Skip to main content Accessibility help
×
Hostname: page-component-77c89778f8-m42fx Total loading time: 0 Render date: 2024-07-17T06:59:56.797Z Has data issue: false hasContentIssue false

References

Published online by Cambridge University Press:  22 April 2019

D. Margaret Avery
Affiliation:
Iziko Museums of South Africa

Summary

Type
Chapter
Information
Publisher: Cambridge University Press
Print publication year: 2019
Creative Commons
Creative Common License - CCCreative Common License - BYCreative Common License - NCCreative Common License - ND
This content is Open Access and distributed under the terms of the Creative Commons Attribution licence CC-BY-NC-ND 4.0 https://creativecommons.org/cclicenses/

References

Ackermann, R. R., Mackay, A. & Arnold, M. L. (2016). The hybrid origin of ‘modern’ humans. Evolutionary Biology, 43, 111.CrossRefGoogle Scholar
Adams, J. W. (2012a). 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, 717.Google Scholar
Adams, J. W. (2012b). A revised listing of fossil mammals from the Haasgat cave system ex situ deposits (HGD), South Africa. Palaeontologia Electronica, 15, 188.Google Scholar
Adams, J. W. & Conroy, G. C. (2005). Plio-Pleistocene faunal remains from the Gondolin GD 2 in situ assemblage, North West Province, South Africa. In Lieberman, D., Smith, R. J. & Kelley, J., eds. Interpreting the Past: Essays on Human, Primate and Mammal Evolution in Honor of David Pilbeam. Boston, MA: Brill Academic, pp. 243261.Google Scholar
Adams, J. W., Hemingway, J., Kegley, A. D. T. & Thackeray, J. F. (2007a). Luleche, a new paleontological site in the Cradle of Humankind, North-West Province, South Africa. Journal of Human Evolution, 53, 751754.CrossRefGoogle ScholarPubMed
Adams, J. W., Herries, A. I. R., Kuykendall, K. I. & Conroy, G. C. (2007b). Taphonomy of a South African cave: geological and hydrological influences on the GD 1 fossil assemblage at Gondolin, a Plio-Pleistocene paleocave system in the Northwest Province, South Africa. Quaternary Science Reviews, 26, 25262543.CrossRefGoogle 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, 685691.Google Scholar
Adams, J. W., Kegley, A. D. T. & 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, 693698.CrossRefGoogle ScholarPubMed
Afzelius, A. (1815). De antilopis in genere et speciatim guineensibus commentatio. Nova Acta Reginae Societatis Scientarium Upsaliensis, Series 2, 7, 195270.Google Scholar
Aguilar, G. H. (1993). The karyotype and taxonomic status of Cryptomys hottentotus darlingi (Rodentia: Bathyergidae). South African Journal of Zoology, 28, 201204.CrossRefGoogle Scholar
Alba, D. M., Vinuesa, V. & Madurell-Malapeira, J. (2015). On the original author and year of description of the extinct hyaenid Pachycrocuta brevirostris. Acta Palaeontologica Polonica, 60, 573576.Google Scholar
Allen, G. M. (1939). A checklist of African mammals. Bulletin of the Museum of Comparative Zoology Harvard, 83, 1763.Google Scholar
Andersen, K. (1904). Five new Rhinolophi from Africa. Annals and Magazine of Natural History, Series 7, 14, 378388.CrossRefGoogle Scholar
Andersen, K. (1905). Further descriptions of new Rhinolophi from Africa. Annals and Magazine of Natural History, Series 7, 15, 7076.Google Scholar
Andrews, C. W. (1911). On a new species of Deinotherium from British East Africa. Proceedings of the Zoological Society of London, 1911, 943945.CrossRefGoogle Scholar
Andrews, C. W. (1914). Appendix II: on the lower Miocene vertebrates from British East Africa, collected by Dr Felix Oswald. The Quarterly Journal of the Geological Society of London, 70, 163186.CrossRefGoogle Scholar
Andrews, C. W. (1916). Note on a new baboon (Simopithecus oswaldi gen. et sp. nov.) from the (?)Pliocene of British East Africa. Annals and Magazine of Natural History, Series 8, 18, 410419.CrossRefGoogle Scholar
Andrews, C. W. & Beadnell, H. J. L. (1902). A Preliminary Note on Some New Mammals from the Upper Eocene of Egypt. Cairo: National Printing Department.Google Scholar
Angas, G. F. (1848). Description of Tragelaphus angasii, Gray, with some account of its habits. Proceedings of the Zoological Society of London, 16, 8990.Google Scholar
Anonymous (1827). [Review of] Histoire Naturelle des Mammifères, avec des figures originales, dessinées d’après des animaux vivants; et. Par MM Geoffroy Saint-Hilaire et F. Cuvier. Livraisons 52de et 53ème. Zoological Journal, 3, 140143.Google Scholar
Antonites, A., Antonites, A. R., Kruger, N. & Roodt, F. (2014). Report on excavations at Penge, a first-millennium Doornkop settlement. Southern African Humanities, 26, 177192.Google Scholar
Arambourg, C. (1942). L’Elephas recki Dietrich : Exposition systématique et ses affinités. Bulletin de la Société Géologique de France, Series 5, 12, 7387.CrossRefGoogle Scholar
Arambourg, C. (1947). Fascicle 3: Contribution à l’étude géologique et paléontologique du basin du lac Rodolphe et de la basse vallée de l’Omo. 2. Paléontologie. In Arambourg, C., ed. Mission Scientifiques de l’Omo, 1932–1933, Vol. 1. Géologie-Anthropologie. Paris: Muséum National d’Histoire Naturelle, pp. 232562.Google Scholar
Arambourg, C. (1949). Numidocapra crassicornis, nov. gen, nov. sp., un Ovicapriné nouveau du Villafranchien constaninois. Compte Rendu Sommaire des Séances de la Société Géologique de France, Series 5, 13, 290291.Google Scholar
Armstrong, A. (2016). Small mammal utilization by Middle Stone Age humans at Die Kelders Cave 1 and Pinnacle Point Site 5–6, Western Cape Province, South Africa. Journal of Human Evolution, 101, 1744.Google Scholar
Arnold, E. A. (2008). A consideration of livestock exploitation during the Early Iron Age in the Thukela valley, KwaZulu-Natal. In Badenhorst, S., Mitchell, P., & Driver, J. C., eds. Animals and People: Archaeozoological Papers in Honour of Ina Plug. Oxford: British Archaeological Reports, pp. 152168.Google Scholar
Asher, R. J. & Avery, D. M. (2010). New golden moles (Afrotheria, Chrysochloridae) from the Early Pliocene of South Africa. Palaeontologia Electronica, 13, 112.Google Scholar
Astre, G. (1929). Sur un félin a particularités ursoïdes des limons pliocènes du Roussillon. Compte Rendu Sommaire et Bulletin de la Société Géologique de France, Series 4, 29, 199204.Google Scholar
Avery, D. M. (1977). Past and present distribution of some rodent and insectivore species in the southern Cape Province, South Africa: new information. Annals of the South African Museum, 74, 201209.Google Scholar
Avery, D. M. (1979). Appendix 2: the micromammalian fauna from the Late Stone Age levels at Die Kelders. In Schweitzer, F. R.. Excavations at Die Kelders, Cape Province, South Africa. The Holocene deposits. Annals of the South African Museum, 78, 229232.Google Scholar
Avery, D. M. (1981). Holocene micromammalian faunas from the northern Cape Province, South Africa. South African Journal of Science, 77, 265273.Google Scholar
Avery, D. M. (1982a). The micromammalian fauna from Border Cave, KwaZulu, South Africa. Journal of Archaeological Science, 9, 187204.Google Scholar
Avery, D. M. (1982b). 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. (1984a). Appendix 2: the micromammalian fauna from Fairview Shelter. In Robertshaw, P. T.. Fairview Rockshelter: a contribution to the prehistory of the eastern Cape Province of South Africa. Annals of the Cape Provincial Museums (Human Sciences), 1, 8788.Google Scholar
Avery, D. M. (1984b). Micromammals and environmental change at Zebrarivier, central Namibia. Journal of the South West Africa Scientific Society, 38, 7986.Google Scholar
Avery, D. M. (1985). The dispersal of brown rats Rattus norvegicus and new specimens from 19th century Cape Town. Mammalia, 49, 573576.Google 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, 405421.Google Scholar
Avery, D. M. (1988). Comments on micromammalian fauna from Nkupe Shelter. Mazel, A.D. Nkupe Shelter: report on excavations in the eastern Biggarsberg, Thukela Basin, Natal, South Africa. Annals of the Natal Museum, 29, 376377.Google Scholar
Avery, D. M. (1989). Remarks concerning vertebrate faunal remains from the main house at Paradise. South African Archaeological Bulletin, 44, 114116.Google Scholar
Avery, D. M. (1991a). Micromammals, owls and vegetation change in the eastern Cape Midlands, South Africa, during the last millennium. Journal of Arid Environments, 20, 357369.Google Scholar
Avery, D. M. (1991b). Late Quaternary incidence of some micromammals in Natal. Durban Museum Novitates, 16, 111.Google Scholar
Avery, D. M. (1992a). The environment of early modern humans at Border Cave, South Africa: micromammalian evidence. Palaeogeography, Palaeoclimatology, Palaeoecology, 91, 7187.CrossRefGoogle Scholar
Avery, D. M. (1992b). Micromammals from Collingham Shelter. Natal Museum Journal of Humanities, 4, 6163.Google Scholar
Avery, D. M. (1992c). Micromammals and the environment of early pastoralists at Spoeg River, western Cape Province, South Africa. South African Archaeological Bulletin, 47, 116121.Google Scholar
Avery, D. M. (1995a). A preliminary assessment of the micromammalian remains from Gladysvale Cave, South Africa. Palaeontologia Africana, 32, 110.Google Scholar
Avery, D. M. (1995b). Southern savannas and Pleistocene hominid adaptations: the micromammalian perspective. In Vrba, E. S., Denton, G. H., Partridge, T. C. & Burckle, L. H., eds. Paleoclimate and Evolution with Emphasis on Human Origins. New Haven, CT: Yale University Press, pp. 459478.Google Scholar
Avery, D. M. (1996). Appendix 1: comments on micromammals from Maqonqo, KwaZulu-Natal. In Mazel, A. D., Maqonqo Shelter: the excavation of Holocene deposits in the eastern Biggarsberg, Thukela Basin, South Africa. Natal Museum Journal of Humanities, 8, 39.Google Scholar
Avery, D. M. (1997a). Micromammals and the Holocene environment of Mzinyashana Shelter 1. Natal Museum Journal of Humanities, 9, 3746.Google Scholar
Avery, D. M. (1997b). 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.CrossRefGoogle Scholar
Avery, D. M. (1999). Holocene coastal environments in the Western Cape Province, South Africa: micromammalian evidence from Steenbokfontein. Archaeozoologia, 10, 163180.Google Scholar
Avery, D. M. (2000). Notes on the systematics of micromammals from Sterkfontein, Gauteng, South Africa. Palaeontologia Africana, 36, 8390.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. (2004). Size variation in the common molerat Cryptomys hottentotus from southern Africa and its potential for palaeoenvironmental reconstruction. Journal of Archaeological Science, 31, 273282.CrossRefGoogle 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. & Avery, G. (2011). Micromammals in the Northern Cape Province of South Africa, past and present. African Natural History, 7, 939.Google Scholar
Avery, G. (1992). Appendix 1: faunal remains from Klein Kliphuis Shelter, Clanwilliam District, South Africa. In van Rijssen, W. J.. The Late Holocene deposits at Klein Kliphuis Shelter, Cedarberg, Western Cape Province. South African Archaeological Bulletin, 47, 4043.Google Scholar
Avery, G. (1997). 2nd report on work completed at Spreeuwal and proposed new work (NMC 80/97/12/004/51 Dec. 1997). Cape Town: Iziko South African Museum. Unpublished Report.Google Scholar
Avery, G., Cruz-Uribe, K., Goldberg, P., et al. (1997). The 1992–1993 excavations at the Die Kelders Middle and Later Stone Age cave site, South Africa. Journal of Field Archaeology, 24, 263291.Google Scholar
Avery, G., Halkett, D., Orton, J., et al. (2008). The Ysterfontein 1 Middle Stone Age rock shelter and the evolution of coastal foraging. South African Archaeological Society Goodwin Series, 10, 6689.Google Scholar
Aymard, A. (1846). Essaie monographique sur un nouveau genre de mammifère fossile trouvé dans la Haute-Loire, et nommé Entelodon. Annales de la Société d’Agriculture, Sciences, Arts et Commerce du Puy, 12, 227268.Google Scholar
Azanza, B., Morales, J. & Pickford, M. (2003). On the nature of the multibranched cranial appendages of the climacoceratid Orangemeryx hendeyi. Memoirs of the Geological Survey of Namibia, 19, 345357.Google Scholar
Backwell, L., Bhagwandas-Jogibhai, K., Fenn, G., et al. (1996). Twyfelpoort Shelter: a Later Stone Age sequence. Southern African Field Archaeology, 5, 8495.Google Scholar
Backwell, L., McCarthy, T. S., Wadley, L., et al. (2014). Multiproxy record of late Quaternary climate change and Middle Stone Age human occupation at Wonderkrater, South Africa. Quaternary Science Reviews, 99, 4259.Google Scholar
Badenhorst, S. (2003). The archaeofauna from iNkolimahashi Shelter, a Later Stone Age shelter in the Thukela Basin, KwaZulu-Natal, South Africa. Southern African Humanities, 15, 4557.Google Scholar
Badenhorst, S. (2014). Appendix B: hunters or herders? The fauna from Kuidas Spring, a Late Holocene stone circle site in Namibia. In Veldman, A.. The archaeology of a rock shelter and a stone circle at Kuidas Spring, north-west Namibia. Unpublished MA thesis, University of Johannesburg, pp. 164–188.Google Scholar
Badenhorst, S. & Boshoff, W. S. (2015). Animal remains from an early twentieth century rural farming community at Orange River Railway Station, South Africa. Navorsinge van die Nasionale Museum Bloemfontein, 31, 4964.Google Scholar
Badenhorst, S. & Plug, I. (2001). Appendix: the faunal remains from Mmatshetshele, a Late Iron Age site in the Rustenburg District. In Pistorius, J. C. C.. Late Iron Age sites on Mmatshetshele Mountain in the Central Bankenveld of the North West Province, South Africa. South African Archaeological Bulletin, 56, 5556.Google Scholar
Badenhorst, S. & Plug, I. (2002). Appendix: animal remains from recent excavations at a Late Iron Age site, Simunye, Swaziland. In Ohinata, F.. The beginning of ‘Tsonga’ archaeology: excavations at Simunye, north-eastern Swaziland. Southern African Humanities, 14, 4550.Google Scholar
Badenhorst, S. & Plug, I. (2003). The archaeozoology of goats, Capra hircus (Linnaeus, 1758): their size variation during the last two millennia in southern Africa (Mammalia: Artiodactyla: Caprini). Annals of the Transvaal Museum, 40, 91121.Google Scholar
Badenhorst, S. & Plug, I. (2004/2005). Boleu: faunal analysis from a 19th century site in the Groblersdal area, Mpumalanga, South Africa. Southern African Field Archaeology, 13–14, 1318.Google Scholar
Badenhorst, S. & 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., Plug, I., Pelser, A. J. & Van Vollenhoven, A. C. (2002). Faunal analysis from Steinaecker’s Horse, the northernmost British military outpost in the Kruger National Park during the South African War. Annals of the Transvaal Museum, 39, 5763.Google Scholar
Badenhorst, S., Plug, I. & Boshoff, W. S. (2011a). Faunal remains from test excavations at Middle and Late Iron Age sites in the Limpopo Valley, South Africa. Annals of the Ditsong National Museum of Natural History, 1, 2331.Google Scholar
Badenhorst, S., Sénégas, F., Gommery, D., et al. (2011b). 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, 3340.Google Scholar
Badenhorst, S., Sinclair, P., Ekblom, A. & Plug, I. (2011c). Faunal remains from Chibuene, an Iron Age coastal trading station in central Mozambique. Southern African Humanities, 23, 115.Google Scholar
Badenhorst, S., Parsons, I. & Voigt, E. A. (2015). Fauna from five Later Stone Age sites in the Bushmanland region of South Africa. Annals of the Ditsong National Museum of Natural History, 5, 110.Google Scholar
Badenhorst, S., Van Niekerk, K. L. & Henshilwood, C. S. (2016a). Large mammal remains from the 100 Ka Middle Stone Age layers of Blombos Cave, South Africa. South African Archaeological Bulletin, 71, 4652.Google Scholar
Badenhorst, S., Veldman, A. & Lombard, M. (2016b). Late Holocene fauna from Kuidas Spring in Namibia. African Archaeological Review, 33, 2944.Google Scholar
Bain, A. G. (1839). [Ox from the Modder River]. Proceedings of the Geological Society of London, 3, 152.Google Scholar
Baker, C. M. (1992). Atilax paludinosus. Mammalian Species, 408, 16.Google Scholar
Barry, R. E. & Shoshani, J. (2000). Heterohyrax brucei. Mammalian Species, 645, 17.Google Scholar
Bate, D. M. A. (1949). A new African long-horned buffalo. Annals and Magazine of Natural History, Series 12, 11, 396398.Google Scholar
Beaumont, P. B. & Boshier, A. K. (1974). Report on test excavations in a prehistoric pigment mine near Postmasburg, northern Cape. South African Archaeological Bulletin, 29, 4159.Google Scholar
Beaumont, P. B., De Villiers, H. & Vogel, J. C. (1978). Modern man in sub-Saharan Africa prior to 49 000 years BP: a review and evaluation with particular reference to Border Cave. South African Journal of Science, 74, 409419.Google Scholar
Beck, R. (1906). Mastodon in the Pleistocene of South Africa. Geological Magazine, New Series 5, 3, 4950.Google Scholar
Bender, P. A. (1990). A reconsideration of the fossil Suidae of the Makapansgat Limeworks, Potgietersrus, northern Transvaal. Unpublished MSc thesis, University of the Witwatersrand.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, 166.Google Scholar
Bender, P. A. & Brink, J. S. (1992). Preliminary report on new large-mammal fossil finds from the Cornelia-Uitzoek site, in the north-eastern Orange Free State. South African Journal of Science, 88, 512515.Google Scholar
Benefit, B. R. & Pickford, M. (1986). Miocene fossil cercopithecoids from Kenya. American Journal of Physical Anthropology, 69, 441464.Google Scholar
Bennett, N. C., Maree, S. & Faulkes, C. G. (2006). Georychus capensis. Mammalian Species, 799, 14.Google Scholar
Bennett, N. C., Faulkes, C. G., Hart, L. & Jarvis, J. U. M. (2009). Bathyergus suillus (Rodentia: Bathyergidae). Mammalian Species, 828, 17.CrossRefGoogle Scholar
Bennett, N., Jarvis, J., Visser, J. & Maree, S. (2016). A conservation assessment of Georychus capensis. In Child, M. F., Roxburgh, L., Do Linh San, E., Raimondo, D. & Davies-Mostert, H. T., eds. The Red List of Mammals of South Africa, Swaziland and Lesotho. South Africa: South African National Biodiversity Institute and Endangered Wildlife Trust.Google Scholar
Berger, L. R. (1992). Early hominid fossils discovered at Gladysvale Cave, South Africa. South African Journal of Science, 88, 362.Google Scholar
Berger, L. R. (1993). A preliminary estimate of the age of the Gladysvale australopithecine site. Palaeontologia Africana, 30, 5155.Google Scholar
Berger, L. R. & Brink, J. (1996). Late Middle Pleistocene fossils, including a human patella, from the Riet River gravels, Free State, South Africa. South African Journal of Science, 92, 13.Google Scholar
Berger, L. R. & Brink, J. (2001). An Atlas of Southern African Mammalian Fossil Bearing Sites: Late Miocene to Late Pleistocene. Johannesburg: University of the Witwatersrand.Google Scholar
Berger, L. R. & 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. & Parkington, J. E. (1995). Brief communication: a new Pleistocene hominid-bearing locality at Hoedjiespunt, South Africa. American Journal of Physical Anthropology, 98, 601609.Google Scholar
Berger, L. R. & Tobias, P. V. (1994). New discoveries at the early hominid site of Gladysvale, South Africa. South African Journal of Science, 90, 223226.Google Scholar
Berger, L. R., Keyser, A. W. & Tobias, P. V. (1993). Gladysvale: first early hominid site discovered in South Africa since 1949. American Journal of Physical Anthropology, 92, 107111.Google Scholar
Berger, L. R., Pickford, M. & Thackeray, J. F. (1995). A Plio-Pleistocene hominid upper central incisor from the Cooper’s sites, South Africa. South African Journal of Science, 91, 541542.Google Scholar
Berger, L. R., Lacruz, R. & 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.Google Scholar
Berger, L. R., De Ruiter, D. J., Steininger, C. M. & Hancox, 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., De Ruiter, D., 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, 2015, 135.Google Scholar
Berkenhout, J. (1769). Outlines of the Natural History of Great Britain and Ireland. London: Elmsly.Google Scholar
Bernor, R. L. & Kaiser, T. L. (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, 149185.Google Scholar
Bernor, R. L., Armour-Chelu, A., Gilbert, H., Kaiser, T. M. & Schulz, E. (2010). Equidae. In Werdelin, L. & Sanders, W. J., eds. Cenozoic Mammals of Africa. Berkeley, CA: University of California Press, pp. 685721.Google Scholar
Berta, A. & Galiano, H. (1983). Megantereon hesperus from the Late Hemphillian of Florida, with remarks on the phylogenetic relationships of machairodonts (Mammalia, Felidae, Machairodontinae). Journal of Paleontology, 57, 892899.Google Scholar
Biedermann, W. G. A. (1863). Petrefakten aus der Umgegend von Winterthur. Winterthur: Bleuler-Hausheer.Google Scholar
Binneman, J. (1997). Results from a test excavation at The Havens Cave, Cambria Valley, south-eastern Cape. Southern African Field Archaeology, 6, 93105.Google Scholar
Binneman, J. (1999). Results from a test excavation at Groot Kommandokloof Shelter, in the Baviaanskloof/Kouga region, Eastern Cape. Southern African Field Archaeology, 8, 100107.Google Scholar
Binneman, J. (2000). Results from two test excavations in the Baviaanskloof Mountains, Eastern Cape Province. Southern African Field Archaeology, 9, 8396.Google Scholar
Binneman, J. (2004/2005). Archaeological research along the south-eastern Cape Coast Part I: open-air shell middens. Southern African Field Archaeology, 13–14, 4977.Google Scholar
Binneman, J. (2006/2007). Archaeological research along the south-eastern Cape coast part 2, caves and shelters: Kabeljous River Shelter 1 and associated stone tool industries. Southern African Field Archaeology, 15–16, 5774.Google Scholar
Birungi, J. & Arctander, P. (2001). Molecular systematics and phylogeny of the Reduncini (Artiodactyla: Bovidae) inferred from the analysis of mitochondrial cytochrome b gene sequences. Journal of Mammalian Evolution, 8, 125147.Google Scholar
Bishop, L. C. (2010). Suoidea. In Werdelin, L. & Sanders, W. J., eds. Cenozoic Mammals of Africa. Berkeley, CA: University of California Press, pp. 823842.Google Scholar
Blumenbach, J. F. (1797). D. Joh. Fr. Blumenbach’s Handbuch der Naturgeschichte, 5th edition. Göttingen: Johann Christian Dieterich.Google Scholar
Boddaert, P. (1784 [1785]). Elenchus Animalium. Rotterdam: C.R. Hake.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, 126.Google Scholar
Bonaparte, C. L. (1832–1841). Iconografia della Fauna Italica: per le quattro classi degli animali vertebrati. Rome: Salviucci.Google Scholar
Boné, E. L. & Dart, R. A. (1955). A catalog of the australopithecine fossils found at the Limeworks, Makapansgat. American Journal of Physical Anthropology, 13, 621624.Google Scholar
Boné, E. L. & Singer, R. (1965). Hipparion from Langebaanweg, Cape Province and a revision of the genus in Africa. Annals of the South African Museum, 48, 273397.Google Scholar
Boshoff, W. & Steyn, M. (2000). A human grave from the farm Hamilton in the Limpopo River Valley (South Africa). Southern African Field Archaeology, 9, 6874.Google Scholar
Boshoff, A., Landman, M. & 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
Botha, J. & Gaudin, T. (2007). An Early Pliocene pangolin (Mammalia; Pholidota) from Langebaanweg, South Africa. Journal of Vertebrate Paleontology, 27, 484491.Google Scholar
Bousman, B. & Brink, J. (2014). Excavation of Middle and Later Stone Age sites at Erfkroon, South Africa. Unpublished NSF Final Report 0918074.Google Scholar
Brain, C. K. (1969). 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, 1121117.Google Scholar
Brain, C. K. (1974). Human food remains from the Iron Age at Zimbabwe. South African Journal of Science, 70, 303309.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. (1981). The Hunters or the Hunted? Chicago, IL: University of Chicago Press.Google Scholar
Brain, C. K. & Brain, V. (1977). Microfaunal remains from Mirabib: some evidence of palaeoecological changes in the Namib. Madoqua, 10, 285293.Google Scholar
Brain, C. K. & Watson, V. (1992). A guide to the Swartkrans early hominid site. Annals of the Transvaal Museum, 35, 343365.Google Scholar
Brain, C. K., Vrba, E. S. & Robinson, J. T. (1974). A new hominid innominate bone from Swartkrans. Annals of the Transvaal Museum, 29, 5563.Google Scholar
Brain, C. K., Churcher, C. 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
