). Predator-specific alarm calls in Campbell’s monkeys, Cercopithecus campbelli. Behavioral Ecology and Sociobiology 50, 414422.Google Scholar
Zuberbühler, K., Noë, R. and Seyfarth, R.M. (1997). Diana monkey long-distance calls: messages for conspecifics and predators. Animal Behaviour 53, 589604.Google Scholar
Zwane, B. (2018). A reconstruction of the Late Holocene environment using archaeological charcoal from Klasies River main site cave 1, southern Cape. Unpublished MSc dissertation, University of the Witwatersrand.Google Scholar

Save book to Kindle

To save this book to your Kindle, first ensure coreplatform@cambridge.org is added to your Approved Personal Document E-mail List under your Personal Document Settings on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part of your Kindle email address below. Find out more about saving to your Kindle.

Note you can select to save to either the @free.kindle.com or @kindle.com variations. ‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi. ‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.

Find out more about the Kindle Personal Document Service.

  • Volume References
  • Edited by Sally C. Reynolds, Bournemouth University, René Bobe, University of Oxford
  • Book: African Paleoecology and Human Evolution
  • Online publication: 19 May 2022
  • Chapter DOI: https://doi.org/10.1017/9781139696470.040
Available formats
×

Save book to Dropbox

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Dropbox.

  • Volume References
  • Edited by Sally C. Reynolds, Bournemouth University, René Bobe, University of Oxford
  • Book: African Paleoecology and Human Evolution
  • Online publication: 19 May 2022
  • Chapter DOI: https://doi.org/10.1017/9781139696470.040
Available formats
×

Save book to Google Drive

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Google Drive.

  • Volume References
  • Edited by Sally C. Reynolds, Bournemouth University, René Bobe, University of Oxford
  • Book: African Paleoecology and Human Evolution
  • Online publication: 19 May 2022
  • Chapter DOI: https://doi.org/10.1017/9781139696470.040
Available formats
×