Brants, A. (1827). Het Geslacht der Muizen door Linnaeus Opgesteld. Berlin: Akademische Boekdrukkery.Google Scholar
Bräuer, G. & Singer, S. (1996). The Klasies zygomatic bone: archaic or modern? Journal of Human Evolution, 30, 161165.Google Scholar
Bräuer, G., Deacon, H. J. & Zipfel, B. (1992). Comment on the new maxilliary finds from Klasies River, South Africa. Journal of Human Evolution, 23, 419422.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
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. (1993). Postcranial evidence for the evolution of the black wildebeest, Connochaetes gnou: an exploratory study. Palaeontologia Africana, 30, 6169.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. (2004). The taphonomy of an Early/Middle Pleistocene hyaena burrow at Cornelia-Uitzoek, South Africa. Revue de Paléobiologie, 23, 731740.Google Scholar
Brink, J. S. (2005). The evolution of the Black Wildebeest, Connochaetes gnou, and modern large mammal faunas in central southern Africa. Unpublished DPhil thesis, University of Stellenbosch.Google Scholar
Brink, J. S. (2012). Appendix: the fauna from Lithakong. In Kaplan, J. & Mitchell, P.. The archaeology of the Lesotho Highlands Water Project Phases IA and IB. Southern African Humanities, 24, 3032.Google Scholar
Brink, J. S. & Deacon, H. J. (1982). A study of a last interglacial shell midden and bone accumulation at Herold’s Bay, Cape Province, South Africa. Palaeoecology of Africa, 15, 3140.Google Scholar
Brink, J. S. & Holt, S. (1992). A small goat, Capra hircus, from a Late Iron Age site in the eastern Orange Free State. Southern African Field Archaeology, 1, 8891.Google Scholar
Brink, J. S. & Rossouw, L. (2000). New trial excavations at the Cornelia-Uitzoek type locality. Navorsinge van die Nasionale Museum Bloemfontein, 16, 141156.Google Scholar
Brink, J. S. & Stynder, D. (2009). Morphological and trophic distinction in the dentitions of two early alcelaphine bovids from Langebaanweg (genus Damalacra). Palaeontologia Africana, 44, 139193.Google Scholar
Brink, J. S. & Webley, L. (1996). Faunal evidence for pastoral settlement at Jakkalsberg, Richtersveld, Northern Cape Province. Southern African Field Archaeology, 5, 7078.Google Scholar
Brink, J. S., Dreyer, J. J. B. & Loubser, J. H. N. (1992). Rescue excavations at Pramberg, Jacobsdal, south-western Orange Free State. Southern African Field Archaeology, 1, 5460.Google Scholar
Brink, J. S., De Bruiyn, H., Rademeyer, L. B. & Van der Westhuizen, W. A. (1995). A new find of Megalotragus priscus (Alcelaphini, Bovidae) from the central Karoo, South Africa. Palaeontologia Africana, 32, 1722.Google Scholar
Brink, J. S., Berger, L. R. & Churchill, S. E. (1999). Mammalian fossils from erosional gullies (dongas) in the Doring River drainage, central Free State Province, South Africa. In Becker, C., Manhart, J., Peters, J. & Schibler, J., eds. Historium animalium ex ossibus. Beiträge zur Paläoanatomie, Archäologie, Ägyptologie, Ethnologie und Geschichte der Tiermedizin: Festschrift für Angela von den Driesch. Rahden/Westfalen: Verlag Marie Leidorf, pp. 7990.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., Bousman, C. B. & 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. S., Holt, S. & Kolska Horwitz, L. (2015b). Preliminary findings on macro-faunal taxonomy, taphonomy, biochronology and palaeoecology from the basal layers of Wonderwerk Cave, South Africa. Preserving African Cultural Heritage, 93, 137147.Google Scholar
Brink, J. S., Holt, S. & Kolska Horwitz, L. (2016). The Oldowan and Early Acheulean Mammalian fauna of Wonderwerk Cave (Northern Cape Province, South Africa). African Archaeological Review, 33, 223250.Google Scholar
Brisson, M.-J. (1762). Regnum Animale in Classes IX. Distributum, sive Synopsis Methodica, 2nd edition. Lugduni Batavorum: T. Haak.Google Scholar
Britton-Davidian, J., Catalan, J., Granjon, L. & Duplantier, J. (1995). Chromosomal phylogeny and evolution in the genus Mastomys (Mammalia, Rodentia). Journal of Mammalogy, 76, 248262.Google Scholar
Britton-Davidian, J., Robinson, T. J. & Veyrunes, F. (2012). Systematics and evolution of the African pygmy mice, subgenus Nannomys: a review. Acta Oecologica, 42, 4149.Google Scholar
Bronner, G. N. (1995a). Cytogenetic properties of nine species of golden moles. Journal of Mammalogy, 76, 957971.Google Scholar
Bronner, G. N. (1995b). Systematic revision of the golden mole genera Amblysomus, Chlorotalpa and Calcochloris (Insectivora: Chrysochloromorpha; Chrysochloridae). Unpublished PhD thesis, University of Natal (Durban).Google Scholar
Bronner, G. N. (1996). Non-geographic variation in morphological characteristics of the Hottentot golden mole Amblysomus hottentotus (Insectivora: Chrysochloridae). Mammalia, 60, 707727.Google Scholar
Bronner, G. N. (2013). Neamblysomus gunningi. In Kingdon, J., Happold, D., Butynski, T., et al., eds. Mammals of Africa, Vol. 1. London: Bloomsbury Publishing, pp. 239240.Google Scholar
Bronner, G. N. & Jenkins, E. (2005). Order Afrosoricida. In Wilson, D. E. & Reeder, D. M., eds. Mammal Species of the World, Vol. 1. Baltimore, MD: Johns Hopkins University Press, pp. 7181.Google Scholar
Bronner, G. N. & Meester, J. A. J. (1988). Otomys angoniensis. Mammalian Species, 306, 16.Google Scholar
Bronner, G., Gordon, S. & Meester, J. (1988). Otomys irroratus. Mammalian Species, 308, 16.Google Scholar
Bronner, G. N., Hoffmann, M., Taylor, P. J., et al. (2003). A revised systematic checklist of the extant mammals of the southern African subregion. Durban Museum Novitates, 28, 5696.Google Scholar
Bronner, G. N., Van der Merwe, M. & Njobe, K. (2007). Nongeographic cranial variation in two medically important rodents from South Africa, Mastomys natalensis and Mastomys coucha. Journal of Mammalogy, 88, 11791194.Google Scholar
Brook, G. A., Cherkinsky, A., Marias, E. & Todd, N. B. (2014). Rare elephant molar (Loxodonta africana zulu) from the Windhoek Spring Deposit, Namibia. Transactions of the Royal Society of South Africa, 69, 145150.Google Scholar
Brookes, J. (1827). III. Lycaon. In Griffiths, E., Hamilton Smith, C. & Pidgeon, E., eds. Animal Kingdom Arranged in Conformity with its Organization by the Baron Cuvier, with Additional Descriptions of All the Species Hitherto named and of Many Not Before Noticed, Vol. 5. London: Whittaker.Google Scholar
Brookes, J. (1828). A Catalogue of the Anatomical & Zoological Museums of Joshua Brookes, Esq., F.R.S., F.L.S. etc., Part 1. London: Gold & Walton.Google Scholar
Brooks, A. S. & Yellen, J. E. (1977). Archaeological excavations at ≠gi: a preliminary report on the first two field seasons. Botswana Notes and Records, 9, 2130.Google Scholar
Brooks, A. S., Crowell, A. L. & Yellen, J. E. (1980). Gi: a Stone Age archaeological site in the northern Kalahari Desert, Botswana. In Leakey, R. E. F. & Bogot, B. A., eds. Eighth Panafrican Congress of Prehistory and Quaternary Studies (Nairobi, 1977). Nairobi: TILLMIAP, pp. 304309.Google Scholar
Broom, R. (1907a). On some new species of Chrysochloris. Annals and Magazine of Natural History, Series 7, 19, 262268.Google Scholar
Broom, R. (1907b). A contribution to the knowledge of the Cape golden moles. Transactions of the South African Philosophical Society, 18, 283311.Google Scholar
Broom, R. (1908). Further observations on the Chrysochloridae. Annals of the Transvaal Museum, 1, 1416.Google Scholar
Broom, R. (1909a). On a large extinct species of Bubalis. Annals of the South African Museum, 7, 279280.Google Scholar
Broom, R. (1909b). On evidence of a large horse recently extinct in South Africa. Annals of the South African Museum, 7, 281282.Google Scholar
Broom, R. (1909c). Some observations on the dentition of Chrysochloris and on the tritubercular theory. Annals of the Natal Government Museum, 2, 129139.Google Scholar
Broom, R. (1910). On Chrysochloris namaquensis, Broom. Transactions of the Royal Society of South Africa, 2, 4143.Google Scholar
Broom, R. (1913a). A new species of golden mole. Proceedings of the Zoological Society of London, 1913, 546548.Google Scholar
Broom, R. (1913b). Note on Equus capensis Broom. Bulletin of the American Museum of Natural History, 32, 437439.Google Scholar
Broom, R. (1913c). Man contemporaneous with extinct animals in South Africa. Annals of the South African Museum, 12, 1316.Google Scholar
Broom, R. (1918). The evidence afforded by the Boskop skull of a new species of primitive man (Homo capensis). Anthropological Papers of the American Museum of Natural History, 23, 6379.Google Scholar
Broom, R. (1925). On evidence of a giant pig from the Late Tertiaries of South Africa. Records of the Albany Museum, 3(4), 307308.Google Scholar
Broom, R. (1928). On some new mammals from the Diamond Gravels of the Kimberley District. Annals of the South African Museum, 22, 439444.Google Scholar
Broom, R. (1929). The Transvaal fossil human skeleton. Nature, 123, 415416.Google Scholar
Broom, R. (1931). An extinct giant pig from the gravels of Windsorton, South Africa. Records of the Albany Museum, 4, 167168.Google Scholar
Broom, R. (1934). On the fossil remains associated with Australopithecus africanus. South African Journal of Science, 31, 471480.Google Scholar
Broom, R. (1936a). A new fossil anthropoid skull from South Africa. Nature, 138, 486488.Google Scholar
Broom, R. (1936b). A new fossil baboon from the Transvaal. Annals of the Transvaal Museum, 18, 393396.Google Scholar
Broom, R. (1937a). On some new Pleistocene mammals from limestones caves of the Transvaal. South African Journal of Science, 33, 750768.Google Scholar
Broom, R. (1937b). Notices of a few more new fossil mammals from the caves of the Transvaal. Annals and Magazine of Natural History, Series 10, 20, 509514.Google Scholar
Broom, R. (1938a). The Pleistocene anthropoid apes of South Africa. Nature, 142, 377379.Google Scholar
Broom, R. (1938b). Note on the premolars of the elephant shrews. Annals of the Transvaal Museum, 19, 251252.Google Scholar
Broom, R. (1939a). The dentition of the Transvaal Pleistocene anthropoids, Plesianthropus and Paranthropus. Annals of the Transvaal Museum, 19, 303314.Google Scholar
Broom, R. (1939b). The fossil rodents of the limestone cave at Taungs. Annals of the Transvaal Museum, 19, 315317.Google Scholar
Broom, R. (1939c). A preliminary account of the Pleistocene carnivores of the Transvaal Caves. Annals of the Transvaal Museum, 19, 331338.Google Scholar
Broom, R. (1939d). On the affinities of the South African Pleistocene anthropoids. South African Journal of Science, 36, 408411.Google Scholar
Broom, R. (1939e). A restoration of the Kromdraai skull. Annals of the Transvaal Museum, 19, 327329.Google Scholar
Broom, R. (1940). The South African Pleistocene cercopithecid apes. Annals of the Transvaal Museum, 20, 89100.Google Scholar
Broom, R. (1941). On two Pleistocene golden moles. Annals of the Transvaal Museum, 20, 215216.Google Scholar
Broom, R. (1946). Some new and some rare golden moles. Annals of the Transvaal Museum, 20, 329335.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. (1949a). Another new type of fossil ape-man (Paranthropus crassidens). Nature, 163, 903.Google Scholar
Broom, R. (1949b). The extinct blue buck of South Africa. Nature, 164, 10971098.Google Scholar
Broom, R. (1950). Some further advance in our knowledge of the Cape golden moles. Annals of the Transvaal Museum, 21, 234241.Google Scholar
Broom, R. & Jensen, J. S. (1946). A new fossil baboon from the caves at Potgietersrust. Annals of the Transvaal Museum, 20, 337340.Google Scholar
Broom, R. & Robinson, J. T. (1948). A new sub-fossil baboon from Kromdraai, Transvaal. Annals of the Transvaal Museum, 21, 242245.Google Scholar
Broom, R. & Robinson, J. T. (1949a). A new mandible of the ape-man Plesianthropus transvaalensis. American Journal of Physical Anthropology, 7, 123127.Google Scholar
Broom, R. & Robinson, J. T. (1949b). A new type of fossil baboon, Gorgopithecus major. Proceedings of the Zoological Society of London, 119, 379387.Google Scholar
Broom, R. & Robinson, J. T. (1949c). A new type of fossil man. Nature, 164, 322323.Google Scholar
Broom, R. & Robinson, J. T. (1950a). Note on the skull of the Swartkrans ape-man Paranthropus crassidens. American Journal of Physical Anthropology, 8, 295304.Google Scholar
Broom, R. & Robinson, J. T. (1950b). One of the earliest types of man. South African Journal of Science, 47, 5557.Google Scholar
Broom, R. & Robinson, J. T. (1952). Swartkrans ape-man, Paranthropus crassidens. Transvaal Museum Memoir, 6, 1123.Google Scholar
Broom, R. & Schepers, G. W. H. (1946). The South African fossil ape-men: the Australopithecinae. Transvaal Museum Memoir, 2, 1272.Google Scholar
Brophy, J. K., De Ruiter, D. J., Lewis, P. J., Churchill, S. E. & Berger, L. R. (2006). Preliminary investigation of the new Middle Stone Age site of Plovers Lake, South Africa. Current Research in the Pleistocene, 23, 4143.Google Scholar
Brophy, J. K., De Ruiter, D. J., Athreya, S. & 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. J. V. (1981). Appendix 3: an analysis of faunal remains at Broederstroom 24/73. In Mason, R. J.. Early Iron Age settlement at Broederstroom 24/73, Transvaal, South Africa. South African Journal of Science, 77, 416.Google Scholar
Brown, A. J. V. (1988). The faunal remains from Kalkbank, northern Transvaal. Palaeoecology of Africa, 19, 205212.Google Scholar
Brown, A. J. V. & Verhagen, B. T. (1985). Two Antidorcas bondi individuals from the Late Stone Age site of Kruger Cave 35/83, Olifantsnek, Rustenburg District, South Africa. South African Journal of Science, 81, 102.Google Scholar
Brunton, S., Badenhorst, S. & Schoeman, M. H. (2013). Ritual fauna from Ratho Kroonkop: a second millennium AD rain control site in the Shashe-Limpopo Confluence area of South Africa. Azania, 48, 111132.Google Scholar
Burchell, W. J. (1817). Note sur une nouvelle espèce de Rhinoceros. Bulletin des Sciences par la Société Philomathique de Paris, 1817–1819, 9697.Google Scholar
Burchell, W. J. (1824). Travels in the Interior of Southern Africa. London: Longman, Hurst, Rees, Orme, Brown & Green.Google Scholar
Burmeister, H. (1837). Handbuch der Naturgeschichte: zum Gebrauch bei Vorlesungen. Berlin: Enslin.Google Scholar
Butler, P. M. (1956). Erinaceidae from the Miocene of East Africa. Fossil Mammals of Africa, 11, 175.Google Scholar
Butler, P. M. (1965). Fossil Mammals of Africa No. 18: East African Miocene and Pleistocene chalicotheres. Bulletin of the British Museum (Natural History) Geology, 10, 163237.Google Scholar
Butler, P. M. (1984). Macroscelidea, Insectivora and Chiroptera from the Miocene of East Africa. Palaeovertebrata, 14, 117200.Google Scholar
Butler, P. M. (2010). Neogene Insectivora. In Werdelin, L. & Sanders, W. J., eds. Cenozoic Mammals of Africa. Berkeley, CA: University of California Press, pp. 573580.Google Scholar
Butler, P. M. & Greenwood, M. (1973). The early Pleistocene hedgehog from Olduvai, Tanzania. Fossil Vertebrates of Africa, 3, 742.Google Scholar
Butler, P. M. & Greenwood, M. (1979). Soricidae (Mammalia) from the early Pleistocene of Olduvai Gorge, Tanzania. Zoological Journal of the Linnean Society, 67, 329379.Google Scholar
Butler, P. M. & Hopwood, A. T. (1957). Insectivora and Chiroptera from the Miocene rocks of Kenya Colony. Fossil Mammals of Africa, 13, 135.Google Scholar
Butler, P. M., Thorpe, R. S. & Greenwood, M. (1989). Interspecific relations of African crocidurine shrews (Mammalia: Soricidae) based on multivariate analysis of mandibular data. Zoological Journal of the Linnean Society, 96, 373412.Google Scholar
Butzer, K. W. (1973). On the geology of a late Pliocene Mammuthus site, Virginia, Orange Free State. Navorsinge van die Nasionale Museum Bloemfontein, 2, 386393.Google Scholar
Cable, J. H. C., Scott, K. & Carter, P. L. (1980). Excavations at Good Hope Shelter, Underberg District, Natal. Annals of the Natal Museum, 24, 134.Google Scholar
Cain, C. (2006). Human activity suggested by the taphonomy of 60 ka and 50 ka faunal remains from Sibudu Cave. Southern African Humanities, 18, 241260.Google Scholar
Cain, J. W., Krausman, P. R. & Germaine, H. L. (2004). Antidorcas marsupialis. Mammalian Species, 753, 17.Google Scholar
Capanna, E., Civitelli, M. V., Hickman, G. C. & Nevo, E. (1989). The chromosomes of Amblysomus hottentotus (Smith 1829) and A. iris Thomas; Schwann 1905: first report for the golden moles of Africa (Insectivora, Chrysochloridae). Tropical Zoology, 2, 318322.Google Scholar
Carlson, K. B., De Ruiter, D. J., DeWitt, T. J., et al. (2016). Developmental simulation of the adult cranial morphology of Australopithecus sediba. South African Journal of Science, 112, 19.Google Scholar
Castiglia, R., Solano, E., Makundi, R. H., et al. (2012). Rapid chromosomal evolution in the mesic four-striped grass rat Rhabdomys dilectus (Rodentia, Muridae) revealed by mtDNA phylogeographic analysis. Journal of Zoological Systematics and Evolutionary Research, 50, 165172.Google Scholar
Cavallini, P. (1992). Herpestes pulverulentus. Mammalian Species, 409, 14.Google Scholar
Chacornac, M. (1999). Etude de quelques Gerbillinae (Mammalia, Rodentia): données moléculaires et morphologiques. Unpublished MSc thesis, Université de Montpellier II.Google Scholar
Chatterton, J. F., Collett, D. P. & Swan, J. T. (1979). A Late Iron Age village site in the Letaba District, Northeast Transvaal. South African Archaeological Society Goodwin Series, 3, 109119.Google Scholar
Chevret, P., Denys, C., Jaeger, J.-J., Michaux, J. & Catzeflis, F. (1993a). Molecular evidence that the spiny mouse (Acomys) is more closely related to the gerbils (Gerbillinae) than to true mice (Murinae). Proceedings of the National Academy of Science of the United States of America, 90, 34333436.Google Scholar
Chevret, P., Denys, C., Jaeger, J.-J., Michaux, J. & Catzeflis, F. (1993b). Molecular and paleontological aspects of the tempo and mode of evolution in Otomys (Otomyinae: Muridae: Mammalia). Biochemical Systematics and Ecology, 21, 123131.Google Scholar
Chimimba, C. T. (1997). A systematic revision of southern African Aethomys Thomas, 1915 (Rodentia: Muridae). Unpublished PhD thesis, University of Pretoria.Google Scholar
Chimimba, C. T. (1998). A taxonomic synthesis of southern African Aethomys (Rodentia: Muridae) with a key to species. Mammalia, 62, 427437.Google Scholar
Chimimba, C. T. (2000). Geographic variation in Aethomys chrysophilus (Rodentia: Muridae) from southern Africa. Zeitschrift für Säugetierkunde, 65, 157171.Google Scholar
Chimimba, C. T. (2001). Infraspecific morphometric variation in Aethomys namaquensis (Rodentia: Muridae) from southern Africa. Journal of Zoology, 253, 191210.Google Scholar
Chimimba, C. T. (2005). Phylogenetic relationships in the genus Aethomys (Rodentia: Muridae). African Zoology, 40, 271284.Google Scholar
Chimimba, C. T. & Dippenaar, N. J. (1994). Non-geographic variation in Aethomys chrysophilus (De Winton, 1897) and A. namaquensis (A. Smith, 1834) (Rodentia: Muridae) from southern Africa. South African Journal of Zoology, 29, 107117.Google Scholar
Chimimba, C. T., Dippenaar, N. J. & Robinson, T. J. (1999). Morphometric and morphological delineation of southern African species of Aethomys (Rodentia: Muridae). Biological Journal of the Linnaean Society, 67, 501527.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. (1970). The fossil Equidae of the Krugersdorp caves. Annals of the Transvaal Museum, 26, 144168.Google Scholar
Churcher, C. S. (1974). Sivatherium maurisium (Pomel) from the Swartkrans australopithecine site, Transvaal (Mammalia: Giraffidae). Annals of the Transvaal Museum, 29, 6569.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. Palaeontologia 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. & Watson, V. (1993). 5. Additional fossil Equidae from Swartkrans. In Brain, C. K., ed. Swartkrans: a cave’s chronicle of early man. Transvaal Museum Monographs, 8, 137150.Google Scholar
Churchill, S. E., Pearson, O. M., Grine, F. E., Trinkaus, E. & Holliday, T. W. (1996). Morphological affinities of the proximal tibia from Klasies River main site: archaic or modern? Journal of Human Evolution, 31, 213237.Google Scholar
Churchill, S. E., Berger, L. R. & Parkington, J. E. (2000a). A middle Pleistocene human tibia from Hoedjiespunt, Western Cape, South Africa. South African Journal of Science, 96, 367368.Google Scholar
Churchill, S. E., Brink, J. S., Berger, L. R., et al. (2000b). 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
Clark, H. O. (2005). Otocyon megalotis. Mammalian Species, 766, 15.Google Scholar
Clark, J. L. & 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. (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: Aldine de Ruyter, pp. 285292.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. (2002). Newly revealed information on the Sterkfontein Member 2 Australopithecus skeleton. South African Journal of Science, 98, 523526.Google Scholar
Clarke, R. J. (2008). Latest information on Sterkfontein’s Australopithecus skeleton and a new look at Australopithecus. South African Journal of Science, 104, 443449.Google Scholar
Clarke, R. J. & Tobias, P. V. (1995). Sterkfontein Member 2 foot bones of the oldest South African hominid. Science, 269, 521524.Google Scholar
Clarke, R. J., Howell, F. C. & Brain, C. K. (1970). More evidence of an advanced hominid at Swartkrans. Nature, 225, 12191222.Google Scholar
Cohen, K. M., Finney, S. C., Gibbard, P. L. & Fan, J.-X. (2013, updated). The ICS International Chronostratigraphic Chart. Episodes, 36, 199204.Google Scholar
Collett, D. P. (1982). Excavations of stone-walled ruin types in the Badfontein Valley, Eastern Transvaal, South Africa. South African Archaeological Bulletin, 37, 3443.Google Scholar
Collings, G. E. (1972). A new species of machaerodont from Makapansgat. Palaeontologia Africana, 14, 8792.Google Scholar
Collings, G. E. (1973). Some new machaerodonts from Makapansgat Limeworks. Unpublished MSc thesis, University of the Witwatersrand.Google Scholar
Collings, G. E., Cruickshank, A. R. I., Maguire, J. M. & Randall, R. M. (1976). Recent faunal studies at Makapansgat Limeworks, Transvaal, South Africa. Annals of the South African Museum, 71, 153165.Google Scholar
Collins, B., Wilkins, J. & Ames, C. (2017). Revisiting the Holocene occupations at Grassridge Rockshelter, Eastern Cape, South Africa. South African Archaeological Bulletin, 72, 162170.Google Scholar
Commission on Zoological Nomenclature (2003). Opinion 2030 (Case 3178). Bulletin of Zoological Nomenclature, 60, 9091.Google Scholar
Conard, N. J. & Kandel, A. W. (2006). The economics and settlement dynamics of the later Holocene inhabitants of near coastal environments in the West Coast National Park (South Africa). In Wotzka, H.-P., ed. Grundlegungen. Beiträge zur europäischen und afrikanischen Archäologie für Manfred K. H. Eggert. Tübingen: Francke, pp. 329355.Google Scholar
Conroy, G. C., Pickford, M., Senut, B., Van Couvering, J. & Mein, P. (1992). Otavipithecus namibensis, first Miocene hominoid from southern Africa. Nature, 356, 144148.Google Scholar
Conroy, G. C., Pickford, M., Senut, B. & Mein, P. (1993). Additional Miocene primates from the Otavi Mountains, Namibia. Comptes Rendus de l’Académie des Sciences, Paris, Série II, 317, 987990.Google Scholar
Conroy, G. C., Senut, B., Gommery, D., Pickford, M. & Mein, P. (1996). Brief communication: new primate remains from the Miocene of Namibia southern Africa. American Journal of Physical Anthropology, 99, 487492.Google Scholar
Cooke, H. B. S. (1939). On a collection of fossil Mammalian remains from the Vaal River gravels at Pniel. South African Journal of Science, 36, 412416.Google Scholar
Cooke, H. B. S. (1947). Some fossil hippotragine antelopes from South Africa. South African Journal of Science, 43, 226231.Google Scholar
Cooke, H. B. S. (1949a). Fossil mammals of the Vaal River deposits. Memoirs of the Geological Society of South Africa, 35, 1108.Google Scholar
Cooke, H. B. S. (1949b). The fossil Suina of South Africa. Transactions of the Royal Society of South Africa, 32, 144.Google Scholar
Cooke, H. B. S. (1950). A critical revision of the Quaternary Perissodactyla of southern Africa. Annals of the South African Museum, 31, 393479.Google Scholar
Cooke, H. B. S. (1955). Some fossil mammals in the South African Museum collections. Annals of the South African Museum, 42, 161168.Google Scholar
Cooke, H. B. S. (1961). Further revision of the fossil Elephantidae of southern Africa. Palaeontologia Africana, 46–58.Google Scholar
Cooke, H. B. S. (1962). Appendix 1: notes on the faunal material from the Cave of Hearths and Kalkbank. In Mason, R., ed. Prehistory of the Transvaal. Johannesburg: University of the Witwatersrand Press, pp. 447453.Google Scholar
Cooke, H. B. S. (1974). The fossil mammals of the Cornelia Beds, O.F.S. Memoirs of the National Museum Bloemfontein, 9, 6384.Google Scholar
Cooke, H. B. S. (1985). Ictonyx bolti, a new mustelid from cave breccias at Bolt’s Farm, Sterkfontein area, South Africa. South African Journal of Science, 81, 618619.Google Scholar
Cooke, H. B. S. (1988). The larger mammals from the Cave of Hearths. In Mason, R., ed. Cave of Hearths, Makapansgat, Transvaal. Johannesburg: Archaeological Research Unit, pp. 507523.Google Scholar
Cooke, H. B. S. (1990). Taung fossils in the University of California collections. In Sperber, G. H., ed. Apes to Angels: Essays in Anthropology in Honour of Phillip V. Tobias. New York: Wiley-Liss, pp. 119134.Google Scholar
Cooke, H. B. S. (1991). Dinofelis barlowi (Mammalia, Carnivora, Felidae) cranial material from Bolt’s Farm, collected by the University of California African Expedition. Palaeontologia Africana, 28, 921.Google Scholar
Cooke, H. B. S. (1993a). Undescribed suid remains from Bolt’s Farm and other Transvaal cave deposits. Palaeontologia Africana, 30, 723.Google Scholar
Cooke, H. B. S. (1993b). Fossil proboscidean remains from Bolt’s Farm and other Transvaal cave brecchias. Palaeontologia Africana, 30, 2534.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. Palaeontologia Africana, 41, 131140.Google Scholar
Cooke, H. B. S. & Hendey, Q. B. (1992). Nyanzachoerus (Mammalia: Suidae: Tetraconodontinae) from Langebaanweg, South Africa. Durban Museum Novitates, 17, 120.Google Scholar
Cooke, H. B. S. & Wells, L. H. (1946). The Power collection of mammalian remains from the Vaal River deposits at Pneil. South African Journal of Science, 42, 224235.Google Scholar
Cooke, H. B. S. & Wells, L. H. (1951). Fossil remains from Chelmer, near Bulawayo, Southern Rhodesia. South African Journal of Science, 47, 205209.Google Scholar
Cooke, H. B. S., Malan, B. D. & Wells, L. H. (1945). Fossil man in the Lebombo Mountains, South Africa: the ‘Border Cave’, Ingwavuma District, Zululand. Man, 43, 613.Google Scholar
Coombs, M. C. & Cote, S. M. (2010). Chalicotheriidae. In Werdelin, L. & Sanders, W. J., eds. Cenozoic Mammals of Africa. Berkeley, CA: University of California Press, pp. 659667.Google Scholar
Coppens, Y. (1971). Une nouvelle espèce de Suidé du Villafranchien du Tunisie, Nyanzachoerus jaegeri nov. sp. Comptes Rendus Hebdomadaires des Séances, Académie des Sciences, 272, 32643267.Google Scholar
Corbet, G. B. & Hanks, J. (1968). A revision of the elephant shrews, family Macroscelididae. Bulletin of the British Museum (Natural History) Zoology, 16, 47111.Google Scholar
Corvinus, G. (1978). Palaeontological and archaeological investigations in the lower Orange valley from Arrisdrift to Orib, in the concession area of the Consolidated Diamond Mines of South West Africa (Proprietary Limited). Palaeoecology of Africa, 10, 7591.Google Scholar
Cote, S. M. (2010). Pecora Incertae Sedis. In Werdelin, L. & Sanders, W. J., eds. Cenozoic Mammals of Africa. Berkeley, CA: University of California Press, pp. 731739.Google Scholar
Cotterill, F. P. D. (2003). Insights into the taxonomy of tsessebe antelopes Damaliscus lunatus (Bovidae: Alcelaphini) with a description of a new evolutionary species in south-central Africa. Durban Museum Novitates, 28, 1130.Google Scholar
Crawford-Cabral, J. & Fernandes, A. C. (2001). The rusty-spotted genets as a group with three species in Southern Africa (Carnivora: Viverridae). In Denys, C., Granjon, L. & Poulet, A., eds. Proceedings of the 8th International Symposium on African Small Mammals. Paris: I.R.D., pp. 6580.Google Scholar
Cretzschmar, P. J. (1826). Atlas zu der Reise im nördlichen Afrika von Eduard Rüppell, 1. Zoologie, Säugethiere. Gedruckt und in Commission bei Heinr. Frankfurt am Main: Ludw. Brönner.Google Scholar
Croizet, J. B. & Jobert, A. C. G. (1828). Recherches sur les Ossemens Fossiles du Département du Puy-de-Dome. Paris: Delahays.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. (1981–1983). Faunal remains from some Middle and Later Stone Age sites in South West Africa. Journal of the South West Africa Scientific Society, 36–37, 91114.Google Scholar
Cruz-Uribe, K. & Schrire, C. (1991). Analysis of faunal remains from Oudepost I, an early outpost of the Dutch East India Company, Cape Province. South African Archaeological Bulletin, 46, 92106.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, 559575.Google Scholar
Curnoe, D. (2001). Early Homo from southern Africa: a cladistic perspective. South African Journal of Science, 97, 186190.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: Internationale Zeitschrift fur die vergleichende Forschung am Menschen, 61, 151177.Google Scholar
Curnoe, D. & Tobias, P. V. (2006). Description, reconstruction, comparative anatomy and classification of the Sterkfontein Stw 53 cranium, with discussions about the taxonomy of other southern African early Homo remains. Journal of Human Evolution, 50, 3677.Google Scholar
Curnoe, D., Herries, A., 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
Cuvier, F. (1821). Vervet. In Geoffroy-Saint-Hilaire, E. & Cuvier, F., eds. Histoire Naturelle des Mammifères, Vol. 5. Paris: Belin, 2 pp.Google Scholar
Cuvier, F. (1822). Du sanglier à masque et des Phacochoeres. Mémoires du Muséum d’Histoire Naturelle, 8, 447455.Google Scholar
Cuvier, F. (1823). [Lievre des rochers, Lepus saxatilis]. Dictionnaire des Sciences Naturelles, 26, 309310.Google Scholar
Cuvier, F. (1824). Vansire. In Geoffroy-Saint-Hilaire, E. & Cuvier, F., eds. Histoire Naturelle des Mammifères, Vol. 5. Paris: Belin, pl. XLVIII & 2 pp.Google Scholar
Cuvier, F. (1825). Des Dents des Mammifères Considérées comme Caractères Zoologiques. Strasbourg: Levrault.Google Scholar
Cuvier, F. (1826). [Phacochoeres, Phacochoerus]. Dictionnaire des Sciences Naturelles, 39, 383386.Google Scholar
Cuvier, F. (1829). Otomys cafre. In Geoffroy Saint-Hilaire, E. & Cuvier, F., eds. Histoire Naturelle des Mammifères Vol. 6. Paris: Belin, Pl. LX & pp. 2.Google Scholar
Cuvier, F. (1841). Mémoire sur les gerboises et les gerbilles. Transactions of the Zoological Society of London, 2, 131149.Google Scholar
Cuvier, G. (1798). Tableau Elémentaire de l’Histoire Naturelle des Animaux. Paris: Baudouin.Google Scholar
Cuvier, G. (1817). Le Règne Animal distribué d’après son Organisation, Vol. 1. Paris: Detervile.Google Scholar
Cuvier, G. (1829). Le Règne Animal distribué d’après son Organisation, Vol. 1. Nouvelle édn. Paris: Detervile.Google Scholar
Daams, R. & De Bruijn, H. (1995). A classification of the Gliridae (Rodentia) on the basis of dental morphology. Proc. II. Conf. on Dormice. Hystrix, 6, 350.Google Scholar
Dagg, A. I. (1971). Giraffa camelopardalis. Mammalian Species, 5, 18.Google Scholar
Dale, M. M. (1948). New fossil Suidae from Limeworks Quarry, Makapansgat, Potgietersrust. South African Science, 2, 114117.Google Scholar
Dalton, D. L., Linden, B., Wimberger, K., et al. (2015). New insights into Samango monkey speciation in South Africa. PLoS ONE, 10. DOI: 10.1371/journal.pone.0117003.Google Scholar
Dart, R. A. (1925). Australopithecus africanus: the man-ape of South Africa. Nature, 115, 195199.Google Scholar
Dart, R. A. (1927). Mammoths and man in the Transvaal. Nature, 120, 4148.Google Scholar
Dart, R. A. (1929a). A note on the Taungs skull. South African Journal of Science, 26, 648658.Google Scholar
Dart, R. A. (1929b). Mammoths and other fossil elephants of the Vaal and Limpopo watersheds. South African Journal of Science, 26, 698731.Google Scholar
Dart, R. A. (1940). Recent discoveries bearing on human history in southern Africa. Journal of the Royal Anthropological Institute of Great Britain and Ireland, 70, 1327.Google Scholar
Dart, R. A. (1948a). The Makapansgat proto-human Australopithecus prometheus. American Journal of Physical Anthropology, 6, 259284.Google Scholar
Dart, R. A. (1948b). The first human mandible from the Cave of Hearths, Makapansgat. South African Archaeological Bulletin, 3, 9698.Google Scholar
Dart, R. A. (1954). The adult female lower jaw from Makapansgat. American Anthropologist, NS, 56, 884888.Google Scholar
Dart, R. A. (1959). A tolerably complete australopithecine cranial from the Makapansgat Pink Breccia. South African Journal of Science, 55, 325327.Google Scholar
Davies, C. (1987). Fossil Pedetidae (Rodentia) from Laetoli. In Leakey, M. D. & Harris, J. M., eds. Laetoli: A Pliocene Site in Northern Tanzania. Oxford: Clarendon Press, pp. 171189.Google Scholar
Davis, D. H. S. (1949). The affinities of the South African gerbils of the genus Tatera. Proceedings of the Zoological Society of London, 118, 10021018.Google Scholar
Davis, D. H. S. (1961). Appendix B: report on the microfauna in the University of California collections from the South African cave breccias (excluding the Soricidae). Unpublished report, Transvaal Museum.Google Scholar
Davis, D. H. S. (1965). The affinities of the South African gerbils of the gerbils Tatera: corrections and notes. Proceedings of the Zoological Society of London, 144, 323326.Google Scholar
De Blainville, H. (1816). Sur plusieurs espèces d’animaux mammifères, de l’ordre des ruminans. Bulletin des Sciences par la Société Philomathique de Paris, 1816, 7282.Google Scholar
De Blainville, H. (1839). Sur l’Hyaenodon leptorhynchus (De Laizer) nouveau genre de Carnassiers fossiles d’Auvergne. Annales François et Etrange d’Anatomie et Physiologie, 3, 1731.Google Scholar
De Christol, J. (1832). [Description of Hipparion]. Annales des Sciences et de l’Industrie du Midi de la France, 1, 180181.Google Scholar
De Graaff, G. (1958). A new chrysochlorid from Makapansgat. Palaeontologia Africana, 5, 2127.Google Scholar
De Graaff, G. (1961a). A preliminary investigation of the Mammalian microfauna in Pleistocene deposits in the Transvaal System. Palaeontologia Africana, 7, 59118.Google Scholar
De Graaff, G. (1961b). A short survey of investigations of fossil rodents in African deposits. South African Journal of Science, 57, 191196.Google Scholar
De Graaff, G. (1961c). On the fossil mammalian microfauna collected at Kromdraai by Draper in 1895. South African Journal of Science, 57, 259260.Google Scholar
De Graaff, G. (1965). A Systematic Revision of the Bathyergidae (Rodentia) of Southern Africa. Unpublished PhD thesis, University of Pretoria.Google Scholar
De Graaff, G. (1988). The smaller mammals of the Cave of Hearths from basal guano underlying the Acheulean deposits (circa 200 000 BP). In Mason, R., ed. Cave of Hearths, Makapansgat, Transvaal. Johannesburg: Archaeological Research Unit, pp. 535548.Google Scholar
De Meneses Cabral, J. C. (1966). Note on the taxonomy of Genetta. Zoologica Africana, 2, 2526.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). Undescribed hominin fossils from the Transvaal Museum faunal collections. Annals of the Transvaal Museum, 41, 2940.Google Scholar
De Ruiter, D. J., Steininger, C. M. & Berger, L. R. (2006). A cranial base of Australopithecus robustus from the Hanging Remnant of Swartkrans, South Africa. American Journal of Physical Anthropology, 130, 435444.Google Scholar
De Ruiter, D. J., Brophy, J. K., Lewis, P. J., Churchill, S. E. & Berger, L. R. (2008a). 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., Sponheimer, M. & Lee Thorp, J. A. (2008b). Indications of habitat association of Australopithecus robustus in the Bloubank Valley, South Africa. Journal of Human Evolution, 55, 10151030.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., Brophy, J. K., Lewis, P. J., et al. (2010). Preliminary investigation of Matjhabeng, a Pliocene fossil locality in the Free State of South Africa. Palaeontologia Africana, 45, 1122.Google Scholar
De Ruiter, D. J., Churchill, S. E., Brophy, J. K. & Berger, L. R. (2011). Regional survey of Middle Stone Age fossil vertebrate deposits in the Virginia-Theunissen area of the Free State, South Africa. Navorsinge van die Nasionale Museum Bloemfontein, 27, 120.Google Scholar
De Ruiter, D. J., DeWitt, T. J., Carlson, K. B., et al. (2013). Mandibular remains support taxonomic validity of Australopithecus sediba. Science, 340. DOI: 10.1126/science.1232997.Google Scholar
De Villiers, H. (1965). Part II: skeletal remains from the Gamtoos Valley. South African Archaeological Bulletin, 20, 201205.Google Scholar
De Villiers, H. (1972a). The first fossil human skeleton from South West Africa. Transactions of the Royal Society of South Africa, 40, 187196.Google Scholar
De Villiers, H. (1972b). Appendix 1: the Welgegund human skeleton – physical description. In Voigt, E. A.. Preliminary report on Welgegund: an Iron Age burial site. South African Archaeological Bulletin, 27, 163.Google Scholar
De Villiers, H. (1973). Human skeletal remains from Border Cave, Ingwavuma District, KwaZulu, South Africa. Annals of the Transvaal Museum, 28, 229246.Google Scholar
De Villiers, H. (1974). Human skeletal remains from Cape St Francis, Cape Province. South African Archaeological Bulletin, 29, 8991.Google Scholar
De Villiers, H. (1976a). A second adult human mandible from Border Cave, Ingwavuma District, KwaZulu, South Africa. South African Journal of Science, 72, 121125.Google Scholar
De Villiers, H. (1976b). Human skeletal remains from Tautswemogala Hill, Botswana. Botswana Notes and Records, 8, 724.Google Scholar
De Villiers, H. (1980). Appendix: human skeletal remains from Iron Age burials in the Limpopo/Shashi Valley. In Hanisch, E. O. M.. An archaeological interpretation of certain Iron Age sites in the Limpopo/Shashi Valley. Unpublished MA thesis, University of Pretoria, pp. 1–20.Google Scholar
De Villiers, H. (1982). Appendix II: report on human skeletal remains: 2539 AB4, JVIIm5. In Evers, T. M.. Excavations at the Lydenburg Heads site, Eastern Transvaal, South Africa. South African Archaeological Bulletin, 37, 33.Google Scholar
De Villiers, H. (1984). Appendix: Early Iron Age human skeletal remains from Mhlopeni and Msuluzi Confluence. In Maggs, T. & Ward, V.. Early Iron Age sites in the Muden area of Natal. Annals of the Natal Museum, 26, 138140.Google Scholar
De Villiers, H. (1987). Appendix 1: report on human skeletal remains from Matanga (07-D3-9), Botswana. In Van Waarden, C.. Matanga, a late Zimbabwe cattle post. South African Archaeological Bulletin, 42, 123124.Google Scholar
De Villiers, H. & Wilson, M. L. (1982). Human burials from Byneskranskop, Bredasdorp District, Cape Province, South Africa. Annals of the South African Museum, 88, 205248.Google Scholar
De Wet-Bronner, E. (1994). The faunal remains from four Late Iron Age sites in the Soutpansberg region: Part 1 – Tavhatshena. Southern African Field Archaeology, 3, 3343.Google Scholar
De Wet-Bronner, E. (1995a). The faunal remains from four Late Iron Age sites in the Soutpansberg region: Part II – Tshitheme and Dzata. Southern African Field Archaeology, 4, 1829.Google Scholar
De Wet-Bronner, E. (1995b). The faunal remains from four Late Iron Age sites in the Soutpansberg region: Part III – Tshirululuni. Southern African Field Archaeology, 4, 109119.Google Scholar
De Winton, W. E. (1896). On collections of rodents made by Mr J. ffolliot Darling in Mashunaland and Mr F.C. Selous in Matabeleland with short field notes by the collectors. Proceedings of the Zoological Society of London, 1896, 798809.Google Scholar
De Winton, W. E. (1898). On the nomenclature and distribution of some of the rodents of South Africa, with descriptions of new species. Annals and Magazine of Natural History, Series 7, 2, 18.Google Scholar
Deacon, H. J. (1976). Where Hunters Gathered. Claremont Cape: South African Archaeological Society.Google Scholar
Dechow, P. C. & Singer, R. (1984). Additional fossil Theropithecus from Hopefield, South Africa: a comparison with other African sites and a re-evaluation of its taxonomic status. American Journal of Physical Anthropology, 63, 405435.Google Scholar
Dembo, M., Radovčić, D., Garvin, H. M., et al. (2016). The evolutionary relationships and age of Homo naledi: an assessment using dated Bayesian phylogenetic methods. Journal of Human Evolution, 97, 1726.Google Scholar
Dempster, E. R., Perrin, M. R. & Downs, C. T. (1999). Gerbillurus vallinus. Mammalian Species, 605, 14.Google Scholar
Denbow, J. (2011). Excavations at Divuyu, Tsodilo Hills. Botswana Notes and Records, 43, 7694.Google Scholar
Denbow, J., Smith, J. D. B., Ndobochani, N. M., Atwood, K. & Miller, D. (2008). Archaeological excavations at Bosutswe, Botswana: cultural chronology, paleo-ecology and economy. Journal of Archaeological Science, 35, 459480.Google Scholar
Denys, C. (1988a). Nouvelles observations de la structure dentaire de spécimens juvéniles de Cryptomys hottentotus (Rongeurs, Batherygidés). Mammalia, 52, 292294.Google Scholar
Denys, C. (1988b). Apports de l’analyse morphologiques à la determination des espèces actuelles et fossiles du genre Saccostomus (Cricetomyinae, Rodentia). Mammalia, 52, 497532.Google Scholar
Denys, C. (1990a). The oldest Acomys (Rodentia, Muridae) from the Lower Pliocene of South Africa and the problem of its murid affinities. Palaeontographica Abt A, 210, 7991.Google Scholar
Denys, C. (1990b). Deux nouvelles espèces d’Aethomys (Rodentia, Muridae) à Langebaanweg (Pliocène, Afrique du Sud): implications phylogénétiques. Annales de Paléontologie, 76, 4169.Google Scholar
Denys, C. (1991). Un nouveau rongeur Mystromys pocockei sp. nov. (Cricetinae) du Pliocène inférieur de Langebaanweg (Région du Cap, Afrique du Sud). Comptes Rendus de l’Académie des Sciences de Paris, Série IIa, 313, 13351341.Google Scholar
Denys, C. (1994a). Nouvelles espèces de Dendromus (Rongeurs, Muroidea) à Langebaanweg (Pliocène, Afrique du Sud). Conséquences stratigraphiques et paléoécologiques. Palaeovertebrata, 23, 153176.Google Scholar
Denys, C. (1994b). Affinités systématiques de Stenodontomys (Mammalia, Rodentia) rongeur Muroidea du Pliocène de Langebaanweg (Afrique du Sud). Comptes Rendus de l’Académie des Sciences de Paris, Série IIa, 318, 411416.Google Scholar
Denys, C. (1998). Phylogenetic implications of the existence of two modern genera of Bathyergidae (Mammalia, Rodentia) in the Pliocene site of Langebaanweg (South Africa). Annals of the South African Museum, 105, 265286.Google Scholar
Denys, C. (1999). Of mice and men. In Bromage, T. G. & Schrenk, F., eds. African Biogeography, Climate Change and Human Evolution. New York: Oxford University Press, pp. 226257.Google Scholar
Denys, C. & Jaeger, J. J. (1992). Rodents of the Miocene site of Fort Ternan (Kenya): first part – phiomyids, bathyergids, sciurids and anomalurids. Neues Jahrbuch fűr Geologie und Paläontologie, Abhandlungen, 185, 6384.Google Scholar
Denys, C. & Matthews, T. (2017). A new Desmodillus (Gerbillinae, Rodentia) species from the early Pliocene site of Langebaanweg (South-western Cape, South Africa). Palaeovertebrata, 41. DOI: 10.18563/pv.41.1.e1.Google Scholar
Denys, C., Michaux, J. & Hendey, B. (1987). Les rongeurs (Mammalia) Euryotomys et Otomys: un exemple d’évolution parallèle en Afrique tropicale. Comptes Rendus de l’Academie des Sciences de Paris, Série IIa, 305, 13891395.Google Scholar
Depéret, C. (1897). Découverte du Mastodon angustidens dans l’étage cartennien de Kabylie. Bulletin de la Société Géologique de France, Series 3, 25, 518521.Google Scholar
Deraniyagala, P. E. P. (1951). A hornless rhinoceros from the Mio-Pliocene deposits of East Africa. Spolia Zeylanica, 26, 133135.Google Scholar
DeSilva, J. M., Steininger, C. & Patel, B. A. (2013). Cercopithecoid primate postcranial fossils from Cooper’s D, South Africa. Geobios, 46, 381394.Google Scholar
Desmarest, A. G. (1822). Mammalogie ou Descriptions des Espèces de Mammifères. Paris: Mme Veuve Agasse.Google Scholar
Dewar, G. I. (2007). The archaeology of the coastal desert of Namaqualand, South Africa: a regional synthesis. Unpublished PhD thesis, University of Cape Town.Google Scholar
Dewar, G. & Jerardino, A. (2007). Micromammals: when humans are the hunters. Journal of Taphonomy, 5, 114.Google Scholar
Dewar, G. & Stewart, B. A. (2012). Preliminary results of excavations at Spitzkloof Rockshelter, Richtersveld, South Africa. Quaternary International, 270, 3039.Google Scholar
Dewar, G. & Stewart, B. A. (2016). Paleoenvironments, sea levels and land use in Namaqualand, South Africa, during MIS6–2. In Jones, S. C. & Stewart, B. A., eds. Africa from MIS 6-2: Population Dynamics and Paleoenvironments. Dordrecht: Springer Science+Business Media, pp. 195212.Google Scholar
Dewar, G. & Stewart, B. A. (2017). Early Maritime desert dwellers in Namaqualand, South Africa: a Holocene perspective on Pleistocene peopling. Journal of Island & Coastal Archaeology, 12, 4464.Google Scholar
Dewar, G., Halkett, D., Hart, T., Orton, J. & Sealy, J. (2006). Implications of a mass kill site of springbok (Antidorcas marsupialis) in South Africa: hunting practices, gender relations and sharing in the Later Stone Age. Journal of Archaeological Science, 33, 12661275.Google Scholar
Dietrich, W. O. (1915). Elephas antiquus recki n.f. aus dem Diluvium Deutch-Ostafrikas. In Reck, H., ed. Wissenschaftliche Ergebnisse, Oldoway Expedition. Leipzig: Boerntraeger, pp. 180.Google Scholar
Dietrich, W. O. (1926). Fortschritte der Säugetierpaläontologie Afrikas. Forschungen und Fortschritte, 15, 121122.Google Scholar
Dietrich, W. O. (1928). Pleistocäne Deutsch-Ostafrikaische Hippopotamus-reste. In Reck, H., ed. Wissenschaftliche Ergeebnisse des Oldoway Expedition herausgeben von Prof. Dr. Reck. Neue Folge, Heft 3. Leipzig: Boerntraeger, pp. 241.Google Scholar
Dietrich, W. O. (1942). Altestquartare Saugetiere aus der sudlichen Serengeti, Deutsch-Ostafrika. Palaeontographica (A), 94, 43133.Google Scholar
Dineur, H. (1982). Le genre Brachyodus, Anthracotheriidae (Artiodactyla, Mammalia) du Miocene inférieur d’Europe et d’Afrique. Mémoires des Sciences de la Terre, Université de Paris, 6, 1186.Google Scholar
Dippenaar, N. J. (1977). Variation in Crocidura mariquensis (A. Smith, 1844) in southern Africa. Part 1 (Mammalia: Soricidae). Annals of the Transvaal Museum, 30, 163206.Google Scholar
Dippenaar, N. J. (1979). Variation in Crocidura mariquensis (A. Smith, 1844) in southern Africa, Part 2 (Mammalia: Soricidae). Annals of the Transvaal Museum, 32, 134.Google Scholar
Dippenaar, N. J. & Rautenbach, I. L. (1986). Morphometrics and karyology of the southern African species of the genus Acomys Geoffroy Saint-Hilaire, 1838 (Rodentia: Muridae). Annals of the Transvaal Museum, 34, 129183.Google Scholar
Dippenaar, N. J., Swanepoel, P. & Gordon, D. H. (1993). Diagnostic morphometrics of two medically important southern African rodents, Mastomys natalensis and M. coucha (Rodentia: Muridae). South African Journal of Science, 89, 300303.Google Scholar
Dirks, P. H. G., Kibii, J. M., Kuhn, B. F., et al. (2010). Geological age and setting of Australopithecus sediba from southern Africa. Science, 328, 205208.Google Scholar
Discamps, E. & Henshilwood, C. S. (2015). Intra-site variability in the Still Bay fauna at Blombos Cave: implications for explanatory models of the Middle Stone Age cultural and technological evolution. PLoS ONE, 10, 121.Google Scholar
Drennan, M. R. (1937). Archaeology of the Oakhurst Shelter, George: Part III. The cave-dwellers. Transactions of the Royal Society of South Africa, 25, 259280.Google Scholar
Drennan, M. R. (1953). The Saldanha skull and its associations. Nature, 172, 791793.Google Scholar
Drennan, M. R. (1955). The special features and status of the Saldanha skull. American Journal of Physical Anthropology, 13, 625634.Google Scholar
Dreyer, T. F. (1933). The archaeology of the Matjies River Rock Shelter. Transactions of the Royal Society of South Africa, 21, 187209.Google Scholar
Dreyer, T. F. (1935). A human skull from Florisbad. Proceedings of the Academy of Sciences, Amsterdam, 38, 119128.Google Scholar
Dreyer, T. F. & Lyle, A. (1931). New Fossil Mammals and Man from South Africa. Bloemfontein: Nasionale Pers.Google Scholar
Dubois, E. (1894). Pithecanthropus erectus: ein meschenähnliche Übergangsform aus Java. Batavia: Landesdruckerei.Google Scholar
Dusseldorp, G., Lombard, M. & Wurz, S. (2013). Pleistocene Homo and the updated Stone Age sequence of South Africa. South African Journal of Science, 109. DOI: 10.1590/sajs.2013/20120042.Google Scholar
Duvernoy, G. L. (1838). Supplément au mémoire sur les musaraignes. Mémoire de la Société d’ Histoire Naturelle de Strasbourg, 2, 17.Google Scholar
Duvernoy, G. L. (1851). Note sur une espèce de Buffle fossile [Bubalus (Arni) antiquus], découverte en Algérie. Comptes Rendus Hebdomadaires des Séances de l’Académie des Sciences, 33, 595597.Google Scholar
Edwards, S. (2009). Phylogeographic variation of the Karoo bush rat, Otomys unisulcatus: a molecular and morphological perspective. Unpublished MSc thesis, University of Stellenbosch.Google Scholar
Edwards, S., Claude, J., Van Vuuren, B. J. & Matthee, C. A. (2011). Evolutionary history of the Karoo bush rat, Myotomys unisulcatus (Rodentia: Muridae): disconcordance between morphology and genetics. Biological Journal of the Linnean Society, 102, 510526.Google Scholar
Éhik, J. (1930). Prodinotherium hungaricum n. g., n. sp. Geologica Hungarica Series Palaeontologica, 6, 124.Google Scholar
Ehrenberg, C. G. (1832). Herpestes leucurus H. et E. Colloraria. In Hemprich, F. W. & Ehrenberg, C. G.. Symbolae Physicae, seu, Icones et Descriptiones Corporum Naturalium Novorum aut Minus Cognitorum Zoologica 1, Mammalium 2 . Berlin: Officina Academica.Google Scholar
Ehrenberg, C. G. (1833). Sciurus, Xerus, bracchyotus H. et E. In Hemprich, F. W. & Ehrenberg, C. G.. Symbolae Physicae, seu, Icones et Descriptiones Corporum Naturalium Novorum aut Minus Cognitorum Zoologica 1, Mammalium 1. Berlin: Officina Academica.Google Scholar
Eisenmann, V. (2000). Equus capensis (Mammalia, Perissodactyla) from Elandsfontein. Palaeontologia Africana, 36, 9196.Google Scholar
Eisenmann, V. & Baylac, M. (2000). Extant and fossil Equus (Mammalia, Perissodactyla) skulls: a morphometric definition of the subgenus Equus. Zoologica Scripta, 29, 89100.Google Scholar
Eisenmann, V. & Brink, J. S. (2000). Koffiefontein quaggas and true Cape quaggas: the importance of basic skull morphology. South African Journal of Science, 96, 529533.Google Scholar
Ellerman, J. R. (1941). The Families and Genera of Living Rodents. London: British Museum (Natural History).Google Scholar
Ellerman, J. R., Morrison-Scott, T. C. S. & Hayman, R. W. (1953). Southern African Mammals 1758–1951: A Reclassification. London: British Museum (Natural History).Google Scholar
Engelbrecht, A., Taylor, P. J., Daniels, S. R. & Rambau, R. V. (2011). Chromosomal polymorphisms in African Vlei Rats, Otomys irroratus (Muridae: Otomyini), detected by banding techniques and chromosome painting: inversions, centromeric shifts and diploid number variation. Cytogenetic and Genome Research, 133, 815.Google Scholar
Ennouchi, E. (1953). Un nouveau genre d’ovicapriné dans le gisement pléistocène de Rabat. Compte Rendu Sommaire des Séances de la Société Géologique de France, 8, 126128.Google Scholar
Erxleben, J. C. P. (1777). Systema Regni Animales per Classes, Ordines, Genera, Species, Varietates cum Synomia et Historia Animalium. Lipsiae [Leipzig]: Impensis Weygandianis.Google Scholar
Esterhuysen, A. B., Behrens, J. & Harper, P. (1994). Leliehoek Shelter: a Holocene sequence from the eastern Orange Free State. South African Archaeological Bulletin, 49, 7378.Google Scholar
Evans, F. G. (1942). The osteology and relationships of the elephant shrews (Macroscelididae). Bulletin of the American Museum of Natural History, 80, 85125.Google Scholar
Evers, T. M. (1980). Klingbeil Early Iron Age sites, Lydenburg, Eastern Transvaal, South Africa. South African Archaeological Bulletin, 35, 4657.Google Scholar
Ewer, R. F. (1954). The fossil carnivores of the Transvaal caves: the Hyaenidae of Kromdraai. Proceedings of the Zoological Society of London, 124, 565585.Google Scholar
Ewer, R. F. (1955a). The fossil carnivores of the Transvaal caves: the Hyaenidae, other than Lycyaena, of Swartkrans and Sterkfontein. Proceedings of the Zoological Society of London, 124, 815837.Google Scholar
Ewer, R. F. (1955b). The fossil carnivores of the Transvaal caves: the Lycyaenas of Sterkfontein and Swartkrans, together with some general considerations of the Transvaal fossil hyaenids. Proceedings of the Zoological Society of London, 124, 839857.Google Scholar
Ewer, R. F. (1955c). The fossil carnivores of the Transvaal caves: Machairodontinae. Proceedings of the Zoological Society of London, 125, 587615.Google Scholar
Ewer, R. F. (1956a). The fossil carnivores of the Transvaal caves: two new viverrids, together with some general considerations. Proceedings of the Zoological Society of London, 126, 259274.Google Scholar
Ewer, R. F. (1956b). The fossil suids of the Transvaal caves. Proceedings of the Zoological Society of London, 127, 527544.Google Scholar
Ewer, R. F. (1957a). Some fossil carnivores from the Makapansgat Valley. Palaeontologia Africana, 4, 5767.Google Scholar
Ewer, R. F. (1957b). The fossil pigs of Florisbad. Navorsinge van die Nasionale Museum Bloemfontein, 1, 239257.Google Scholar
Ewer, R. F. (1958a). A collection of Phacochoerus aethiopicus teeth from the Kalkbank Middle Stone Age site, central Transvaal. Palaeontologia Africana, 5, 520.Google Scholar
Ewer, R. F. (1958b). Appendix A: faunal lists for the sites of Sterkfontein, Swartkrans, Kromdraai A and Makapansgat Limeworks. In Brain, C. K., ed. The Transvaal ape-man-bearing cave deposits. Transvaal Museum Memoir, 11, 128130.Google Scholar
Ewer, R. F. (1958c). A note on some South African fossil otters. Navorsinge van die Nasionale Museum Bloemfontein, 1, 275280.Google Scholar
Ewer, R. F. (1958d). The fossil Suidae of Makapansgat. Proceedings of the Zoological Society, 130, 329372.Google Scholar
Ewer, R. F. (1962). Appendix 2: Kalkbank Suidae. In Mason, R. J., ed. Prehistory of the Transvaal. Johannesburg: Witwatersrand University Press, p. 454.Google Scholar
Ewer, R. F. & Singer, R. (1956). Fossil Carnivora from Hopefield. Annals of the South African Museum, 42, 335342.Google Scholar
Fagan, B. M. (1960). The Glentyre shelter and Oakhurst re-examined. South African Archaeological Bulletin, 15, 8094.Google Scholar
Fagan, B. M. (1966). Appendix: Vertebrate fauna from Harleigh Farm. In Robins, P. A. & Whitty, A.. Excavations at Harleigh Farm, near Rusape, Rhodesia, 1958–1962. South African Archaeological Bulletin, 21, 7880.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., 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
Falconer, H. & Cautley, P. T. (1836). Sivatherium giganteum, a new fossil ruminant gnu, from the Valley of the Markanda, in the Sivalik branch of the Sub-Himalayan Mountains. Journal of the Asiatic Society of Bengal, 5, 3850.Google Scholar
Faulkes, C. G., Bennett, N. C., Bruford, M. W., et al. (1997). Ecological constraints driving social evolution in the African mole-rat. Proceedings of the Royal Society of London B: Biological Sciences, 264, 16191627.Google Scholar
Faulkes, C. G., Verheyen, E., Verheyen, W., Jarvis, J. U. M. & Bennett, N. C. (2004). Phylogeographical patterns of genetic divergence and speciation in African mole-rats (Family: Bathyergidae). Molecular Ecology, 13, 613629.Google Scholar
Feakins, S. J. & DeMenocal, P. (2010). Global and African climate during the Cenozoic. In Werdelin, L. & Sanders, W. J., eds. Cenozoic Mammals of Africa. Berkeley, CA: University of California Press, pp. 4555.Google Scholar
Fernandez-Jalvo, Y. & Avery, D. M. (2015). Pleistocene micromammals and their predators at Wonderwerk Cave, South Africa. African Archaeological Review, 32, 751791.Google Scholar
Ficcarelli, G., Torre, D. & Turner, A. (1984). First evidence of a species of raccoon dog, Nyctereutes Temminck, 1838, in South African Plio-Pleistocene deposits. Bolletino della Società Palaeontologica Italiana, 23, 125130.Google Scholar
Fitzinger, L. J. (1867). Versuch einer natűrlichen Anordnung der Nagthiere (Rodentia). Sitzungsberichte der Akademie der Wissenschafliche Wein, 56, 57168.Google Scholar
Flynn, L. J. & Sabatier, M. (1984). A muroid rodent of Asian affinity from the Miocene of Kenya. Journal of Paleontology, 3, 160165.Google Scholar
Forssman, T. (2014). Dzombo Shelter: a contribution to the Later Stone Age sequence of the Greater Mapungubwe Landscape. South African Archaeological Bulletin, 69, 182191.Google Scholar
Forster, J. R. (1778). Beskrifning på Djuret Yerbua Capensis, med Anmårkningar (*) om Genus Yerbuae. Kungl. Svenska Vetenskaps Akademiens Handlingar Series 1, 39, 108119.Google Scholar
Forster, J. R. (1790). Le Vaillant’s Erste Reise in das Innere von Afrika, während der Jahre 1780 bis 1782, Bd 1(v 3). Berlin: Fleischer.Google Scholar
Fourtau, R. (1918). Contribution à l’Etude des Vertébrés Miocènes de l’Egypte. Cairo: Geological Survey of Egypt.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, 366377.Google Scholar
Fraas, E. (1907). Pleistocäne Fauna aus den Diamantseifen von Südafrika. Zeitschrift der Deutschen Geologischen Gesellschaft, 59, 232243.Google Scholar
Franz-Odendaal, T. A. & Salounias, N. (2004). Comparative dietary evaluations of an extinct giraffid (Sivatherium hendeyi) (Mammalia, Giraffidae, Sivatheriinae) from Langebaanweg, South Africa (early Pliocene). Geodiversitas, 26, 675685.Google Scholar
Franz-Odendaal, T., Kaiser, T. M. & Bernor, R. L. (2003). Systematics and dietary evaluation of a fossil equid from South Africa. South African Journal of Science, 99, 453459.Google Scholar
Freedman, L. (1954). The status of Papio rhodesiae (Haagner) 1918. Annals of the Transvaal Museum, 22, 267270.Google Scholar
Freedman, L. (1957). The fossil Cercopithecoidea of South Africa. Annals of the Transvaal Museum, 23, 8262.Google Scholar
Freedman, L. (1961). Some new fossil cercopithecoid specimens from Makapansgat, South Africa. Palaeontologia Africana, 7, 745.Google Scholar
Freedman, L. (1965). Fossil and subfossil primates from the limestone deposits at Taungs, Bolt’s Farm and Witkrans, South Africa. Palaeontologia Africana, 9, 1948.Google Scholar
Freedman, L. (1970). A new checklist of fossil Cercopithecoidea of South Africa. Palaeontologia Africana, 13, 109110.Google Scholar
Freedman, L. (1976). South African fossil Cercopithecoidea: a re-assessment including a description of new material from Makapansgat, Sterkfontein and Taung. Journal of Human Evolution, 5, 297315.Google Scholar
Freedman, L. & Brain, C. K. (1972). Fossil cercopithecoid remains from the Kromdraai australopithecine site (Mammalia: Primates). Annals of the Transvaal Museum, 28, 116.Google Scholar
Freedman, L. & Brain, C. K. (1977). A re-examination of the cercopithecoid fossils from Swartkrans (Mammalia: Cercopithecidae). Annals of the Transvaal Museum, 30, 211218.Google Scholar
Freedman, L. & Stenhouse, N. S. (1972). The Parapapio species of Sterkfontein, Transvaal, South Africa. Palaeontologia Africana, 14, 93111.Google Scholar
Frisch, J. L. (1775). Das Natur-System der Vierfüssigen Thiere in Tabellen darinnen alle Ordnungen, Geschlechte und Arten, nicht nur mit bestimmenden Benennungen, sondern beygesetzten unterschiedenden Kennzeichen angezeigt werden, zum Nutzen der erwachsenen Schuljugend. Glogau: Günther.Google Scholar
Frost, S., Saanane, C., Starkovich, B., et al. (2017). New cranium of the large cercopithecid primate Theropithecus oswaldi leakeyi (Hopwood, 1934) from the paleoanthropological site of Makuyuni, Tanzania. Journal of Human Evolution, 109, 4556.Google Scholar
Galloway, A. (1936). Some prehistoric skeletal material from the Natal coast. Transactions of the Royal Society of South Africa, 23, 277295.Google Scholar
Galloway, A. (1937a). Man in Africa in the light of recent discoveries. South African Journal of Science, 34, 89120.Google Scholar
Galloway, A. (1937b). The characteristics of the skull of the Boskop physical type. American Journal of Physical Anthropology, 32, 3147.Google Scholar
Galloway, A. & Wells, L. H. (1934). Report on the human skeletal remains from the Karridene site. Transactions of the Royal Society of South Africa, 22, 225233.Google Scholar
Gaubert, P., Tranier, M., Veron, G., et al. (2003a). Case 3204: Viverra maculata Gray, 1830 (currently Genetta maculata; Mammalia, Carnivora) proposed conservation of the specific name. Bulletin of Zoological Nomenclature, 60, 4547.Google Scholar
Gaubert, P., Tranier, M., Veron, G., et al. (2003b). Nomenclatural comments on the rusty-spotted genet (Carnivora, Viverridae) and designation of a neotype. Zootaxa, 160, 114.Google Scholar
Gaubert, P., Taylor, P. J., Fernandes, C. A., Bruford, M. W. & Veron, G. (2005). Patterns of cryptic hybridization revealed using an integrative approach: a case study on genets (Carnivora, Viverridae, Genetta spp.) from the southern African subregion. Biological Journal of the Linnaean Society, 86, 1133.Google Scholar
Gaudin, T. (2010). Pholidota. In Werdelin, L. & Sanders, W. J., eds. Cenozoic Mammals of Africa. Berkeley, CA: University of California Press, pp. 599602.Google Scholar
Gaudin, T. J., Emry, R. J. & Wible, J. R. (2009). The phylogeny of living and extinct pangolins (Mammalia, Pholidota) and associated taxa: a morphology based analysis. Journal of Mammalogy, 16, 235305.Google Scholar
Gaudry, A. (1863). Animaux Fossiles et Géologie de l’Attique. Paris: Savy.Google Scholar
Gautier, A. & Van Waarden, C. (1981). The subsistence patterns at the Leeukop site, eastern Tuli Block. Botswana Notes and Records, 13, 111.Google Scholar
Gear, J. H. S. (1926). A preliminary account of the baboon remains from Taungs. South African Journal of Science, 23, 731747.Google Scholar
Gentry, A. W. (1965). New evidence on the systematic position of Hippotragus niro Hopwood, 1936 (Mammalia). Journal of Natural History, Series 13, 8, 335338.Google Scholar
Gentry, A. W. (1970). Revised classification for Makapania broomi Wells and Cooke (Bovidae, Mammalia). Palaeontologia Africana, 13, 6367.Google Scholar
Gentry, A. W. (1974). A new genus and species of the Pliocene boselaphine (Bovidae, Mammalia) from South Africa. Annals of the South African Museum, 65, 145188.Google Scholar
Gentry, A. W. (1980). Fossil Bovidae (Mammalia) from Langebaanweg, South Africa. Annals of the South African Museum, 79, 213337.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, L. & Sanders, W. J., eds. Cenozoic Mammals of Africa. Berkeley, CA: University of California Press, pp. 741796.Google Scholar
Gentry, A. W. & Gentry, A. (1978). Fossil Bovidae (Mammalia) of Olduvai Gorge, Tanzania: Part I. Bulletin of the British Museum (Natural History) Geology, 29, 289446.Google Scholar
Gentry, A. W., Gentry, A. & Mayr, H. (1995). Rediscovery of fossil antelope holotypes (Mammalia, Bovidae) collected from Olduvai Gorge, Tanzania, in 1913. Mitteilungen aus dem Bayerischen Staatssammlung für Paläontologie und historische Geologie, 35, 125135.Google Scholar
Geoffroy Saint-Hilaire, E. (1796). Mammifères : Mémoire sur les rapports naturels des Makis Lemur. Magasin Encyclopedique, 1, 2050.Google Scholar
Geoffroy Saint-Hilaire, E. (1803). Catalogue des Mammifères du Muséum National d’Histoire Naturelle. Paris: [publisher not given].Google Scholar
Geoffroy Saint-Hilaire, E. (1810). Description des roussettes et des cephalotes, deux nouveaux genres de la famille des chauve-souris. Annales du Muséum d’Histoire Naturelle, 15, 86108.Google Scholar
Geoffroy Saint-Hilaire, E. (1812). Tableau des quadrumanes, ou des animaux composant le premier Order de la Classe des Mammifères. Annales du Muséum d’Histoire Naturelle, 19, 85170.Google Scholar
Geoffroy Saint-Hilaire, E. (1813). Sur un genre de chauve-souris, sous le nom de rhinolopes (1). Annales du Muséum d’Histoire Naturelle Paris, 20, 254266.Google Scholar
Geoffroy Saint-Hilaire, E. (1818). Des chauve-souris. In Description des Mammifères qui se trouvent en Egypte. Déscription de l’Egypte, Vol. 2 Histoire Naturelle. Paris: Imprimerie Imperiale, pp. 99144.Google Scholar
Geoffroy Saint-Hilaire, E. & Cuvier, F. (1795). Mémoire sur une nouvelle division des Mammifères, et sur les principes qui doivent servis de base dans cette sorte de travail. Magasin Encyclopedique, 1, 164190.Google Scholar
Geoffroy Saint-Hilaire, I. (1827). Musaraigne. Dictionnaire Classique d’Histoire Naturelle, 11, 313329.Google Scholar
Geoffroy Saint-Hilaire, I. (1832). II. Déscription de trois espèces du genre Lièvre. Magasin de Zoologie, 2, cl. 1, pl. 9–10.Google Scholar
Geoffroy Saint-Hilaire, I. (1837). Notice sur deux nouveaux genres de Mammifères carnassiers, les Ichneumies, du continent africain, et les Galidies, de Madagascar. Annales des Sciences Naturelles (Zoologie) Paris, Series 2, 8, 249252.Google Scholar
Geoffroy Saint-Hilaire, I. (1838). Notice sur les rongeurs épineux désignés par les auteurs sous les noms d’Echumys, Lonchères, Heteromys et Nelomys. Annales des Sciences Naturelles (Zoologie) Paris, Series 2, 10, 122127.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. (1992). Phylogenetic analysis of the tribe Bovini (Mammalia: Artiodactyla). Zoological Journal of the Linnean Society, 104, 193207.Google Scholar
Geraads, D. (1998). Rongeurs du Miocène supérieur de Chorora (Ethiopie): Cricetidae, Rhizomyidae, Phiomyidae, Thryonomyidae, Sciuridae. Palaeovertebrata, 27, 203216.Google Scholar
Geraads, D. (2001). Rongeurs du Miocene supérieur de Chorora, Ethiopia: Murinae, Dendromurinae et conclusions. Palaeovertebrata, 30, 89109.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. (2010a). Rhinocerotidae. In Werdelin, L. & Sanders, W. J., eds. Cenozoic Mammals of Africa. Berkeley, CA: University of California Press, pp. 669683.Google Scholar
Geraads, D. (2010b). Tragulidae. In Werdelin, L. & Sanders, W. J., eds. Cenozoic Mammals of Africa. Berkeley, CA: University of California Press, pp. 723729.Google Scholar
Gervais, P. (1848–1852). Zoologie et Paléontologie Françaises. Nouvelles recherches sur les animaux vivants et fossiles de la France. Paris: Betrand.Google Scholar
Gheerbrandt, E., Schmitt, A. & Kocsis, L. (2018). Early African fossils elucidate the origin of embrithopod mammals. Current Biology, 28, 21672173.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. Annals of the Transvaal Museum, 44, 110.Google Scholar
Gilbert, C. C. (2013). Cladistic analysis of extant and fossil African papionins using craniodental data. Journal of Human Evolution, 64, 399433.Google Scholar
Gilbert, C. C., Steininger, C. M., Kibii, J. M. & 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, 115.Google Scholar
Gilbert, C. C., Frost, P. & Delson, E. (2016a). 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, C. C., Takahashi, M. Q. & Delson, E. (2016b). Cercopithecoid humeri from Taung support the distinction of major papionin clades in the South African fossil record. Journal of Human Evolution, 90, 88104.Google Scholar
Gill, T. (1884). Order III: Insectivora. In Kingsley, J. S., ed. The Standard Natural History, Vol. 5. Mammals. Boston, MA: Cassino.Google Scholar
Gingerich, P. D. (1974). Proteles cristatus Sparrman from the Pleistocene of South Africa, with a note on tooth replacement in the aardwolf (Mammalia: Hyaenidae). Annals of the Transvaal Museum, 29, 4954.Google Scholar
Ginsburg, L. (1965). L’Amphicyon ambiguus des phosphorites du Quercy. Bulletin de Museum National d’Histoire Naturelle, Series 2, 37, 724730.Google Scholar
Glenny, W. (2006). Report on the micromammal assemblage analysis from Sibudu Cave, KwaZulu-Natal. South African Humanities, 18, 279288.Google Scholar
Gmelin, J. F., ed. (1788). Linnaeus, C. Systema Naturae, 1, 13th edition. Lipsiae [Leipzig]: G. E. Beer.Google Scholar
Gommery, D. (2000). Superior cervical vertebrae of a Miocene hominoid and a Plio-Pleistocene hominid from southern Africa. Palaeontologia Africana, 36, 139145.Google Scholar
Gommery, D. (2008). A new hominid hip bone from Swartkrans (SKW 8012) in relation to the anatomy of the anterior inferior iliac spine. Annals of the Transvaal Museum, 45, 5566.Google Scholar
Gommery, D. & Bento da Costa, L. (2016). Les primates non-humains pliocènes et plio-pléistocènes d’Afrique du Sud. Revue de Primatologie, 7. DOI: 10.4000/primatologie.2698.Google Scholar
Gommery, D., Sénégas, F., Thackeray, J. F., et al. (2008a). Plio-Pleistocene fossils from Femur Dump, Bolt’s Farm, Cradle of Humankind World Heritage Site. Annals of the Transvaal Museum, 45, 6776.Google Scholar
Gommery, D., Thackeray, J. F., Potze, S. & Braga, J. (2008b). The first recorded occurrence of honey badger of the genus Mellivora (Carnivora: Mustelidae) at Kromdraai B, South Africa: scientific notes. Annals of the Transvaal Museum, 45, 145148.Google Scholar
Gommery, D., Thackeray, J. F., Sénégas, F., Potze, S. & Kgasi, L. (2008c). 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, 405408.Google Scholar
Gommery, D., Thackeray, J. F., Sénégas, F., Potze, S. & Kgasi, L. (2009). Additional fossils of Parapapio sp. from Waypoint 160 (Bolt’s Farm, South Africa), dated between 4 and 4.5 million years ago. Annals of the Transvaal Museum, 46, 6372.Google Scholar
Gommery, D., Badenhorst, S., Potze, 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, 3345.Google Scholar
Gommery, D., Badenhorst, S., Sénégas, F., Potze, S. & 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, 1931.Google Scholar
Gommery, D., Sénégas, F., Potze, S., Kgasi, L. & Thackeray, J. F. (2014). Cercopithecoidea material from the Middle Pliocene site, Waypoint 160, Bolt’s Farm, South Africa. Annals of the Ditsong National Museum of Natural History, 4, 18.Google Scholar
Gommery, D., Sénégas, F., Kgazi, L., et al. (2016). Bolt’s Farm cave system dans le Cradle of Humankind (Afrique du Sud): un exemple d’approche multidisciplinaire dans l’étude des sites à primates fossiles. Revue de Primatologie, 7. DOI: 10.4000/primatologie.2725.Google Scholar
Goodwin, A. J. H. (1937a). Archaeology of the Oakhurst Shelter, George. Part I. Course of the excavation. Transactions of the Royal Society of South Africa, 25, 229245.Google Scholar
Goodwin, A. J. H. (1937b). Archaeology of the Oakhurst Shelter, George. Part II. Disposition of the skeletal remains. Transactions of the Royal Society of South Africa, 25, 247257.Google Scholar
Goodwin, A. J. H. & Van Riet Lowe, C. (1929). The Stone Age cultures of South Africa. Annals of the South African Museum, 27, 1289.Google Scholar
Gordon, D. H. (1991). Chromosomal variation in the water rat Dasymys incomtus (Rodentia: Muridae). Journal of Mammalogy, 72, 411414.Google Scholar
Granjon, L., Duplantier, J.-M., Catalan, J. M. & Britton-Davidian, J. (1997). Systematics of the genus Mastomys (Thomas, 1915) (Rodentia: Muridae), a review. Belgian Journal of Zoology, 127 (Suppl.), 718.Google Scholar
Gray, J. E. (1824). A revision of the family Equidae. Zoological Journal, 1, 241248.Google Scholar
Gray, J. E. (1830). Spicilegia Zoologica; or Original Figures and Short Systematic Descriptions of New and Unfigured Animals. London: Treüttel, Würtz.Google Scholar
Gray, J. E. (1831). Zoological Miscellany. London: Treüttel, Würtz.Google Scholar
Gray, J. E. (1837). [Arrangement of the Sorices]. Proceedings of the Zoological Society of London, 5, 123126.Google Scholar
Gray, J. E. (1838). A revision of the genera of bats (Vespertilionidae) and the description of some new genera and species. Magazine of Zoology and Botany, 2, 483505.Google Scholar
Gray, J. E. (1842). Descriptions of some new genera and fifty unrecorded species of Mammalia. Annals and Magazine of Natural History, 10, 255267.Google Scholar
Gray, J. E. (1849). Description of a new species of Herpestes. Annals and Magazine of Natural History, 17, 11.Google Scholar
Gray, J. E. (1850a). Synopsis of the species of antelopes and strepsiceres, with descriptions of some new species. Proceedings of the Zoological Society of London, 18, 111146.Google Scholar
Gray, J. E. (1850b). Gleanings from the Menagerie and Aviary at Knowsley Hall. Knowsley: [private publication].Google Scholar
Gray, J. E. (1862a). Notice of a new species of bosh-buck (Cephalophus bicolor) from Natal. Annals and Magazine of Natural History, Series 3, 10, 400.Google Scholar
Gray, J. E. (1862b). Description of some new species of Mammalia. Proceedings of the Zoological Society of London, 1862, 261265.Google Scholar
Gray, J. E. (1864a). Notice of a new species of zorilla. Proceedings of the Zoological Society of London, 1864, 6970.Google Scholar
Gray, J. E. (1864b). Notes on the species of sand-moles (Georychus). Proceedings of the Zoological Society of London, 1864, 123125.Google Scholar
Gray, J. E. (1864c). A revision of the genera and species of viverrine animals (Viverridae), founded on the collection in the British Museum. Proceedings of the Zoological Society of London, 1864, 502578.Google Scholar
Gray, J. E. (1865a). A revision of the species of golden moles (Chrysochloris). Proceedings of the Zoological Society of London, 1865, 678680.Google Scholar
Gray, J. E. (1865b). Revision of the genera and species of entomophagous Edentata, founded on the examination of the specimens in the British Museum. Proceedings of the Zoological Society of London, 1865, 359386.Google Scholar
Gray, J. E. (1866). Notice of a new bat (Scotophilus welwitschii) from Angola. Proceedings of the Zoological Society of London, 1866, 211.Google Scholar
Gray, J. E. (1867). Observations on the preserved specimens and skeletons of the Rhinocerotidae in the collection of the British Museum and Royal College of Surgeons, including the descriptions of three new species. Proceedings of the Zoological Society of London, 1867, 10031032.Google Scholar
Gray, J. E. (1868). Revision of the species of Hyrax, founded on the specimens in the British Museums. Annals and Magazine of Natural History, Series 4, 1, 3551.Google Scholar
Gray, J. E. (1873). On the boomdas (Dendrohyrax arboreus). Annals and Magazine of Natural History, Series 4, 11, 154155.Google Scholar
Gray, P. A., Fenton, M. B. & Van Cakenberghe, V. (1999). Nycteris thebiaca. Mammalian Species, 612, 18.Google Scholar
Green, C. A., Keogh, H., Gordon, D. H., Pinto, M. & Hartwig, E. K. (1980). The distribution, identification and naming of the Mastomys natalensis species complex in southern Africa (Rodentia: Muridae). Journal of Zoology, 192, 1723.Google Scholar
Greenwood, M. (1955). Fossil Hystricoidea from the Makapan Valley. Palaeontologia Africana, 3, 7785.Google Scholar
Greenwood, M. (1958). Fossil Hystricoidea from the Makapan Valley, Transvaal: Hystrix makapanensis nom. nov. for Hystrix major Greenwood. Annals and Magazine of Natural History, Series 13, 1, 365.Google Scholar
Griffin, M. (1990). A review of taxonomy and ecology of gerbilline rodents of the central Namib Desert, with keys to the species (Rodentia: Muridae). In Seeley, M.K.. Namib ecology: 25 years of Namib Research. Transvaal Museum Monograph, 7, 8398.Google Scholar
Grine, F. E. (1981). Description of some juvenile hominid specimens from Swartkrans, Transvaal. Annals of the South African Museum, 86, 4371.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. (1989). New hominid fossils from the Swartkrans Formation (1979–1986 excavations): cranio-dental specimens. American Journal of Physical Anthropology, 79, 409450.Google Scholar
Grine, F. E. (1993). Description and preliminary analysis of new hominid craniodental fossils from the Swartkrans Formation. In Brain, C. K., ed. Swartkrans: a cave’s chronicle of early man. Transvaal Museum Monograph, 8, 75116.Google Scholar
Grine, F. E. (1998). Additional human fossils from the Middle Stone Age of Die Kelders Cave South Africa: 1995 excavation. South African Journal of Science, 94, 229235.Google Scholar
Grine, F. E. (2000). Middle Stone Age human fossils from Die Kelders Cave 1, Western Cape Province, South Africa. Journal of Human Evolution, 38, 129145.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, 750758.Google Scholar
Grine, F. E. & Daegling, W. L. (1993). New mandible of Paranthropus robustus from Member 1, Swartkrans Formation, South Africa. Journal of Human Evolution, 24, 319333.Google Scholar
Grine, F. E. & Hendey, Q. B. (1981). Earliest primate remains from South Africa. Journal of Mammalogy, 20, 6264.Google Scholar
Grine, F. E. & Klein, R. G. (1985). Pleistocene and Holocene human remains from Equus Cave, South Africa. Anthropology, 8, 5598.Google Scholar
Grine, F. E. & Klein, R. G. (1993). Late Pleistocene human remains from the Sea Harvest site, Saldanha Bay, South Africa. South African Journal of Science, 89, 145152.Google Scholar
Grine, F. E. & Strait, D. (1994). New hominid fossils from Member 1 Hanging Remnant, Swartkrans Formation, South Africa. Journal of Human Evolution, 26, 5775.Google Scholar
Grine, F. E. & Susman, R. L. (1991). Radius of Paranthropus robustus from Member 1, Swartkrans Formation, South Africa. American Journal of Physical Anthropology, 84, 229248.Google Scholar
Grine, F. E., Klein, R. G. & 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, 363395.Google Scholar
Grine, F. E., Jungers, W. L., Tobias, P. V. & Pearson, O. M. (1995). Fossil Homo femur from Berg Aukas, northern Namibia. American Journal of Physical Anthropology, 97, 151185.Google Scholar
Grine, F. E., Pearson, O. M., Klein, R. G. & Rightmire, G. P. (1998). Additional human fossils from Klasies River Mouth, South Africa. Journal of Human Evolution, 35, 95107.Google Scholar
Grine, F. E., Bailey, R. M., Harvati, K., et al. (2007). Late Pleistocene human skull from Hofmeyr, South Africa and modern human origins. Science, 315, 226229.Google Scholar
Grine, F. E., Gunz, P., Betti-Nash, L., Neubauer, S. & Morris, A. G. (2010). Reconstruction of the late Pleistocene human skull from Hofmeyr, South Africa. Journal of Human Evolution, 59, 115.Google Scholar
Grine, F. E., Jacobs, R. L., Reed, K. E. & Plavcan, J. M. (2012). The enigmatic molar from Gondolin, South Africa: implications for Paranthropus paleobiology. Journal of Human Evolution, 63, 597609.Google Scholar
Groves, C. P. (1972). Ceratotherium simum. Mammalian Species, 8, 16.Google Scholar
Groves, C. P. & Grubb, P. (2011). Ungulate Taxonomy. Baltimore, MD: Johns Hopkins University Press.Google Scholar
Grubb, P. (1981). Equus burchelli. Mammalian Species, 157, 19.Google Scholar
Grubb, P. (2004). Controversial scientific names of African mammals. African Zoology, 39, 91109.Google Scholar
Grubb, P. & d’Huart, J.-P. (2010). Rediscovery of the Cape warthog Phacochoerus aethiopicus: a review. Journal of the East African Natural History Society, 99, 77102.Google Scholar
Guérin, C. (1987). Fossil Rhinocerotidae (Mammalia, Perissodactyla) from Laetoli. In Leakey, M. D. & Harris, J. M., eds. Laetoli: A Pliocene site in northern Tanzania. Oxford: Clarendon Press, pp. 320348.Google Scholar
Guérin, C. (2000). The Neogene rhinoceroses of Namibia. Palaeontologia Africana, 36, 119138.Google Scholar
Guérin, C. (2003). Miocene Rhinocerotidae of the Orange River valley, Namibia. Memoirs of the Geological Survey of Namibia, 19, 257281.Google Scholar
Guérin, C. (2008). The Miocene Rhinocerotidae (Mammalia) from the northern Sperrgebiet, Namibia. Memoirs of the Geological Survey of Namibia, 20, 331341.Google Scholar
Haas, S. K., Hatssen, V. & Krausman, P. R. (2005). Panthera leo. Mammalian Species, 762, 111.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
Hamilton, W. R. (1973). A Lower Miocene Mammalian fauna from Siwa, Egypt. Palaeontology, 16, 275281.Google Scholar
Hamilton, W. R. (1978). Fossil giraffes from the Miocene of Africa and a revision of the phylogeny of the Giraffoidea. Philosophical Transactions of the Royal Society Series B, 283, 165229.Google Scholar
Hamilton, W. R. & Van Couvering, J. A. (1977). Lower Miocene mammals from South West Africa. Bulletin of Desert Ecological Research Unit, 2, 911.Google Scholar
Hanisch, E. O. M. (1980). An archaeological interpretation of certain Iron Age sites in the Limpopo/Shashi valley. Unpublished MA thesis, University of Pretoria.Google Scholar
Harris, J. M. (1976). Pliocene Giraffoidea (Mammalia, Artiodactyla) from the Cape Province. Annals of the South African Museum, 69, 325353.Google Scholar
Harris, J. M. (1977). Deinotheres from southern Africa. South African Journal of Science, 73, 281282.Google Scholar
Harris, J. M. & White, T. D. (1979). Evolution of the Plio-Pleistocene African Suidae. Transactions of the American Philosophical Society, 69, 1128.Google Scholar
Harris, J. M., Solounias, N. & Geraads, D. (2010). Giraffoidea. In Werdelin, L. & Sanders, W. J., eds. Cenozoic Mammals of Africa. Berkeley, CA: University of California Press, pp. 797811.Google Scholar
Harris, W. C. (1838a). [Aigererus niger]. Atheneum, 535, 71.Google Scholar
Harris, W. C. (1838b). [A new antelope from the Cape]. Proceedings of the Zoological Society of London, 1838, 13.Google Scholar
Harrison, T. (2010). Dendropithecoidea, Proconsuloidea and Hominoidea. In Werdelin, L. & Sanders, W. J., eds. Cenozoic Mammals of Africa. Berkeley, CA: University of California Press, pp. 429469.Google Scholar
Hartstone-Rose, A., De Ruiter, D., Berger, L. R. & Churchill, S. E. (2007). A sabre-tooth felid from Coopers Cave (Gauteng, South Africa) and its implications for Megantereon (Felidae: Machairodontinae) taxonomy. Palaeontologia Africana, 42, 99108.Google Scholar
Hartstone-Rose, A., Werdelin, L., De Ruiter, D., Berger, L. R. & Churchill, S. E. (2010). The Plio-Pleistocene ancestor of wild dogs, Lycaon sekoweni n. sp. Journal of Paleontology, 84, 299308.Google Scholar
Hartstone-Rose, A., Kuhn, B. F., Nalla, S. & Werdelin, L. (2013). A new species of fox from the Australopithecus sediba type locality, Malapa, South Africa. Transactions of the Royal Society of South Africa, 68, 19.Google Scholar
Harvati, K., Bauer, C. C., Grine, F. E., et al. (2015). A human deciduous molar from the Middle Stone Age (Howiesons Poort) of Klipdrift Shelter, South Africa. Journal of Human Evolution, 82, 190196.Google Scholar
Haughton, S. H. (1921). A note on some fossils from the Vaal River Gravels. Transactions of the Geological Society of South Africa, 24, 1116.Google Scholar
Haughton, S. H. (1924). A note on the occurrence of a species of baboon in limestone deposits near Taungs. Transactions of the Royal Society of South Africa, 12, lxviii.Google Scholar
Haughton, S. H. (1932a). The fossil Equidae of South Africa. Annals of the South African Museum, 28, 407427.Google Scholar
Haughton, S. H. (1932b). On some South African fossil Proboscidea. Transactions of the Royal Society of South Africa, 21, 118.Google Scholar
Haughton, S. H., Thomson, R. B. & Péringuey, L. (1917). Preliminary note on the ancient human skull-remains from the Transvaal. Transactions of the Royal Society of South Africa, 6, 113.Google Scholar
Häusler, M. & Berger, L. R. (2001). StW 441/465: a new fragmentary ilium of a small-bodied Australopithecus africanus from Sterkfontein, South Africa. Journal of Human Evolution, 40, 411417.Google Scholar
Hautier, L., Mackaye, H. T., Lihoreau, F. & Tassy, P. (2009). New material of Anancus kenyensis (Proboscidea, Mammalia) from Toros-Menalla (Late Miocene, Chad): contribution to the systematics of African anancines. Journal of African Earth Sciences, 53, 171176.Google Scholar
Hawks, J., Elliott, M., Schmid, P., et al. (2017). New fossil remains of Homo naledi from the Lesedi Chamber, South Africa. eLife, 6, e24232.Google Scholar
Heaton, J. L. (2006). Taxonomy of the Sterkfontein fossil Cercopithecinae: the Papionini of Members 2 and 4 (Gauteng, South Africa). Unpublished PhD thesis, Indiana University.Google Scholar
Heissig, K. (1971). Brachypotherium aus dem Miozän von Südwestafrikas. Mitteilungen aus dem Bayerischen Staatssammlung für Paläontologie und historische Geologie, 11, 125128.Google Scholar
Helbing, H. (1924). Das Genus Hyaenaelurus Biedermann. Eclogae Geologicae Helvetiae, 19, 214245.Google Scholar
Heller, E. (1912a). New genera and races of African ungulates. Smithsonian Miscellaneous Collections, 60(8), 116.Google Scholar
Heller, E. (1912b). New races of insectivores, bats and lemurs from British East Africa. Smithsonian Miscellaneous Collections, 60(12), 113.Google Scholar
Helm, C., Cawthra, H. C., Cowling, R., et al. (2018). Palaeoecology of giraffe tracks in Late Pleistocene aeolianites on the Cape south coast. South African Journal of Science, 114, 8.Google Scholar
Hemmer, H. (1965). Zur nomenklatur und verbreitung des genus Dinofelis Zdansky, 1924 (Therailurus Piveteau, 1948). Palaeontologia Africana, 9, 7589.Google Scholar
Hemprich, F. W. & Ehrenberg, C. G. (1828–1845). Symbolae Physicae, seu, Icones et Descriptiones Corporum Naturalium Noorum aut Minus Cognitorum. Berlin: Officina Academica.Google Scholar
Henderson, Z. (1992). The context of some Middle Stone Age hearths at Klasies River Shelter 1B: implications for understanding human behaviour. Southern African Field Archaeology, 1, 1426.Google Scholar
Hendey, Q. B. (1967). A specimen of ‘Archidiskodon’ cf transvaaalensis from the south-western Cape Province. South African Archaeological Bulletin, 22, 5356.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. (1969). Quaternary vertebrate fossil sites in the south-western Cape Province. South African Archaeological Bulletin, 24, 96105.Google Scholar
Hendey, Q. B. (1970). A review of the geology and palaeontology of the Plio/Pleistocene deposits at Langebaanweg, Cape Province. Annals of the South African Museum, 56, 75117.Google Scholar
Hendey, Q. B. (1972). A Pliocene ursid from South Africa. Annals of the South African Museum, 59, 115132.Google Scholar
Hendey, Q. B. (1973a). Carnivore remains from the Kromdraai australopithecine site (Mammalia: Carnivora). Annals of the Transvaal Museum, 28, 99112.Google Scholar
Hendey, Q. B. (1973b). Fossil occurrences at Langebaanweg, Cape Province. Nature, 244, 1314.Google Scholar
Hendey, Q. B. (1974a). The Late Cenozoic Carnivora of the south-western Cape Province. Annals of the South African Museum, 63, 1369.Google Scholar
Hendey, Q. B. (1974b). New fossil carnivores from the Swartkrans australopithecine site (Mammalia: Carnivora). Annals of the Transvaal Museum, 29, 2748.Google Scholar
Hendey, Q. B. (1976a). Fossil pecary from the Pliocene of South Africa. Science, 192, 787789.Google Scholar
Hendey, Q. B. (1976b). The Pliocene fossil occurrences in ‘E’ Quarry, Langebaanweg, South Africa. Annals of the South African Museum, 69, 215247.Google Scholar
Hendey, Q. B. (1978a). Late Tertiary Hyaenidae from Langebaanweg, South Africa and their relevance to the phylogeny of the family. Annals of the South African Museum, 76, 265297.Google Scholar
Hendey, Q. B. (1978b). Late Tertiary Mustelidae (Mammalia, Carnivora) from Langebaanweg, South Africa. Annals of the South African Museum, 76, 329357.Google Scholar
Hendey, Q. B. (1978c). Preliminary report on the Miocene vertebrates from Arrisdrift, South West Africa. Annals of the South African Museum, 76, 141.Google Scholar
Hendey, Q. B. (1978d). The age of the fossils from Baard’s Quarry, Langebaanweg, South Africa. Annals of the South African Museum, 75, 124.Google Scholar
Hendey, Q. B. (1980). Agriotherium (Mammalia, Ursidae) from Langebaanweg, South Africa and relationships of the genus. Annals of the South African Museum, 81, 1109.Google Scholar
Hendey, Q. B. (1981). 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. (1984). Southern African late Tertiary vertebrates. In Klein, R. G., ed. Southern African Prehistory and Paleoenvironments. Rotterdam: Balkema, pp. 81106.Google Scholar
Hendey, Q. B. & Cooke, H. B. S. (1985). Kolpochoerus paiceae (Mammalia, Suidae) from Skuurwerug, near Saldanha, South Africa and its palaeoenvironmental implications. Annals of the South African Museum, 97, 956.Google Scholar
Hendey, Q. B. & Hendey, H. (1968). New Quaternary fossil sites near Swartklip, Cape Province. Annals of the South African Museum, 52, 4373.Google Scholar
Hendey, Q. B. & Singer, R. (1965). Part III: The faunal assemblages from the Gamtoos Valley shelters. South African Archaeological Bulletin, 20, 206213.Google Scholar
Henshilwood, C. S. (1995). Holocene archaeology of the coastal Garcia State Forest, southern Cape, South Africa. Unpublished PhD thesis, University of Cambridge.Google Scholar
Henshilwood, C. S. (1996). A revised chronology for pastoralism in southernmost Africa: new evidence of sheep at c. 2000 b.p. from Blombos Cave, South Africa. Antiquity, 70, 945949.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., Van Niekerk, K. L., Wurz, S., et al. (2014). Klipdrift Shelter, southern Cape, South Africa: preliminary report on the Howiesons Poort layers. Journal of Archaeological Science, 45, 284303.Google Scholar
Herries, A. I. R., Adams, J. W., Kuykendall, K. I. & Shaw, J. (2006). 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, 617631.Google Scholar
Heuglin, T. (1863). Beiträge zur Zoologie Afrika’s. Über einige Säugethiere des Bäschlo-Gebietes. Novorum Actorum Academiae Caesareae Leopoldino-Carolinae Germanicae Naturae Curiosorum, 30, 114.Google Scholar
Heuglin, T. (1865). Beschreibung eines centralafrikanischen Leporinen. Novorum Actorum. Academiae Caesareae Leopoldino-Carolinae Germanicae Naturae Curiosorum. Leopoldina, 32-24, 3236.Google Scholar
Hill, J. & Harrison, D. L. (1987). The baculum in the Vespertilioninae (Chiroptera: Vespertilionidae) with a systematic review, a synopsis of Pipistrellus and Eptesicus and the descriptions of a new genus and subgenus. Bulletin of the British Museum (Natural History) Zoological Series, 52, 225305.Google Scholar
Hillestad Nel, T. & 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. DOI: 10.1371/journal.pone.0159817.Google Scholar
Hillman-Smith, A. K. K. & Groves, C. P. (1994). Diceros bicornis. Mammalian Species, 455, 18.Google Scholar
Hodgson, B. H. (1847). On the various genera of the ruminants. Journal of the Asiatic Society of Bengal, Series 3, 16, 685711.Google Scholar
Hoffman, A. C. (1953). The fossil alcelaphines of South Africa: genera Peloceras, Lunatoseras and Alcelaphus. Navorsinge van die Nasionale Museum Bloemfontein, 1, 4156.Google Scholar
Holden, M. E. (1996). Systematic revision of sub-Saharan African dormice (Rodentia: Myoxidae: Graphiurus) part 1: an introduction to the generic revision and a revision of Graphiurus surdus. American Museum Novitates, 3157, 144Google Scholar
Holloway, R. L., Hurst, S. D., Garvin, H. M., et al. (2018). Endocast morphology of Homo naledi from the Dinaledi Chamber, South Africa. Proceedings of the National Academy of Sciences of the United States of America. DOI: 10.1073/pnas.1720842115.Google Scholar
Holroyd, P. A. (1999). New Pterodontinae (Creodonta: Hyaenodontidae) from the late Eocene–early Oligocene Jebel Qatrani Formation, Fayum province, Egypt. PaleoBios, 19, 118.Google Scholar
Holroyd, P. A. (2010a). Macroscelidea. In Werdelin, L. & Sanders, W. J., eds. Cenozoic Mammals of Africa. Berkeley, CA: University of California Press, pp. 8998.Google Scholar
Holroyd, P. A. (2010b). Tubulidentata. In Werdelin, L. & Sanders, W. J., eds. Cenozoic mammals of Africa. Berkeley, CA: University of California Press. pp. 107–111.Google Scholar
Holroyd, P. A., Lihoreau, F., Gunnell, G. F. & Miller, E. R. (2010). Anthracotheriidae. In Werdelin, L. & Sanders, W. J., eds. Cenozoic Mammals of Africa. Berkeley, CA: University of California Press, pp. 843851.Google Scholar
Holt, S. (2009). The faunal remains from the Makgabeng Plateau, Limpopo Province. Unpublished MSc thesis, University of the Witwatersrand.Google Scholar
Hooijer, D. A. (1945). Note on the subfossil teeth of Equus zebra L. from the Orange Free State. Zoologische Mededelingen Museum Leiden, 25, 101108.Google Scholar
Hooijer, D. A. (1958). Pleistocene remains of hippopotamus from the Orange Free State. Navorsinge van die Nasionale Museum Bloemfontein, 1, 259266.Google Scholar
Hooijer, D. A. (1959). Fossil rhinoceroses from the Limeworks Cave, Makapansgat. Palaeontologia Africana, 6, 113.Google Scholar
Hooijer, D. A. (1963). Miocene Mammalia of Congo. Annales du Musée Royal de l’Afrique Centrale Sciences Géologiques, 46, 177.Google Scholar
Hooijer, D. A. (1971). A new rhinoceros from the Late Miocene of Loporot, Turkana District, Kenya. Bulletin of the Museum of Comparative Zoology, 142, 339392.Google Scholar
Hooijer, D. A. (1972). A late Pliocene rhinoceros from Langebaanweg. Annals of the South African Museum, 59, 151191.Google Scholar
Hooijer, D. A. (1973). Additional Miocene to Pleistocene rhinoceroses of Africa. Zoologische Mededelingen, 46, 149177.Google Scholar
Hooijer, D. A. (1975). Miocene to Pleistocene hipparions of Kenya, Tanzania and Ethiopia. Zoologische Verhandelingen, 142, 380.Google Scholar
Hooijer, D. A. (1976). The late Pliocene Equidae of Langebaanweg, Cape Province, South Africa. Zoologische Verhandelingen, 148, 339.Google Scholar
Hooijer, D. A. & Patterson, B. (1972). Rhinoceroses from the Pliocene of Northwestern Kenya. Bulletin on the Museum of Comparative Zoology, 144, 126.Google Scholar
Hooijer, D. A. & Singer, R. (1960). Fossil rhinoceroses from Hopefield, South Africa. Zoologische Mededelingen, 37, 113128.Google Scholar
Hooijer, D. A. & Singer, R. (1961). The fossil hippopotamus from Hopefield, South Africa. Zoologische Mededelingen, 37, 157165.Google Scholar
Hopley, P. J., Latham, A. G. & Marshall, J. D. (2006). Palaeoenvironments and palaeodiets of mid-Pliocene micromammals from Makapansgat Limeworks, South Africa: a stable isotope and dental microwear approach. Palaeogeography, Palaeoclimatology, Palaeoecology, 233, 235251.Google Scholar
Hopwood, A. T. (1926). Some Mammalia from the Pliocene of Homa Mountain, Victoria Nyanza. Annals and Magazine of Natural History, Series 9, 18, 266272.Google Scholar
Hopwood, A. T. (1929). New and little known mammals from the Miocene of Africa. American Museum Novitates, 76, 19.Google Scholar
Hopwood, A. T. (1934). New fossil mammals from Olduvai, Tanganyika Territory. Annals and Magazine of Natural History, Series 10, 14, 546550.Google Scholar
Hopwood, A. T. (1936). New and little-known fossil mammals from Kenya Colony and Tanganyika Territory I. Annals and Magazine of Natural History, Series 10, 17, 636641.Google Scholar
Horsburgh, K. A. & Moreno-Mayar, V. J. (2015). Molecular identification of sheep at Blydefontein Rock Shelter, South Africa. Southern African Humanities, 27, 6580.Google Scholar
Horsburgh, K. A. & Rhines, A. (2010). Genetic characterization of an archaeological sheep assemblage from South Africa’s Western Cape. Journal of Archaeological Science, 37, 29062910.Google Scholar
Horsburgh, K. A., Orton, J. & Klein, R. G. (2016). Beware the springbok in sheep’s clothing: how secure are the faunal identifications upon which we build our models? African Archaeological Review, 33, 353361.Google Scholar
Huchon, D., Catzeflis, F. & Douzery, E. J. P. (2000). Variance of molecular datings, evolution of rodents, and the phylogenetic affinities between Ctenodactylidae and Hystricognathi. Proceedings of the Royal Society of London B: Biological Sciences, 267, 393402.Google Scholar
Huchon, D., Madsen, O., Sibbald, M. J. J. B., et al. (2002). Rodent phylogeny and a timescale for the evolution of Glires: evidence from an extensive taxon sampling using three nuclear genes. Molecular Biology and Evolution, 19, 10531065.Google Scholar
Huffman, T. N. (1975). Cattle from Mabveni. South African Archaeological Bulletin, 30, 2324.Google Scholar
Huffman, T. N. (1979a). Test excavations at Naba and Lanlory, Northern Mashonaland. South African Archaeological Society Goodwin Series, 3, 1446.Google Scholar
Huffman, T. N. (1979b). Test excavations at Chamabvefva, southern Mashonaland. South African Archaeological Bulletin, 34, 5770.Google Scholar
Huffman, T. N. (2008). Zhizo and Leopard’s Kopje: test excavations at Simamwe and Mtanye, Zimbabwe. In Badenhorst, S., Mitchell, P. & Driver, J. C., eds. Animals and People: Archaeozoological Papers in Honour of Ina Plug. Oxford: Archaeopress, pp. 200214.Google Scholar
Huffman, T. N. & Kinahan, J. (2002/2003). Archaeological mitigation of the Letsibogo Dam: agropastoralism in southeastern Botswana. Southern African Field Archaeology, 11–12, 463.Google Scholar
Hughes, A. (1990). The Tuinplaas human skeleton from the Springbok Flats, Transvaal. In Sperber, G. H., ed. From Apes to Angels: Essays in Honour of Phillip V. Tobias. New York: Wiley-Liss, pp. 197214.Google Scholar
Humphreys, A. J. B. (1974). A preliminary report on test excavations at Dikbosch Shelter 1, Herbert District, northern Cape. South African Archaeological Bulletin, 29, 115119.Google Scholar
Humphreys, A. J. B. (1975). Burchell’s Shelter: the history and archaeology of a Northern Cape rock shelter. South African Archaeological Bulletin, 30, 318.Google Scholar
Humphreys, A. J. B. (1978). The re-excavation of Powerhouse Cave and an assessment of Dr Frank Peabody’s work on Holocene deposits in the Taung area. Annals of the Cape Provincial Museums, 11, 217244.Google Scholar
Humphreys, A. J. B. & Thackeray, A. I. (1983). Ghaap and Gariep. Cape Town: South African Archaeological Society.Google Scholar
Hutson, J. M. (2006). Taphonomy at Kalkbank: a Late Pleistocene site in the Limpopo Province, South Africa. Unpublished MSc thesis, University of the Witwatersrand.Google Scholar
Hutson, J. M. (2016). The faunal remains from Bundu Farm and Pniel 6: examining the problematic Middle Stone Age archaeological record within the southern African interior. Quaternary International. DOI: 10.1016/j.quaint.2016.04.030.Google Scholar
Hutson, J. M. & Cain, C. R. (2008). Reanalysis and reinterpretation of the Kalkbank faunal accumulation, Limpopo Province, South Africa. Journal of Taphonomy, 6, 399428.Google Scholar
Hutterer, R. (1983). Taxonomy and distribution of Crocidura fuscomurina (Heuglin, 1865). Mammalia, 47, 221227.Google Scholar
Illiger, J. K. W. (1811). Prodromus Systematis Mammaliam et Avium. Berlin: C. Salfeld.Google Scholar
Illiger, J. K. W. (1815). Ueberblick der Säugethiere nach ihrer Vertheilung über die Welheile. Abhandlungen der physikalischen Klasse der Königlich-Preussischen Akademie der Wissenschaften, 1804–1811, 39159.Google Scholar
Ingram, C. M., Burda, H. & Honeycutt, R. L. (2004). Molecular phylogenetics and taxonomy of the African mole-rats, genus Cryptomys and the new genus Coetomys Gray, 1864. Molecular Phylogenetics and Evolution, 31, 9971014.Google Scholar
Jablonsky, N. G. & Frost, S. (2010). Cercopithecoidea. In Werdelin, L. & Sanders, W. J., eds. Cenozoic Mammals of Africa. Berkeley, CA: University of California Press, pp. 393428.Google Scholar
Jacobs, B. F., Pan, A. D. & Scotese, C. R. (2010). A review of the Cenozoic vegetation history of Africa. In Werdelin, L. & Sanders, W. J., eds. Cenozoic Mammals of Africa. Berkeley, CA: University of California Press, pp. 5772.Google Scholar
Jacobson, L. (1978). Report on archaeological and palaeoecological studies in the Gobabis District, South West Africa. Palaeoecology of Africa, 10, 9394.Google Scholar
Jaeger, J. J., Michaux, J. & Sabatier, M. (1980). Premières données sur les rongeurs de la Formation de Cho’rora (Ethiopie) d’âge Miocène supérieur. I: Thryonomyidés. Palaeovertebrata, 9 Ext. Mémoire Jubilaire en Hommage à R. Lavocat, 365374.Google Scholar
Jaeger, J. J., Denys, C. & Coiffait, B. (1985). New Phiomorpha and Anomaluridae from the late Eocene of north-west Africa: phylogenetic implications. In Luckett, W. P. & Hartenberger, J.-L., eds. Evolutionary Relationships among Rodents, a Multidisciplinary Analysis. New York: Plenum, pp. 567588.Google Scholar
Jaeger, J. J., Marivaux, L., Salem, M., et al. (2010). New rodent assemblages from the Eocene Dur At-Talah escarpment (Sahara of central Libya): systematic, biochronological and palaeobiogeographical implications. Zoological Journal of the Linnean Society, 160, 195213.Google Scholar
Jameson, H. L. (1907). On a new hare from the Transvaal. Annals and Magazine of Natural History, Series 7, 20, 404406.Google Scholar
Jameson, H. L. (1909). On a sub-fossil hare from a cave deposit at Godwan River. Annals of the Transvaal Museum, 1, 195196.Google Scholar
Jansen van Vuuren, B. (1999). Molecular phylogeny of duiker antelope (Mammalia: Cephalophini). Unpublished PhD thesis, University of Pretoria.Google Scholar
Jenkins, P., Ruedi, M. & Catzeflis, F. (1998). A biochemical and morphological investigation of Suncus dayi (Dobson, 1888) and discussion of relationships in Suncus Hemprich & Ehrenberg, 1833, Crocidura Wagler, 1832 and Sylvisorex Thomas, 1904 (Insectivora: Soricidae). Bonner Zoologische Beitragen, 47, 257276.Google Scholar
Jerardino, A. (1998). Excavations at Pancho’s Kitchen Midden, western Cape coast, South Africa: further observations into the Megamidden Period. South African Archaeological Bulletin, 53, 1625.Google Scholar
Jerardino, A. (2007). Excavations at a hunter-gatherer site known as ‘Grootrif G’ Shell Midden, Lamberts Bay, Western Cape Province. South African Archaeological Bulletin, 62, 162170.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. Journal of Island and Coastal Archaeology, 7, 76101.Google Scholar
Jerardino, A. & Yates, R. (1996). Preliminary results from excavations at Steenbokfontein Cave: implications for past and future research. South African Archaeological Bulletin, 51, 716.Google Scholar
Jerardino, A., Yates, R., Morris, A. G. & Sealy, J. C. (1992). A dated human burial from the Namaqualand coast: observations on culture, biology and diet. South African Archaeological Bulletin, 47, 7581.Google Scholar
Jerardino, A., Dewar, G. & Navarro, R. (2009a). Opportunistic subsistence strategies among Late Holocene coastal hunter-gatherers, Elands Bay, South Africa. Journal of Island and Coastal Archaeology, 4, 3760.Google Scholar
Jerardino, A., Kolska Horwitz, L., Mazel, A. & Navarro, R. (2009b). Just before Van Riebeeck: glimpses into terminal LSA lifestyle at Connies Limpet Bar, west coast of South Africa. South African Archaeological Bulletin, 64, 7586.Google Scholar
Jerardino, A., Klein, R. G., Navarro, R., Orton, J. & Kolska Horwitz, L. (2013). Settlement and subsistence patterns since the terminal Pleistocene in the Elands Bay and Lamberts Bay areas. In Jerardino, A., Malan, A. & Braun, D., eds. The Archaeology of the West Coast of South Africa. Oxford: Archaeopress, pp. 85108.Google Scholar
Jerardino, A., Kaplan, J., Navarro, R. & Nilssen, P. (2016). Filling-in the gaps and testing past scenarios on the central West Coast: hunter-gatherer subsistence and mobility at ‘Deurspring 16’ shell midden, Lamberts Bay, South Africa. South African Archaeological Bulletin, 71, 7186.Google Scholar
Jones, T. R. (1937). A new fossil primate from Sterkfontein, Krugersdorp, Transvaal. South African Journal of Science, 33, 709728.Google Scholar
Kandel, A. W. & 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
Kaplan, J. (1987). Settlement and subsistence at Renbaan Cave. In Parkington, J. & Hall, M., eds. Papers in the Prehistory of the Western Cape, South Africa. Oxford: Archaeopress, pp. 237261.Google Scholar
Kaplan, J. (1990). The Umhlatuzana Rock Shelter sequence: 100 000 years of Stone Age history. Natal Museum Journal of Humanities, 2, 194.Google Scholar
Katsamudanga, S. (2007a). Environment and culture: a study of prehistoric settlement patterns in the Eastern Highlands of Zimbabwe. Unpublished PhD thesis, University of Zimbabwe.Google Scholar
Katsamudanga, S. (2007b). Archaeological surveys in Zimunya, Burma Valley, Vumba and Tsetsera Mountains, central Eastern Highlands of Zimbabwe. Zimbabwea, 9, 919.Google Scholar
Kaup, J. J. (1828). Ueber Hyaena, Uromastix, Basiliscus, Corythaeolus, Acontoias. Isis von Oken, 21, 11441150.Google Scholar
Kaup, J. J. (1829). Skizzirte Entwickelungs-Geschichte und natürliches System der europäischen Thierwelt, Vol. 1. Darmstadt: Leske.Google Scholar
Kaup, J. J. (1833). Déscription d’Ossements Fossiles de Mammifères Inconnus jusqu’à Présent, qui se trouvent au Museum Grand-Ducal de Darmstadt, Vol. 2. Darmstadt: Stahl and Bekker.Google Scholar
Kaup, J. J. (1835). Das Thierreich in seinen Hauptformen. Darmstadt: J. P. Biehl.Google Scholar
Kearney, T. & Van Schalkwyk, E. (2009). Variation in the position within the tooth row of the minute premolars of Cistugo lesueuri and C. seabrae (Chiroptera: Vespertilionidae) and re-identification of some museum voucher specimens. Annals of the Transvaal Museum, 46, 117120.Google Scholar
Kearney, T., Volleth, M., Contrafatto, G. & Taylor, P. (2002). Systematic implications of chromosome GTG-band and bacula morphology for southern African Eptesicus and Pipistrellus and several other species of Vespertilioninae (Chiroptera: Vespertilionidae). Acta Chiropterologica, 4, 5576.Google Scholar
Keen, E. N. & Singer, R. (1956). Further fossil Suidae from Hopefield. Annals of the South African Museum, 42, 351360.Google Scholar
Kegley, A. D. T., Hemingway, J. & Adams, J. W. (2011). Odontometric analysis of the reanalyzed and expanded Cercopithecoides from the Haasgat fossil assemblage, Cradle of Humankind, South Africa. American Journal of Physical Anthropology, S144, 183.Google Scholar
Keith, A. (1933). A descriptive account of the human skulls from Matjes River Cave, Cape Province. Transactions of the Royal Society of South Africa, 21, 151185.Google Scholar
Kerr, R. (1792). The Animal Kingdom, or Zoological System of the Celebrated Sir Charles Linnaeus. Edinburgh: Strahan & Cadell and Creech.Google Scholar
Keyser, A. (1991). The palaeontology of Haasgat: a preliminary account. Palaeontologia Africana, 28, 2933.Google Scholar
Keyser, A. W. (2000). The Drimolen skull: the most complete australopithecine cranium and mandible to date. South African Journal of Science, 96, 189193.Google Scholar
Keyser, A. W. & Martini, J. E. J. (1991). Haasgat: a new Plio-Pleistocene fossil occurrence. Palaeoecology of Africa, 21, 119129.Google Scholar
Keyser, A. W., Menter, C., Moggi-Cecchi, J., Pickering, T. R. & Berger, L. R. (2000). Drimolen: a new hominid-bearing site in Gauteng, South Africa. South African Journal of Science, 96, 193197.Google Scholar
Kiberd, P. (2006). Bundu Farm: a report on archaeological and palaeoenvironmental assemblages from a pan site in Bushmanland, Northern Cape, South Africa. South African Archaeological Bulletin, 61, 189201.Google Scholar
Kibii, J. M. (2006). Comparative taxonomic, taphonomic and palaeoenvironmental analysis of 4–2.3 million year old australopithecine cave infills at Sterkfontein. Unpublished PhD thesis, University of the Witwatersrand.Google Scholar
Kinahan, J. & Kinahan, J. (2003). Excavation of a late Holocene cave deposit in the southern Namib Desert, Namibia. Cimbebasia, 18, 110.Google Scholar
Kitching, J. W. (1963). A fossil Orycteropus from the Limeworks quarry, Makapansgat, Potgietersrus. Palaeontologia Africana, 8, 119121.Google Scholar
Kitching, J. W. (1965). A new giant hyracoid from the Limeworks Quarry, Makapansgat, Potgietersrus. Palaeontologia Africana, 9, 9196.Google Scholar
Klatzow, S. (1994). Roosfontein, a contact site in the Eastern Orange Free State. South African Archaeological Bulletin, 49, 915.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 for environmental and cultural change. Quaternary Research, 2, 135142.Google Scholar
Klein, R. G. (1974a). On the taxonomic status, distribution and ecology of the blue antelope, Hippotragus leucophaeus (Pallas, 1766). Annals of the South African Museum, 65, 99143.Google Scholar
Klein, R. G. (1974b). Environment and subsistence of prehistoric man in the southern Cape Province. World Archaeology, 5, 249284.Google Scholar
Klein, R. G. (1975a). Paleoanthropological implications of the non-archeological bone assemblage from Swartklip 1, south-western Cape Province, South Africa. Quaternary Research, 5, 275288.Google Scholar
Klein, R. G. (1975b). Middle Stone Age man–animal relationships in southern Africa: evidence from Die Kelders and Klasies River Mouth. Science, 190, 265267.Google Scholar
Klein, R. G. (1976a). 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. (1976b). A preliminary report on the ‘Middle Stone Age’ open-air site of Duinefontein 2 (Melkbosstrand, south-western Cape Province, South Africa). South African Archaeological Bulletin, 31, 1220.Google Scholar
Klein, R. G. (1976c). The fossil history of Raphicerus H. Smith, 1827 (Bovidae, Mammalia) in the Cape Biotic Zone. Annals of the South African Museum, 71, 169191.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, 32, 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). Preliminary analysis of the mammalian fauna from the Redcliff Stone Age site, Rhodesia. Occasional Papers of the National Museums and Monuments of Rhodesia Series A Human Sciences, 4, 7480.Google Scholar
Klein, R. G. (1978c). 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. (1978d). Appendix: vertebrate fauna from the Buffelskloof Rock Shelter. In Opperman, H., Excavations in the Buffelskloof Rock Shelter near Calitzdorp, Southern Cape. South African Archaeological Bulletin, 33, 3538.Google Scholar
Klein, R. G. (1979a). Palaeoenvironmental and cultural implications of late Holocene archaeological faunas from the Orange Free State and north-central Cape Province, South Africa. South African Archaeological Bulletin, 34, 3449.Google Scholar
Klein, R. G. (1979b). Appendix III: macromammals. In Cooke, C. K., Excavations at Diana’s Vow, Rock Shelter: Makomi District, Zimbabwe-Rhodesia. Salisbury: National Museums and Monuments, 147148.Google Scholar
Klein, R. G. (1980). Appendix 1: larger mammals. In Maggs, T. & Ward, V., Driel Shelter: rescue at a Late Stone Age site on the Tugela River. Annals of the Natal Museum, 24, 6267.Google Scholar
Klein, R. G. (1981). Later Stone Age subsistence at Byeneskranskop Cave, South Africa. In Harding, R. S. O. & Teleki, G., eds. Omnivorous Primates: Gathering and Hunting in Human Evolution. New York: Columbia University Press, pp. 166190.Google Scholar
Klein, R. G. (1982). Appendix 1: Byneskranskop 1 – mammals. Tables of the minimum numbers of individuals represented by various skeletal parts. In Schweitzer, F. R. & Wilson, M. L.. Byneskranskop 1: a late Quaternary living site in the southern Cape Province, South Africa. Annals of the South African Museum, 82, 189196.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). Appendix 1: the remains of larger mammals from Fairview Shelter. In Robertshaw, P. T.. Fairview Rockshelter: a contribution to the prehistory of the eastern Cape Province of South Africa. Annals of the Cape Provincial Museums (Human Sciences), 1, 8286.Google Scholar
Klein, R. G. (1986). The prehistory of Stone Age herders in the Cape Province of South Africa. South African Archaeological Society Goodwin Series, 5, 512.Google Scholar
Klein, R. G. (1988). The archaeological significance of animal bones from Acheulean sites in southern Africa. The African Archaeological Review, 6, 325.Google Scholar
Klein, R. G. (1994a). The long-horned African buffalo (Pelorovis antiquus) is an extinct species. Journal of Archaeological Science, 21, 725733.Google Scholar
Klein, R. G. (1994b). Southern Africa before the Iron Age. In Corruccini, R. S. & Ciochon, R. L., eds. Integrative Paths to the Past: Paleoanthropological Advances in Honor of F. Clark Howell. Englewood Cliffs, NJ: Prentice Hall, pp. 471519.Google Scholar
Klein, R. G. & Cruz-Uribe, K. (1987). Large mammal and tortoise bones from Elands Bay Cave and nearby sites, western Cape Province, South Africa. In Parkington, J. & Hall, M., eds. Papers in the Prehistory of the Western Cape, South Africa. Oxford: Archaeopress, pp. 132163.Google Scholar
Klein, R. G. & Cruz-Uribe, K. (1989). Faunal evidence for prehistoric herder-forager activities at Kasteelberg, Western Cape Province, South Africa. South African Archaeological Bulletin, 44, 8297.Google Scholar
Klein, R. G. & 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. & Cruz-Uribe, K. (1999). Craniometry of the genus Equus and the taxonomic affinities of the extinct South African quagga. South African Journal of Science, 95, 8186.Google Scholar
Klein, R. G. & Cruz-Uribe, K. (2000). Middle and Later Stone Age large mammal and tortoise remains from Die Kelders Cave 1, Western Cave Province, South Africa. Journal of Human Evolution, 38, 169195.Google Scholar
Klein, R. G. & Scott, K. (1974). The fauna of Scott’s Cave, Gamtoos Valley, south-eastern Cape Province. South African Journal of Science, 70, 186187.Google Scholar
Klein, R. G. & Steele, T. E. (2008). The faunal remains from Diepkloof Rock Shelter, South Africa. In P.-J. Texier. Diepkloof (Western Cape, République d’Afrique du Sud). Rapport sur les travaux effectués du 16 octobre 2007 au 15 octobre 2008. Unpublished CNRS report.Google Scholar
Klein, R. G., Cruz-Uribe, K. & 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. (1999a). Duinefontein 2: an Acheulean site in the Western Cape Province of South Africa. Journal of Human Evolution, 37, 153190.Google Scholar
Klein, R. G., Cruz-Uribe, K., Halkett, D., Hart, T. & Parkington, J. E. (1999b). Paleoenvironmental and human behavioral implications of the Boegoeberg 1 Late Pleistocene hyena den, Northern Cape Province, South Africa. Quaternary Research, 52, 393403.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 of the United States of America, 101, 57085715.Google Scholar
Klein, R. G., Avery, G., Cruz-Uribe, K. & Steele, T. E. (2007). The mammalian fauna associated with an archaic hominin skullcap and later Acheulian artifacts at Elandsfontein, Western Cape Province, South Africa. Journal of Human Evolution, 52, 164186.Google Scholar
Kock, D., Ingram, C. M., Frabotta, L. J., Honeycutt, R. L. & Burda, H. (2006). On the nomenclature of Bathyergidae and Fukomys n. gen. (Mammalia: Rodentia). Zootaxa, 1142, 5155.Google Scholar
Koehler, C. E. & Richardson, P. R. K. (1990). Proteles cristatus. Mammalian Species, 363, 16.Google Scholar
Kolbe, F. F. (1948). On a hitherto unrecorded subspecies of South African bush hare (Lepus saxatilis orangensis n. subsp.). Annals of the Transvaal Museum, 21, 7172.Google Scholar
Korsman, S. & Plug, I. (1994). Two Later Stone Age sites on the Farm Honingklip in the Eastern Transvaal. South African Archaeological Bulletin, 49, 2432.Google Scholar
Krausman, P. R. & Morales, S. M. (2005). Acinonyx jubatus. Mammalian Species, 771, 16.Google Scholar
Kretzoi, N. (1929). Materialien zur phylogenetischen Klassifikation der Aeluroideen. In 10th International Congress of Zoology (1927), Vol. 2. Budapest: Stephaneum, pp. 12931355.Google Scholar
Kryštufek, B., Haberl, W., Baxter, R. M. & Zima, J. (2004). Morphology and karyology of two populations of the woodland dormouse Graphiurus murinus in the Eastern Cape, South Africa. Folia Zoologica, 53, 339350.Google Scholar
Kryštufek, B., Baxter, R. M., Haberl, W., Zima, J. & Bužan, E. V. (2008). Systematics and biogeography of the Mozambique thicket rat, Grammomys cometes, in Eastern Cape Province, South Africa. Journal of Mammalogy, 89, 325335.Google Scholar
Kuhl, H. (1817). Die Deutschen Fledermäuse. Hanau: Universitätsbibliothek Johann Christian Senckenberg.Google Scholar
Kuhn, B. F., Werdelin, L., Hartstone-Rose, A., Lacruz, R. & Berger, L. R. (2011). Carnivoran remains from the Malapa Hominin Site, South Africa. PLoS ONE, 6. DOI: 10.1371/journal.pone.0026940.Google Scholar
Kuhn, B. F., Herries, A. I. R., Price, G. J., et al. (2016). Renewed investigations at Taung: 90 years after the discovery of Australopithecus africanus. Palaeontologia Africana, 51, 1026.Google Scholar
Kuhn, B. F., Werdelin, L. & Steininger, C. (2017). Fossil Hyaenidae from Cooper’s Cave, South Africa and the palaeoenvironmental implications. Palaeobiodiversity and Palaeoenvironments, 97, 355365.Google Scholar
Kuman, K. & Clarke, R. J. (1986). Florisbad: new investigations at a Middle Stone Age site in South Africa. Geoarchaeology, 1, 103125.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. & Conroy, G. C. (1999). Description of the Gondolin teeth: hyper-robust hominids in South Africa? American Journal of Physical Anthropology, 28, 176177.Google Scholar
Kuykendall, K. L. & Rae, T. C. (2008). Presence of the maxillary sinus in fossil Colobinae (Cercopithecoides williamsi) from South Africa. The Anatomical Record, 291, 14991506.Google Scholar
Kuykendall, K. L., Toich, C. A. & McKee, J. K. (1995). Preliminary analysis of the fauna from Buffalo Cave, northern Transvaal, South Africa. Palaeontologia Africana, 32, 2731.Google Scholar
L’Abbé, E. N., Loots, M. & Keough, N. (2008). The Matjes River Rock Shelter: a description of the skeletal assemblage. South African Archaeological Bulletin, 63, 6168.Google Scholar
Lacépède, E. B. G. (1799). Tableau des Divisions, Sous-divisions, Ordres, et Genres des Mammifères. Paris: Plasson.Google Scholar
Lacruz, R. (2009). Panthera leo (Mammalia: Felidae) remains from the Gladysvale Cave, South Africa: scientific note. Annals of the Transvaal Museum, 46, 121124.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 the Gladysvale Cave, South Africa. Palaeontologia Africana, 38, 99114.Google Scholar
Lacruz, R., Ungar, P., Hancox, P. J., Brink, J. S. & Berger, L. R. (2003). Gladysvale: fossils, strata and GIS analysis. South African Journal of Science, 99, 283285.Google Scholar
Lacruz, R., Turner, A. & Berger, L. R. (2006). New Dinofelis (Carnivora: Machairodontinae) remains from Sterkfontein Valley sites and a taxonomic revision of the genus in southern Africa. Annals of the Transvaal Museum, 43, 89106.Google Scholar
Larivière, S. (2001a). Aonyx capensis. Mammalian Species, 671, 16.Google Scholar
Larivière, S. (2001b). Poecilogale albinucha. Mammalian Species, 681, 14.Google Scholar
Larivière, S. (2002). Ictonyx striatus. Mammalian Species, 698, 15.Google Scholar
Larivière, S. & Calzada, J. (2001). Genetta genetta. Mammalian Species, 680, 16.Google Scholar
Laubscher, N. F., Steffens, F. E. & Vrba, E. (1972). Statistical evaluation of the taxonomic status of a fossil member of the Bovidae (Mammalia: Artiodactyla). Annals of the Transvaal Museum, 28, 1726.Google Scholar
Laurillard, C. L. (1842). Antilope. In d’Orbigny, C. D., ed. Dictionnaire Universel d’Histoire Naturelle, Vol. 1. Paris: Bureau Principal des Editeurs, pp. 612626.Google Scholar
Lavocat, R. (1952). Sur une faune de mammifères miocènes découverte à Beni-Mellal (Atlas Marocain). Comptes Rendus de l’Academie des Sciences, Paris, 235, 189191.Google Scholar
Lavocat, R. (1956). La faune des rongeurs des grottes à Australopithèques. Palaeontologia Africana, 4, 6975.Google Scholar
Lavocat, R. (1957). Sur l’âge des faunes de rongeurs des grottes à Australopithèques. In Clark, J. D. & Cole, S., eds. Third Pan-African Congress of Prehistory, Livingstone, 1955. Livingstone: Chatto & Windus, pp. 133134.Google Scholar
Lavocat, R. (1961). Le gisement de vertébrés Miocènes de Beni Mallal (Maroc), 2. Etude systématique de la faune de mammifères. Notes et Mémoires du Service Géologique du Maroc, 155, 2994.Google Scholar
Lavocat, R. (1973). Les Rongeurs du Miocene d’Afrique Orientale, 1. Miocene Inférieur. Mémoires et travaux de l’Institut de Montpellier de l’Ecole Pratique des Hautes Etudes, 1, 1284.Google Scholar
Lawes, M. J. (1990). The distribution of the samango monkey (Cercopithecus mitis erythrarchus Peters, 1852 and Cercopithecus mitis labiatus I. Geoffroy, 1843) and the forest history in southern Africa. Journal of Biogeography, 17, 669680.Google Scholar
Le Grange, A., Bastos, A. D. S., Brettschneider, H. & Chimimba, C. T. (2015). Evidence of a contact zone between two Rhabdomys dilectus (Rodentia: Muridae) mitotypes in Gauteng province, South Africa. African Zoology, 50, 6368.Google Scholar
Le Roux, A. & Badenhorst, S. (2016). Iron Age fauna from Sibudu Cave in KwaZulu-Natal, South Africa. Azania, 51, 307326.Google Scholar
Le Roux, A., Badenhorst, S., Esterhuysen, A. & Cain, C. (2013). Faunal remains from the 1854 siege of Mugombane, Makapans Valley, South Africa. Journal of African Archaeology, 11, 97110.Google Scholar
Leakey, L. S. B. (1942). Fossil Suidae of Oldoway. Journal of the East Africa Natural History Society, 16, 178196.Google Scholar
Leakey, L. S. B. (1943a). Notes on Simopithecus oswaldi Andrews from the type site. Journal of the East Africa Natural History Society, 17, 3944.Google Scholar
Leakey, L. S. B. (1943b). New fossil Suidae from Shungura, Omo. Journal of the East Africa Natural History Society, 17, 4561.Google Scholar
Leakey, L. S. B. (1958). Some East African fossil Suidae. Fossil Mammals of Africa, 14, 1133.Google Scholar
Leakey, L. S. B. (1961). A new Lower Pliocene fossil primate from Kenya. Annals and Magazine of Natural History, Series 13, 4, 689696.Google Scholar
Leakey, L. S. B. (1965). Olduvai Gorge, Vol. 1. Cambridge: Cambridge University Press.Google Scholar
Leece, A. B., Kegley, A. D. T., Lacruz, R. S., et al. (2016). The first hominin from the early Pleistocene paleocave of Haasgat, South Africa. PeerJ, 4. DOI: 10.7717/peerj.2024. eCollection 2016.Google Scholar
Lehmann, T. (2004). Fossil aardvark (Orycteropus) from Swartkrans Cave, South Africa. South African Journal of Science, 100, 311314.Google Scholar
Lehmann, T. (2007). Amended taxonomy of the order Tubulidentata (Mammalia, Eutheria). Annals of the Transvaal Museum, 44, 179196.Google Scholar
Lehmann, T. (2009). Phylogeny and systematics of the Orycteropodidae (Mammalia, Tubulidentata). Zoological Journal of the Linnean Society, 155, 649702.Google Scholar
Lesson, R. P. (1826). Voyage autour du Monde sur la Coquille pendant 1822. Paris: Bertrand.Google Scholar
Lesson, R. P. (1827). Manuel de Mammalogie ou l’Histoire Naturelle des Mammifères. Paris: Roret.Google Scholar
Lesson, R. P. (1842). Nouveau Tableau du Règne Animal. Paris: Bertrand.Google Scholar
Lewis, M. E. & Morlo, M. (2010). Creodonta. In Werdelin, L. & Sanders, W. J., eds. Cenozoic Mammals of Africa. Berkeley, CA: University of California Press, pp. 543560.Google Scholar
Lewis, P. J., Brink, J. S., Kennedy, A. M. & Campbell, T. L. (2011). Examination of the Florisbad microvertebrates. South African Journal of Science, 107, 14.Google Scholar
Lichtenstein, W. H. C. (1812a). Reisen im sűdlichen Africa in en Jahren 1803, 1804, 1805 und 1806. Berlin: Salfeld.Google Scholar
Lichtenstein, W. H. C. (1812b). Die Gattung Antilope. Sitzungsberichte der Gesellschaft Naturforschender Freunde zu Berlin, 6, 147160.Google Scholar
Lichtenstein, W. H. C. (1823). Verzeichniss der Doubletten des zoologischen Museums der Königl. Universität zu Berlin. Berlin: Trautwein.Google Scholar
Lichtenstein, W. H. C. (1835). Ueber Lutra maculicollis Lichtenst: aus dem Kafferlande. Archiv für Naturgeschichte, 1, 8992.Google Scholar
Lihoreau, F. & Ducrocq, S. (2007). Family Anthracotheriidae. In Prothero, D. R. & Foss, S. E., eds. The Evolution of Artiodactyls. Baltimore, MD: Johns Hopkins University Press, pp. 89105.Google Scholar
Lindsay, E. H. (1988). Cricetid rodents from Siwalik deposits near Chinji village: part I. Megacricetodontinae, Myocricetodontinae and Dendromurinae. Palaeovertebrata, 18, 95154.Google Scholar
Linnaeus, C. (1758). Systema Naturae Regnum Animale, 10th edition. Leipzig: Engelmann.Google Scholar
Lockwood, C. A. & Tobias, P. V. (1999). A large male cranium from Sterkfontein, South Africa and the status of Australopithecus africanus. Journal of Human Evolution, 36, 637685.Google Scholar
Lockwood, C. A. & Tobias, P. V. (2002). Morphology and affinities of new hominin cranial remains from Member 4 of the Sterkfontein Formation, Gauteng Province, South Africa. Journal of Human Evolution, 42, 389450.Google Scholar
Loubser, J. H. N. (1985). Buffelshoek: an ethnoarchaeological consideration of a Late Iron Age settlement in the southern Transvaal. South African Archaeological Bulletin, 40, 8187.Google Scholar
Loubser, J. H. N. (1994). Ndebele archaeology of the Petersburg area. Navorsinge van die Nasionale Museum Bloemfontein, 10, 61147.Google Scholar
Louw, J. T. (1960). Prehistory of the Matjes River Rock Shelter. Bloemfontein: National Museum Bloemfontein.Google Scholar
Low, M. E. Y. & Evenhuis, N. L. (2014). Additional dates of Sir Andrew Smith’s illustrations of the zoology of South Africa. Zootaxa, 3795, 483488.Google Scholar
Lundholm, B. (1951). A skull of the true quagga (Equus quagga) in the collection of the Transvaal Museum. South African Journal of Science, 47, 307312.Google Scholar
Lundholm, B. G. (1952). Equus zebra greatheadi, n. subsp., a new South African fossil zebra. Annals of the Transvaal Museum, 22, 2527.Google Scholar
Lundholm, B. G. (1954). A taxonomic study of Cynictis penicillata (G. Cuvier). Annals of the Transvaal Museum, 22, 305319.Google Scholar
Lynch, C. D. (1981). The status of the Cape grey mongoose Herpestes pulverulentus Wagner 1839 (Mammalia: Viverridae). Navorsinge van die Nasionale Museum Bloemfontein, 4, 121168.Google Scholar
Lyon, M. W. (1904). Classification of the hares and their allies. Smithsonian Miscellaneous Collections, 45, 321447.Google Scholar
MacFadden, B. J. (1984). Systematics and phylogeny of Hipparion, Neohipparion, Nannippus and Cormohipparion (Mammalia, Equidae) from the Miocene and Pliocene of the New World. Bulletin of the American Museum of Natural History, 179, 1195.Google Scholar
MacInnes, D. G. (1936). A new genus of fossil deer from the Miocene of Africa. Zoological Journal of the Linnean Society, 39, 521530.Google Scholar
MacInnes, D. G. (1942). Miocene and post-Miocene Proboscidia [sic] from East Africa. Transactions of the Zoological Society of London, 25, 33106.Google Scholar
MacInnes, D. G. (1951). Miocene Anthracotheriidae from East Africa. Fossil Mammals of Africa, 4, 124.Google Scholar
Mackay, A., Jacobs, Z. & Steele, T. E. (2015). Pleistocene archaeology and chronology of Putslaagte 8 (PL8) Rockshelter, Western Cape, South Africa. Journal of African Archaeology, 13, 7198.Google Scholar
MacLatchy, L., DeSilva, J. M., Sanders, W. J. & Wood, B. (2010). Hominini. In Werdelin, L. & Sanders, W. J., eds. Cenozoic Mammals of Africa. Berkeley, CA: University of California Press, pp. 473540.Google Scholar
Madden, C. T., Schmidt, D. L. & Whitmore, F. C. (1983). Mastritherium (Artiodactyla, Anthracotheriidae) from Wadi Sabya, southwestern Saudi Arabia: an earliest Miocene age for continental Rift-Valley volcanic deposits of the Red Sea margin. US Department of the Interior Geological Survey Report 83–83-488. Washington, DC: USGS.Google Scholar
Maggs, T. M. O. C. (1975). Faunal remains and hunting patterns from the Iron Age of the southern highveld. Annals of the Natal Museum, 22, 449459.Google Scholar
Maggs, T. M. O. C. & Michael, M. A. (1976). Ntshekane: an Early Iron Age site in the Tugela Basin, Natal. Annals of the Natal Museum, 22, 705740.Google Scholar
Maglio, V. J. & Hendey, Q. B. (1970). New evidence relating to the supposed stegolophodont ancestry of the Elephantidae. South African Archaeological Bulletin, 25, 8587.Google Scholar
Maguire, J. M. (1976). A taxonomic and ecological study of the living and fossil Hystricidae with particular reference to southern Africa. Unpublished 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: Liss, pp. 151164.Google Scholar
Maier, W. (1970). New fossil Cercopithecoidea from the lower Pleistocene cave deposits of the Makapansgat Limeworks, South Africa. Palaeontologia Africana, 13, 69107.Google Scholar
Maier, W. (1971). Two new skulls of Parapapio antiquus from Taung and a suggested phylogenetic arrangement of the genus Parapapio. Annals of the South African Museum, 59, 116.Google Scholar
Maier, W. (1972). The first complete skull of Simopithecus darti from Makapansgat, South Africa and its systematic position. Journal of Human Evolution, 1, 395405.Google Scholar
Malan, B. D. (1970). Remarks and reminiscences on the history of archaeology in South Africa. South African Archaeological Bulletin, 25, 8892.Google Scholar
Malan, B. D. & Cooke, H. B. S. (1941). A preliminary account of the Wonderwerk Cave, Kuruman District, South Africa. South African Journal of Science, 37, 300312.Google Scholar
Malan, B. D. & Wells, L. H. (1943). A further report on the Wonderwerk Cave, Kuruman. South African Journal of Science, 40, 258270.Google Scholar
Manhire, A. (1993). A report on the excavations at Faraoskop Rock Shelter in the Graafwater District of the south-western Cape. Southern African Field Archaeology, 2, 323.Google Scholar
Manthi, F. K. (2002). Saldanha Bay Yacht Club micromammals. Unpublished MA thesis, University of Cape Town.Google Scholar
Manyanga, M. (2001). Choices and constraints: animal resource exploitation in south-eastern Zimbabwe c. AD 900–1500. Studies in African Archaeology, 18, 1139.Google Scholar
Manyanga, M. (2006). Resilient landscapes. Unpublished PhD thesis, Uppsala University.Google Scholar
Manyanga, M., Pikirayi, I. & Ndoro, W. (2000). Coping with dryland environments: preliminary results from Mapungubwe and Zimbabwe phase sites in the Mateke Hills, south-eastern Zimbabwe. South African Archaeological Society Goodwin Series, 8, 6977.Google Scholar
Marean, C. W. (1985). The faunal remains from Smitswinkelbaai Cave, Cape Peninsula. South African Archaeological Bulletin, 40, 100102.Google Scholar
Marean, C. W., Abe, Y., Frey, C. J. & Randall, R. C. (2000). Zooarchaeological and taphonomic analysis of the Die Kelders Cave 1 layers 10 and 11 Middle Stone Age larger mammal fauna. Journal of Human Evolution, 38, 197233.Google Scholar
Marean, C. W., Nilssen, P. J., Brown, K. S., Jerardino, A. & Stynder, D. D. (2004). Palaeoanthropological investigations of Middle Stone Age sites at Pinnacle Point, Mossel Bay (South Africa): archaeology and hominid remains from the 2000 field season. Paleoanthropology, 1, 1483.Google Scholar
Maree, S. (2002). Phylogenetic relationships and mitochondrial DNA sequence evolution in the African rodent subfamily Otomyinae (Muridae). Unpublished PhD thesis, University of Pretoria.Google Scholar
Maree, S. & Faulkes, C. (2016). Cryptomys hottentotus. The IUCN Red List of Threatened Species, 2016, e.T5755A115079767. DOI: 10.2305/IUCN.UK.2016-3.RLTS.T5755A22185187.en.Google Scholar
Maree, S., Visser, J., Bennett, N. C. & Jarvis, J. (2017). Georychus capensis. The IUCN Red List of Threatened Species, 2017, e.T9077A110019425. DOI: 10.2305/IUCN.UK.2017-2.RLTS.T9077A110019425.en.Google Scholar
Marivaux, L., Adaci, M., Bensalah, M., et al. (2011). Zegdoumyidae (Rodentia, Mammalia), stem anomaluroid rodents from the Early to Middle Eocene of Algeria (Gour Lazib, Western Sahara): new dental evidence. Journal of Systematic Palaeontology, 9, 563588.Google Scholar
Marivaux, L., Essid, E. M., Marzougui, W., et al. (2014). A new and primitive species of Protophiomys (Rodentia, Hystricognathi) from the late middle Eocene of Djebel el Kébar, Central Tunisia. Palaeovertebrata, 38, 117.Google Scholar
Marivaux, L., Essid, E. M., Marzougui, W., et al. (2015). The early evolutionary history of anomaluroid rodents in Africa: new dental remains of a zegdoumyid (Zegdoumyidae, Anomaluroidea) from the Eocene of Tunisia. Zoologica Scripta, 44, 117134.Google Scholar
Martínez-Navarro, B., Pérez-Claros, J. A., Palombo, M. R., Rook, L. & Palmqvist, P. (2007). The Olduvai buffalo Pelorovis and the origin of Bos. Quaternary Research, 68, 220226.Google Scholar
Maswanganye, K. A., Cunningham, M. J., Bennett, N. C., Chimimba, C. T. & Bloomer, P. (2017). Life on the rocks: multilocus phylogeography of rock hyrax (Procavia capensis) from southern Africa. Molecular Phylogenetics and Evolution, 114, 4962.Google Scholar
Matthews, T. & Stynder, D. D. (2011a). An analysis of the Aethomys (Murinae) community from Langebaanweg (Early Pliocene, South Africa) using geometric morphometrics. Palaeogeography, Palaeoclimatology, Palaeoecology, 302, 230242.Google Scholar
Matthews, T. & Stynder, D. D. (2011b). An analysis of two Myosorex species (Soricidae) from the Early Pliocene site of Langebaanweg (West coast, South Africa) using geometric morphometrics, linear measurements and non-metric characters. Geobios, 44, 8799.Google Scholar
Matthews, T., Denys, C. & Parkington, J. E. (2005). The palaeoecology of the micromammals from the late middle Pleistocene site of Hoedjiespunt 1 (Cape Province, South Africa). Journal of Human Evolution, 49, 432451.Google Scholar
Matthews, T., Parkington, J. E. & Denys, C. (2006). The taphonomy of the micromammals from the Late Middle Pleistocene site of Hoedjiespunt 1 (Cape Province, South Africa). Journal of Taphonomy, 4, 1126.Google Scholar
Matthews, T., Denys, C. & Parkington, J. E. (2007). Community evolution of Neogene micromammals from Langebaanweg ‘E’ Quarry and other west coast fossil sites, south-western Cape, South Africa. Palaeogeography, Palaeoclimatology, Palaeoecology, 245, 332352.Google Scholar
Matthews, T., Marean, C. & Nilssen, P. J. (2009). Micromammals from the Middle Stone Age (92–167 ka) at Cave PP13B, Pinnacle Point, south coast, South Africa. Palaeontologia Africana, 44, 112120.Google Scholar
Matthews, T., Rector, A., Jacobs, Z., Herries, A. I. R. & 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, 213229.Google Scholar
Mazel, A. D. (1984a). Gehle Shelter: report on excavations in the uplands ecological zone, Tugela Basin, Natal, South Africa. Annals of the Natal Museum, 26, 124.Google Scholar
Mazel, A. D. (1984b). Diamond I and Clarke’s Shelter: report on excavations in the northern Drakensberg, Natal, South Africa. Annals of the Natal Museum, 26, 2570.Google Scholar
Mazel, A. D. (1986a). Mgede Shelter: a mid- and late Holocene observation in the western Biggarsberg, Thukela Basin, South Africa. Annals of the Natal Museum, 27, 357387.Google Scholar
Mazel, A. D. (1986b). Mbabane Shelter and eSinhlonhweni Shelter: the last two thousand years of hunter-gatherer settlement in the central Thukela Basin, Natal, South Africa. Annals of the Natal Museum, 27, 389453.Google Scholar
Mazel, A. D. (1988a). Nkupe Shelter: report on excavations in the eastern Biggarsberg, Thukela Basin, Natal, South Africa. Annals of the Natal Museum, 29, 321377.Google Scholar
Mazel, A. D. (1988b). Sikhanyisweni Shelter: report on excavations in the Thukela Basin, Natal, South Africa. Annals of the Natal Museum, 29, 379406.Google Scholar
Mazel, A. D. (1990). Mhlawazini Cave: the excavation of Late Holocene deposits in the northern Natal Drakensberg, Natal, South Africa. Natal Museum Journal of Humanities, 2, 95133.Google Scholar
Mazel, A. D. (1992). Collingham Shelter: the excavation of late Holocene deposits, Natal, South Africa. Natal Museum Journal of Humanities, 4, 151.Google Scholar
Mazel, A. D. (1993). KwaThwaleyakhe Shelter: the excavation of mid and late Holocene deposits in the central Thukela Basin, Natal, South Africa. Natal Museum Journal of Humanities, 5, 136.Google Scholar
McCrae, C. & Potze, S. (2006). Analysis of microfauna-bearing breccia from Kromdraai A in the Cradle of Humankind World Heritage Site. Annals of the Transvaal Museum, 43, 107110.Google Scholar
McCrossin, M. L. (1992). Human molars from Later Pleistocene deposits of Witkrans Cave, Gaap Escarpment, Kalahari Margin. Human Evolution, 7, 110.Google Scholar
McCrossin, M. L. & Benefit, B. R. (1997). On the relationships and adaptations of Kenyapithecus, a large-bodied hominoid from the Middle Miocene of eastern Africa. In Begun, D. R., Ward, C. V. & Rose, M. D., eds. Function, Phylogeny and Fossils: Miocene Hominoid Evolution and Adaptations. New York: Plenum, pp. 241267.Google Scholar
McGrath, J. R., Cleghorn, N., Gennari, B., et al. (2015). The Pinnacle Point shell midden complex: a Mid- to Late Holocene record of Later Stone Age coastal foraging along the southern Cape coast of South Africa. South African Archaeological Bulletin, 70, 209219.Google Scholar
McKee, J. K. (1991). Palaeo-ecology of the Sterkfontein hominids: a review and synthesis. Palaeoecology of Africa, 28, 4151.Google Scholar
McKee, J. K. (1993a). Faunal dating of the Taung fossil hominid deposit. Journal of Human Evolution, 25, 363376.Google Scholar
McKee, J. K. (1993b). Taxonomic and evolutionary affinities of Papio izodi fossils from Taung and Sterkfontein. Palaeontologia Africana, 30, 4349.Google Scholar
McKee, J. K. (1994). Catalogue of fossil sites at the Buxton Limeworks, Taung. Palaeontologia Africana, 31, 7381.Google Scholar
McKee, J. K. & Keyser, A. W. (1994). Craniodental remains of Papio angusticeps from the Haasgat cave site, South Africa. International Journal of Primatology, 15, 823841.Google Scholar
McKee, J. K., Thackeray, J. F. & Berger, L. R. (1995). Faunal assemblage seriation of southern African Pliocene and Pleistocene fossil deposits. American Journal of Physical Anthropology, 96, 235250.Google Scholar
McKee, J. K., Von Mayer, A. & Kuykendall, K. I. (2011). New species of Cercopithecoides from Haasgat, North West Province, South Africa. Journal of Human Evolution, 60, 8393.Google Scholar
McKenna, M. C. & Bell, S. K. (1997). Classification of Mammals above the Species Level. New York: Columbia University Press.Google Scholar
McMahon, C. R. & Thackeray, J. F. (1994). Plio-Pleistocene Hyracoidea from Swartkrans Cave, South Africa. South African Journal of Zoology, 29, 4045.Google Scholar
Meester, J. (1953a). The genera of African shrews. Annals of the Transvaal Museum, 22, 205214.Google Scholar
Meester, J. (1953b). A new golden mole from Spitzkop, Sabie. South African Journal of Science, 49, 207208.Google Scholar
Meester, J. (1954). Fossil shrews of South Africa. Annals of the Transvaal Museum, 22, 271278.Google Scholar
Meester, J. (1958). Variation in the shrew genus Myosorex in southern Africa. Journal of Mammalogy, 39, 325339.Google Scholar
Meester, J. (1961a). Appendix A: report on the fossil shrews in the University of California collections from the South African cave breccias. Transvaal Museum unpublished report, 2 pp.Google Scholar
Meester, J. (1961b). A taxonomic revision of southern African Crocidura (Mammalia: Insectivora). Annals and Magazine of Natural History, Series 13, 4, 561571.Google Scholar
Meester, J. (1963). A systematic revision of the shrew genus Crocidura in southern Africa. Transvaal Museum Memoir, 13, 1127.Google Scholar
Meester, J. (1964a). Revision of the Chrysochloridae: 1. The desert golden mole, Eremitalpa Roberts. Scientific Papers of the Namib Desert Research Station, 26, 19.Google Scholar
Meester, J. (1964b). The status of Crocidura mariquensis (Smith) (Mammalia: Insectivora). Puku, 2, 7880.Google Scholar
Meester, J. (1980). Appendix 2: small mammals. In Maggs, T. M. & Ward, V.. Driel Shelter: rescue at a Late Stone Age site on the Tugela River. Annals of the Natal Museum, 24, 67.Google Scholar
Meester, J. & Lambrechts, A. von W. (1971). The southern African species of Suncus Ehrenberg (Mammalia: Soricidae). Annals of the Transvaal Museum, 27, 114.Google Scholar
Meester, J. & Meyer, I. J. (1972). Fossil Suncus (Mammalia: Soricidae) from southern Africa. Annals of the Transvaal Museum, 27, 269277.Google Scholar
Meester, J., Richardson, E. J. & Work, D. (1985). Multivariate analysis of southern African Crocidura (Soricidae). Acta Zoologica Fennica, 173, 219222.Google Scholar
Meester, J., Taylor, P. J., Confratto, G.-C., et al. (1992). Chromosomal speciation in southern African Otomyinae (Rodentia: Muridae): a review. Durban Museum Novitates, 17, 5863.Google Scholar
Mein, P. & Pickford, M. (2003a). Fossil bat (Microchiroptera, Mammalia) from Arrisdrift, Namibia. Memoirs of the Geological Survey of Namibia, 19, 115117.Google Scholar
Mein, P. & Pickford, M. (2003b). Insectivora from Arrisdrift, a basal Middle Miocene locality in southern Namibia. Memoirs of the Geological Survey of Namibia, 19, 143146.Google Scholar
Mein, P. & Pickford, M. (2003c). Rodentia (other than Pedetidae) from the Orange River deposits, Namibia. Memoirs of the Geological Survey of Namibia, 19, 147160.Google Scholar
Mein, P. & Pickford, M. (2003d). Fossil picas (Ochotonidae, Lagomorpha, Mammalia) from the basal Middle Miocene of Arrisdrift, Namibia. Memoirs of the Geological Survey of Namibia, 19, 171176.Google Scholar
Mein, P. & Pickford, M. (2008a). Early Miocene insectivores from the northern Sperrgebiet, Namibia. Memoirs of the Geological Survey of Namibia, 20, 169183.Google Scholar
Mein, P. & Pickford, M. (2008b). Early Miocene Lagomorpha from the northern Sperrgebiet, Namibia. Memoirs of the Geological Survey of Namibia, 20, 227233.Google Scholar
Mein, P. & Pickford, M. (2008c). Early Miocene Rodentia from the northern Sperrgebiet, Namibia. Memoirs of the Geological Survey of Namibia, 20, 235290.Google Scholar
Mein, P. & Pickford, M. (2018). Reithroparamyine rodent from the Eocene of Namibia. Communications of the Geological Survey of Namibia, 18, 3847.Google Scholar
Mein, P. & Senut, B. (2003). The Pedetidae from the Miocene site of Arrisdrift (Namibia). Memoirs of the Geological Survey of Namibia, 19, 161170.Google Scholar
Mein, P., Pickford, M. & Senut, B. (2000a). Late Miocene micromammals from the Harasib karst deposits, Namibia. Part 1 – large muroids and non-muroid rodents. Communications of the Geological Survey of Namibia, 12, 375390.Google Scholar
Mein, P., Pickford, M. & Senut, B. (2000b). Late Miocene micromammals from the Harasib karst deposits, Namibia. Part 2a – Myocricetodontinae, Petromyscinae and Namibimyinae (Muridae, Gerbillidae). Communications of the Geological Survey of Namibia, 12, 391401.Google Scholar
Mein, P., Pickford, M. & Senut, B. (2004). Late Miocene micromammals from the Harasib karst deposits, Namibia. Part 2b – Cricetomyidae, Dendromuridae and Muridae, with an addendum on the Myocricetodontinae. Communications of the Geological Survey of Namibia, 13, 4361.Google Scholar
Meiring, A. J. D. (1955). Fossil proboscidean teeth and ulna from Virginia, O.F.S. Navorsinge van die Nasionale Museum Bloemfontein, 1, 187201.Google Scholar
Mendrez, C. (1966). On Equus (Hippotigris) cf. burchelli (Gray) from ‘Sterkfontein extension’, Transvaal, South Africa. Annals of the Transvaal Museum, 25, 9197.Google Scholar
Menter, C., Kuykendall, K. I., Keyser, A. W. & 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
Miller, E. R., Gunnell, G. F., Gawad, M. A., et al. (2014). Anthracotheres from Wadi Moghra, early Miocene, Egypt. Journal of Paleontology, 88, 967981.Google Scholar
Miller-Butterworth, C. M., Eick, G., Jacobs, D. S., Schoeman, M. C. & Harley, E. C. (2005). Genetic and phenotypic differences between South African long-fingered bats, with a global miniopterine phylogeny. Journal of Mammalogy, 86, 11211135.Google Scholar
Miller-Butterworth, C. M., Murphy, W. J., O’Brien, S. J., et al. (2007). A family matter: conclusive resolution of the taxonomic position of the long-fingered bats, Miniopterus. Molecular Biology and Evolution, 24, 15531561.Google Scholar
Mills, M. G. L. (1982). Hyaena brunnea. Mammalian Species, 194, 15.Google Scholar
Mitchell, P. J. (1993). The archaeology of Tloutle rockshelter, Maseru District, Lesotho. Navorsinge van die Nasionale Museum Bloemfontein, 9, 77132.Google Scholar
Mitchell, P. J., Parkington, J. E. & Yates, R. (1994). Recent Holocene archaeology in western and southern Lesotho. South African Archaeological Bulletin, 49, 3352.Google Scholar
Moggi-Cecchi, J., Tobias, P. V. & Benyon, A. D. (1998). The mixed dentition and associated skull fragments of a juvenile fossil hominid from Sterkfontein, South Africa. American Journal of Physical Anthropology, 106, 425466.Google Scholar
Moggi-Cecchi, J., Grine, F. E. & Tobias, P. V. (2006). Early hominid dental remains from Members 4 and 5 of the Sterkfontein Formation (1966–1996 excavations): catalogue, individual associations, morphological descriptions and initial metrical analysis. Journal of Human Evolution, 50, 239328.Google Scholar
Moggi-Cecchi, J., Menter, C., Boccone, S. & Keyser, A. (2010). Early hominin dental remains from the Plio-Pleistocene site of Drimolen, South Africa. Journal of Human Evolution, 58, 374405.Google Scholar
Mollett, O. D.. (1947). Fossil mammals from the Makapan Valley, Potgietersrust: I. Primates. South African Journal of Science, 43, 295303.Google Scholar
Montgelard, C., Matthee, C. A. & Robinson, T. J. (2003). Molecular systematics of dormice (Rodentia, Gliridae) and the radiation of Graphiurus in Africa. Proceedings of the Royal Society of London Series B, 270, 19471955.Google Scholar
Montoya, P., Morales, J. & Abella, J. (2011). Musteloidea (Carnivora, Mammalia) from the Late Miocene of Venta del Moro (Valencia, Spain). Estudios Geológicos, 67, 193206.Google Scholar
Morales, J. & Pickford, M. (2018). A new barbourofelid mandible (Carnivora, Mammalia) from the Early Miocene of Grillental-6, Sperrgebiet, Namibia. Communications of the Geological Survey of Namibia, 18, 113123.Google Scholar
Morales, J., Soria, D. & Pickford, M. (1995). Sur les origines de la famille des Bovidae (Artiodactyla, Mammalia). Comptes Rendus de l’Académie des Sciences de Paris, Sciences de la Terre et des Planètes, 321, 12111217.Google Scholar
Morales, J., Pickford, M. & Soria, D. (1998a). A new creodont Metapterodon stromeri nov. sp. (Hyaenodontidae, Mammalia) from the Early Miocene of Langental (Sperrgebiet, Namibia). Comptes Rendus de l’Académie des Sciences de Paris, Sciences de la Terre et des Planètes, 327, 633638.Google Scholar
Morales, J., Pickford, M., Soria, D. & Fraile, S. (1998b). New carnivores from the basal Middle Miocene of Arrisdrift, Namibia. Eclogae Geologicae Helvetiae, 91, 2740.Google Scholar
Morales, J., Soria, D. & Pickford, M. (1999). New stem giraffoid ruminants from the early and middle Miocene of Namibia. Geodiversitas, 21, 229253.Google Scholar
Morales, J., Pickford, M., Soria, D. & Fraile, S. (2001a). New Viverrinae (Carnivora: Mammalia) from the basal Middle Miocene of Arrisdrift, Namibia. Palaeontologia Africana, 37, 99102.Google Scholar
Morales, J., Salesa, M. J., Pickford, M. & Soria, D. (2001b). A new tribe, new genus and two new species of Barbourofelinae (Felidae, Carnivora, Mammalia) from the Early Miocene of East Africa and Spain. Transactions of the Royal Society of Edinburgh: Earth Sciences, 92, 97102.Google Scholar
Morales, J., Pickford, M., Fraile, S., Salesa, M. & Soria, D. (2003a). Creodonta and Carnivora from Arrisdrift, early Middle Miocene of southern Namibia. Memoirs of the Geological Survey of Namibia, 19, 177194.Google Scholar
Morales, J., Soria, D., Pelaez-Campomanes, P. & Pickford, M. (2003b). New data regarding Orangemeryx hendeyi Morales et al., 2000, from the type locality, Arrisdrift, Namibia. Memoirs of the Geological Survey of Namibia, 19, 305344.Google Scholar
Morales, J., Soria, D., Pickford, M. & Nieto, M. (2003c). A new genus and species of Bovidae (Artiodactyla, Mammalia) from the early Middle Miocene of Arrisdrift, Namibia and the origins of the family Bovidae. Memoirs of the Geological Survey of Namibia, 19, 371384.Google Scholar
Morales, J., Soria, D., Sánchez, I. M., Quiralte, V. & Pickford, M. (2003d). Tragulidae from Arrisdrift, basal Middle Miocene, southern Namibia. Memoirs of the Geological Survey of Namibia, 19, 359369.Google Scholar
Morales, J., Pickford, M. & Soria, D. (2005). Carnivores from the late Miocene and basal Pliocene of the Tugen Hills, Kenya. Revista de la Sociedad Geológica de España, 18, 3961.Google Scholar
Morales, J., Pickford, M. & Salesa, M. J. (2008a). Creodonta and Carnivora from the early Miocene of the northern Sperrgebiet, Namibia. Memoirs of the Geological Survey of Namibia, 20, 291310.Google Scholar
Morales, J., Soria, D. & Pickford, M. (2008b). Pecoran ruminants from the Early Miocene of the Sperrgebiet, Namibia. Memoirs of the Geological Survey of Namibia, 20, 397464.Google Scholar
Morales, J., Senut, B. & Pickford, M. (2011). Crocuta dietrichi from Meob, Namibia: implications for the age of the Tsondab Sandstone in the coastal part of the Namib Desert. Estudios Geológicos, 67, 207215.Google Scholar
Morales, J., Pickford, M. & Valenciano, A. (2016). Systematics of African Amphicyonidae, with descriptions of new material from Napak (Uganda) and Grillental (Namibia). Journal of Iberian Geology, 42, 131150.Google Scholar
Morlo, M., Miller, E. R. & El-Barkooky, A. N. (2007). Creodonta and Carnivora from Wadi Moghra, Egypt. Journal of Vertebrate Paleontology, 27, 145159.Google Scholar
Morris, A. G. (1994). Appendix 1: human skeletal remains from Wosi – an Early Iron Age site in the Thukela Basin, Natal. In Van Schalkwyk, L.. Wosi: an Early Iron Age village in the lower Thukela Basin, Natal. Natal Museum Journal of Humanities, 6, 97104.Google Scholar
Morris, A. G., Louw, G. H., Van Wyk, E. & Cooper, C. (1995). A brief report on the rescue excavation on a human skeleton from Nooitgedacht, Northern Cape Province, South Africa. Southern African Field Archaeology, 4, 120123.Google Scholar
Morris, A. G., Dlamini, N., Joseph, J., et al. (2004/2005). Later Stone Age burials from the Western Cape Province, South Africa Part 1: Voëlvlei. Southern African Field Archaeology, 13–14, 1926.Google Scholar
Mullin, S. K., Pillay, N. & Taylor, P. J. (2004). Skull size and shape of Dasymys (Rodentia, Muridae) from sub-Saharan Africa. Mammalia, 68, 185220.Google Scholar
Mupira, P. & Katsamudanga, S. (2007). Excavations at Manjowe and Gwenzi rock shelters in Zimunya Communal Lands, central Eastern Highlands of Zimbabwe. Zimbabwea, 9, 2142.Google Scholar
Musser, G. G. & Carleton, M. D. (2005). Superfamily Muroidea. In Wilson, D. E. & Reeder, D. M., eds. Mammal Species of the World, Vol. 2. Baltimore, MD: Johns Hopkins University Press, pp. 8941599.Google Scholar
Mutter, R. J., Berger, L. R. & Schmid, P. (2001). New evidence of the giant hyaena, Pachycrocuta brevirostris (Carnivora, Hyaenidae), from the Gladysvale Cave deposit (Plio-Pleistocene, John Nash Nature Reserve, Gauteng, South Africa). Palaeontologia Africana, 37, 103113.Google Scholar
Mynhardt, S., Maree, S., Pelser, I., et al. (2015). Phylogeography of a morphologically cryptic golden mole assemblage from south-eastern Africa. PLoS ONE, 10. DOI: 10.1371/journal.pone.0144995.Google Scholar
Nelson, C. (2008). An archaeozoological and ethnographic investigation into animal utilisation practices of the Ndzundza Ndebele of the Steelpoort River Valley, South Africa, 1700 AD–1900 AD. Unpublished MA thesis, University of Pretoria.Google Scholar
Nelson, C. (2009). An archaeozoology of the Ndzundza Ndebele in the Steelpoort River valley. Mpumulanga, South Africa. c. 1700 AD – 1883 AD. South African Archaeological Bulletin, 64, 184192.Google Scholar
Noack, T. (1887). Beiträge zur Kenntniss der Säugethier-Fauna von Ost- und Central-Afrika. Zoologische Jahrbücher Systematik, 2, 193302.Google Scholar
Noli, D. (1988). Results of the 1986 excavation at Hailstone Midden (HSM), Eland’s Bay, Western Cape Province. South African Archaeological Bulletin, 43, 4348.Google Scholar
Nyamushosho, R. T. (2016). Living on the margin? The Iron Age communities of Mananzve Hill, Shashi region, South-western Zimbabwe. Unpublished MPhil thesis, University of Cape Town.Google Scholar
Ogilby, W. (1833). [Antilope ellipsyprymnus and Cynictis steedmani in the collection of Mr Steedman]. Proceedings of the Zoological Society of London, 1, 4749.Google Scholar
Ogilby, W. (1836). [On the generic characters of ruminants]. Proceedings of the Zoological Society of London, 4, 131139.Google Scholar
Ogilby, W. (1838). On a collection of Mammalia, procured by Captain Alexander during his journey into the country of the Damaras, on the south-west African coast. Proceedings of the Zoological Society of London, 6, 56.Google Scholar
Ogola, C. A. (2009). The Sterkfontein western breccias: stratigraphy, fauna and artefacts. Unpublished PhD thesis, University of the Witwatersrand.Google Scholar
Opperman, H. (1982). Some research results of excavations in the Colwinton Rock Shelter, north-eastern Cape. South African Archaeological Bulletin, 37, 5156.Google Scholar
Opperman, H. (1987). The Later Stone Age of the Drakensberg Range and its Foothills. Oxford: British Archaeological Reports.Google Scholar
Opperman, H. (1992). A report on the results of a test pit in Strathalan Cave B: Maclear District, north-eastern Cape. Southern African Field Archaeology, 1, 98102.Google Scholar
Opperman, H. (1996). Strathalan Cave B, north-western Cape Province, South Africa: evidence for human behaviour 29,000–26,000 years ago. Quaternary International, 33, 4553.Google Scholar
O’Regan, H. J. (2007). A revision of the Carnivora from Member 5, Sterkfontein, South Africa, based on a reassessment of published material and site stratigraphy. Annals of the Transvaal Museum, 44, 209214.Google Scholar
O’Regan, H. J. & Menter, C. G. (2009). Carnivora from the Plio-Pleistocene hominin site of Drimolen, Gauteng, South Africa. Geobios, 42, 329350.Google Scholar
O’Regan, H. J. & Steininger, C. (2017). Felidae from Cooper’s Cave, South Africa (Mammalia: Carnivora). Geodiversitas, 39, 315332.Google Scholar
O’Regan, H. J., Cohen, B. F. & Steininger, C. M. (2013). Mustelid and viverrid remains from the Pleistocene site of Cooper’s D, Gauteng, South Africa. Palaeontologia Africana, 48, 1923.Google Scholar
Orlando, L., Metcalf, J. L., Alberdi, M. T., et al. (2009). Revising the recent evolutionary history of equids using ancient DNA. Proceedings of the National Academy of Sciences of the United States of America, 106, 2175421759.Google Scholar
Orton, J. (2014). The late pre-colonial site of Komkans 2 (KK002) and an evaluation of the evidence for indigenous copper smelting in Namaqualand, southern Africa. Azania, 47, 386410.Google Scholar
Orton, J., Hart, T. & Halkett, D. (2005). Shell middens in Namaqualand: two Later Stone Age sites at Rooiwalbaai, Northern Cape Province, South Africa. South African Archaeological Bulletin, 60, 2432.Google Scholar
Orton, J., Klein, R. G., Mackay, A., Schwortz, S. & Steele, T. E. (2011). Two Holocene rock shelter deposits from the Knersvlakte, southern Namaqualand, South Africa. Southern African Humanities, 23, 109150.Google Scholar
Orton, J., Mitchell, P., Klein, R., Steele, T. & Horsburgh, K. A. (2013). An early date for cattle from Namaqualand, South Africa: implications for the origins of herding in southern Africa. Antiquity, 87, 108120.Google Scholar
Orton, J., Halkett, D., Hart, T., Patrick, M. & Pfeiffer, S. (2015). An unusual pre-colonial burial from Bloubergstrand, Table Bay, South Africa. South African Archaeological Bulletin, 70, 106112.Google Scholar
Osborn, H. F. (1928). Mammoths and man in the Transvaal. Nature, 121, 672673.Google Scholar
Osborn, H. F. (1934). Primitive Archidiskodon and Palaeoloxodon of South Africa. American Museum Novitates, 741, 115.Google Scholar
Osborn, H. F. (1942). Proboscidea. New York: American Museum of Natural History.Google Scholar
Osgood, W. H. (1910). Diagnoses of new East African mammals, including a new genus of Muridae. Publications of the Field Museum of Natural History, Zoological Series, 10, 513.Google Scholar
Owens, M. D. (1995). Archaeological research at Modipe Hill, Kgatleng District: the burials from 1994 excavations. Botswana Notes and Records, 27, 4148.Google Scholar
Pallas, P. S. (1766). P.S. Pallas Medecinae Doctoris Miscellanea Zoologica. La Haye: Van Cleef.Google Scholar
Pallas, P. S. (1767). De Antilopibus generatim. Spicilegia Zoologica, 1, 144.Google Scholar
Pallas, P. S. (1778). Novae Species Quadrupedum e Glirium Ordine. Erlangen, Walther.Google Scholar
Pappas, L. A. (2002). Taurotragus oryx. Mammalian Species, 689, 15.Google Scholar
Partridge, T. C. (1966). Ficus Cave: an Iron Age living site in the central Transvaal. South African Archaeological Bulletin, 21, 125132.Google Scholar
Partridge, T. C. (2010). Tectonics and geomorphology of Africa during the Phanerozoic. In Werdelin, L. & Sanders, W. J., eds. Cenozoic Mammals of Africa. Berkeley, CA: University of California Press, pp. 317.Google Scholar
Patterson, B. (1965). The fossil elephant shrews (family Macroscelididae). Bulletin of the Museum of Comparative Zoology Harvard, 133, 295335.Google Scholar
Pavlinov, I. J. A. (2001). Current concepts in gerbillid phylogeny and classification. In Denys, C., Granjon, L. & Poulet, A., eds. Proceedings of the 8th International Symposium on African Small Mammals. Paris: I.R.D., pp. 141149.Google Scholar
Pearce, D. G. (2008). Later Stone Age burial practice in the Eastern Cape Province, South Africa. Unpublished DPhil thesis, University of the Witwatersrand.Google Scholar
Pearson, O. M. & Grine, F. E. (1996). Morphology of the Border Cave hominid ulna and humerus. South African Journal of Science, 92, 231236.Google Scholar
Pellatt, A. (1972). Appendix: report on skeletal material from Ruanga. In Garlake, P. S.. Excavations at the Nhunguza and Ruanga Ruins in northern Mashonaland. South African Archaeological Bulletin, 27, 139140.Google Scholar
Pelser, A. J., Teichert, F. & Steyn, M. (2004/2005). A Late Iron Age/Contact Period burial at Stand 1610, Hillside Street, Silver Lakes, Tshwane. Southern African Field Archaeology, 13–14, 2735.Google Scholar
Penzhorn, B. L. (1988). Equus zebra. Mammalian Species, 314, 17.Google Scholar
Péringuey, L. (1911). The Stone Ages of South Africa as represented in the collection of the South African Museum. Annals of the South African Museum, 8, 1201.Google Scholar
Perrin, M. R. & Fielden, L. J. (1999). Eremitalpa granti. Mammalian Species, 629, 14.Google Scholar
Perrin, M. R., Dempster, E. R. & Downs, C. T. (1999). Gerbillurus paeba. Mammalian Species, 606, 16.Google Scholar
Perry, G. (1811). Arcana, or, the Museum of Natural History: Containing the Most Recent Discovered Objects. London: Smicton.Google Scholar
Peters, J. & Brink, J. S. (1992). Comparative postcranial osteomorphology and osteometry of springbok, Antidorcas marsupialis (Zimmerman, 1780) and Grey rhebok, Pelea capreolus (Forster, 1790) (Mammalia, Bovidae). Navorsinge van die Nasionale Museum Bloemfontein, 8, 162207.Google Scholar
Peters, W. C. H. (1846). Über neue Säugethiergattungen aus den Ordnungen der Insectenfresser und Nagelthiere. Bericht über die zur Bekanntmachung geeigneten Verhandlungen der Koniglichen Preussische Akademie der Wissenschaften zu Berlin, 1846, 257259.Google Scholar
Peters, W. C. H. (1851). Derselbe machte eine Mittheilung über zwei neue Insectivoren aus Mossambique. Bericht über die zur Bekanntmachung geeigneten Verhandlungen der Koniglichen Preussische Akademie der Wissenschaften zu Berlin, 1851, 467468.Google Scholar
Peters, W. C. H. (1852a). Einige neue Säugethiere und Flussfische aus Mossambique vor. Bericht über die zur Bekanntmachung geeigneten Verhandlungen der Koniglichen Preussische Akademie der Wissenschaften zu Berlin, 1852, 273276.Google Scholar
Peters, W. C. H. (1852b). Naturwissenschaftliche Reise nach Mossambique. Berlin: Reimer.Google Scholar
Peters, W. C. H. (1866). Über einige neue oder weniger bekannte Flederthiere. Monatsberichte der Königlichen Preussische Akademie des Wissenschaften zu Berlin, 1866, 1625.Google Scholar
Peters, W. C. H. (1875). Über Dasymys, eine neue Gattung von murinen Nagethieren aus Südafrika. Monatsberichte der Königlichen Preussische Akademie des Wissenschaften zu Berlin, 1875, 1214.Google Scholar
Peters, W. C. H. (1878). Über die von Hrn. J.M. Hildebrandt während seiner letzen ostafrikaischen Reise gesammelten Säugethiere und Amphibien. Monatsberichte der Königlichen Preussische Akademie des Wissenschaften zu Berlin, 1878, 194209.Google Scholar
Pether, J. (1994). The sedimentology, palaeontology and stratigraphy of coastal plain deposits at Hondeklip Bay, Namaqualand, South Africa. Unpublished MSc thesis, University of Cape Town.Google Scholar
Petter, F. (1967). Particularités dentaires des Petromyscinae Roberts 1951 (Rongeurs, Cricetides). Mammalia, 31, 217224.Google Scholar
Petter, F. (1981). Remarques sur la systématique des Chrysochloridés. Mammalia, 45, 4953.Google Scholar
Petter, G. (1963). Étude de quelques Viverridés (Mammifères, Carnivores) du Pléistocène inférìeur du Tanganyika (Afrique orientale). Bulletin de la Société Géologique de France, S7-V, 265274.Google Scholar
Petter, G. (1987). Small carnivores (Viverridae, Mustelidae, Canidae) from Laetoli. In Leakey, M. D. & Harris, J. M., eds. Laetoli: A Pliocene Site in Northern Tanzania. Oxford: Clarendon Press, pp. 194234.Google Scholar
Petter, G. & Howell, F. C. (1989). Une nouvelle espèce du genre Crocuta Kaup (Mammalia: Carnivora: Hyaenidae) dans la faune pliocène de Laetoli (Tanzanie): Crocuta dietrichi nov. sp.; origine du genre. Comptes Rendus de l’Académie des Sciences, Paris, Série II, 308, 10311038.Google Scholar
Pfeiffer, S. (2013). Population dynamics in the Southern African Holocene: Human burials from the West Coast. In Jerardino, A., Malan, A. & Braun, D., eds. The Archaeology of the West Coast of South Africa. Oxford: Archaeopress, pp. 143154.Google Scholar
Pfeiffer, S. & Van der Merwe, N. J. (2004). Cranial injuries to Later Stone Age children from the Modder River Mouth, Western Cape Province, South Africa. South African Archaeological Bulletin, 59, 5965.Google Scholar
Pia, J. (1930). Eine neue quartäre Warzenschwein art aus Südwestafrika. Centralblatt für Mineralogie, Geologie und Paläontologie, 1930 B, 7683.Google Scholar
Pickering, R., Dirks, P. H. G. M., Jinnah, Z., et al. (2011). Australopithecus sediba at 1.977 Ma and implications for the origins of the genus Homo. Science, 333, 14211423.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
Pickford, M. (1975). New fossil Orycteropodidae (Mammalia, Tubulidentata) from East Africa: Orycteropus minutus sp. nov. and Orycteropus chemeldoi sp. nov. Netherlands Journal of Zoology, 25, 5788.Google Scholar
Pickford, M. (1984). A revision of the Sanitheriidae (Suiformes, Mammalia). Geobios, 17, 133154.Google Scholar
Pickford, M. (1986). A revision of the Miocene Suidae and Tayassuidae (Artiodactyla, Mammalia) of Africa. Tertiary Research Special Paper, 7, 183.Google Scholar
Pickford, M. (1987). Miocene Suidae from Arrisdrift, South West Africa – Namibia. Annals of the South African Museum, 97, 283314.Google Scholar
Pickford, M. (1988). Un étrange suide nain du Neogene supérieur de Langebaanweg. Annales de Paléontologie, 74, 229250.Google Scholar
Pickford, M. (1990). Some fossiliferous Plio-Pleistocene cave systems of Ngamiland, Botswana. Botswana Notes and Records, 22, 115.Google Scholar
Pickford, M. (1991a). Late Miocene anthracothere (Mammalia, Artiodactyla) from tropical Africa. Comptes Rendus de l’Académie des Sciences de Paris, Série IIa 313, 709715.Google Scholar
Pickford, M. (1991b). Revision of the Neogene Anthracotheriidae of Africa. In Salem, M. J., ed. The Geology of Libya, Vol. 4, Amsterdam: Elsevier, pp. 14911525.Google Scholar
Pickford, M. (1994). 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). Suidae (Mammalia, Artiodactyla) from the Early Middle Miocene of Arrisdrift, Namibia: Namachoerus (gen. nov.) moruoroti and Nguruwe kijivium. Comptes Rendus de l’Académie des Sciences de Paris, Sciences de la Terre et des Planètes, 320, 319326.Google Scholar
Pickford, M. (1996a). Pliohyracids (Mammalia, Hyracoidea) from the upper Middle Miocene at Berg Aukas, Namibia. Comptes Rendus de l’Académie des Sciences de Paris, Série IIa 322, 501505.Google Scholar
Pickford, M. (1996b). Tubulidentata (Mammalia) from the middle and upper Miocene of southern Namibia. Comptes Rendus de l’ Académie des Sciences de Paris, Sciences de la Terre et des Planètes, 322, 805810.Google Scholar
Pickford, M. (1997). Lower Miocene Suiformes from the northern Sperrgebiet, Namibia, including new evidence for the systematic position of the Sanitheriidae. Comptes Rendus de l’ Académie des Sciences de Paris, Sciences de la Terre et des Planètes, 325, 285292.Google Scholar
Pickford, M. (2001a). Afrochoerodon nov. gen. kisumuensis (MacInnes) (Proboscidea, Mammalia) from Cheparawa. Middle Miocene, Kenya. Annales de Paléontologie, 87, 99117.Google Scholar
Pickford, M. (2001b). Africa’s smallest ruminant: a new tragulid from the Miocene of Kenya and the biogeography of East African Tragulidae. Geobios, 34, 437447.Google Scholar
Pickford, M. (2003a). Minute species of Orycteropus from the early Middle Miocene at Arrisdrift, Namibia. Memoirs of the Geological Survey of Namibia, 19, 195198.Google Scholar
Pickford, M. (2003b). Middle Miocene Hyracoidea from the lower Orange River valley, Namibia. Memoirs of the Geological Survey of Namibia, 19, 199205.Google Scholar
Pickford, M. (2003c). New Proboscidea from the Miocene strata in the lower Orange River Valley, Namibia. Memoirs of the Geological Survey of Namibia, 19, 207256.Google Scholar
Pickford, M. (2003d). Early and Middle Miocene Anthracotheriidae (Mammalia, Artiodactyla) from the Sperrgebiet, Namibia. Memoirs of the Geological Survey of Namibia, 19, 283289.Google Scholar
Pickford, M. (2003e). Suidae from the Middle Miocene of Arrisdrift, Namibia. Memoirs of the Geological Survey of Namibia, 19, 291303.Google Scholar
Pickford, M. (2003f). Giant dassie (Hyracoidea, Mammalia) from the middle Miocene of South Africa. South African Journal of Science, 99, 366367.Google Scholar
Pickford, M. (2004). Miocene Sanitheriidae (Suiformes, Mammalia) from Namibia and Kenya: systematic and phylogenetic implications. Annales de Paléontologie, 90, 223278.Google Scholar
Pickford, M. (2005a). Choerolophodon pygmaeus (Proboscidea: Mammalia) from the Middle Miocene of Southern Africa. South African Journal of Science, 101, 175177.Google Scholar
Pickford, M. (2005b). Orycteropus (Tubulidentata, Mammalia) from Langebaanweg and Baard’s Quarry, Early Pliocene of South Africa. Comptes Rendus Paléovol, 4, 715726.Google Scholar
Pickford, M. (2005c). Fossil hyraxes (Hyracoidea: Mammalia) from the Late Miocene and Plio-Pleistocene of Africa and the phylogeny of the Procaviidae. Palaeontologia Africana, 41, 141161.Google Scholar
Pickford, M. (2007). New mammutid proboscidean teeth from the Middle Miocene of tropical and southern Africa. Palaeontologia Africana, 42, 2935.Google Scholar
Pickford, M. (2008a). Tubulidentata from the northern Sperrgebiet, Namibia. Memoirs of the Geological Survey of Namibia, 20, 311313.Google Scholar
Pickford, M. (2008b). Hyracoidea from the Early Miocene of the northern Sperrgebiet, Namibia. Memoirs of the Geological Survey of Namibia, 20, 315325.Google Scholar
Pickford, M. (2008c). Proboscidea from the Early Miocene of the northern Sperrgebiet, Namibia. Memoirs of the Geological Survey of Namibia, 20, 327329.Google Scholar
Pickford, M. (2008d). Anthracotheriidae from the Early Miocene deposits of the northern Sperrgebiet, Namibia. Memoirs of the Geological Survey of Namibia, 20, 343347.Google Scholar
Pickford, M. (2008e). Suidae from the Early Miocene of the northern Sperrgebiet, Namibia. Memoirs of the Geological Survey of Namibia, 20, 349363.Google Scholar
Pickford, M. (2008f). Early Miocene Santheriidae from the northern Sperrgebiet, Namibia. Memoirs of the Geological Survey of Namibia, 20, 365385.Google Scholar
Pickford, M. (2012). Ancestors of Broom’s pigs. Transactions of the Royal Society of South Africa, 67, 1735.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. (2015a). Chrysochloridae (Mammalia) from the Lutetian (Middle Eocene) of Black Crow, Namibia. Communications of the Geological Survey of Namibia, 16, 105113.Google Scholar
Pickford, M. (2015b). Late Eocene Potamogalidae and Tenrecidae (Mammalia) from the Sperrgebiet, Namibia. Communications of the Geological Survey of Namibia, 16, 114152.Google Scholar
Pickford, M. (2015c). Late Eocene Chrysochloridae (Mammalia) from the Sperrgebiet, Namibia. Communications of the Geological Survey of Namibia, 16, 153193.Google Scholar
Pickford, M. (2015d). Late Eocene lorisiform primate from Eocliff, Sperrgebiet, Namibia. Communications of the Geological Survey of Namibia, 16, 194199.Google Scholar
Pickford, M. (2015e). New Titanohyracidae (Hyracoidea: Afrotheria) from the Late Eocene of Namibia. Communications of the Geological Survey of Namibia, 16, 200214.Google Scholar
Pickford, M. (2015f). Bothriogenys (Anthracotheriidae) from the Bartonian of Eoridge, Namibia. Communications of the Geological Survey of Namibia, 16, 215222.Google Scholar
Pickford, M. (2018a). New Zegdoumyidae (Rodentia, Mammalia) from the Middle Eocene of Black Crow, Namibia: taxonomy, dental formula. Communications of the Geological Survey of Namibia, 18, 4863.Google Scholar
Pickford, M. (2018b). Fossil fruit bat from the Ypresian/Lutetian of Black Crow, Namibia. Communications of the Geological Survey of Namibia, 18, 6471.Google Scholar
Pickford, M. (2018c). Additional material of Namahyrax corvus from the Ypresian/Lutetian of Black Crow, Namibia. Communications of the Geological Survey of Namibia, 18, 8186.Google Scholar
Pickford, M. (2018d). Tenrecoid mandible from Elisabethfeld (Early Miocene) Namibia. Communications of the Geological Survey of Namibia, 18, 8792.Google Scholar
Pickford, M. (2018e). Characterising the zegdoumyid rodent Tsaukhaebmys from the Ypresian/Lutetian of Black Crow, Namibia. Communications of the Geological Survey of Namibia, 19, 6670.Google Scholar
Pickford, M. (2018f). Tufamyidae, a new family of hystricognath rodents from the Palaeogene and Neogene of the Sperrgebiet, Namibia. Communications of the Geological Survey of Namibia, 19, 71109.Google Scholar
Pickford, M. & Fischer, M. S. (1987). Parapliohyrax ngororaensis, a new hyracoid from the Miocene of Kenya, with an outline of the classification of Neogene Hyracoidea. Neue Jahrbuch für Geologie und Paläontologie-Abhandlungen, 175, 207234.Google Scholar
Pickford, M. & Gommery, D. (2016). Fossil Suidae (Artiodactyla, Mammalia) from Aves Cave I and nearby sites in Bolt’s Farm Palaeokarst System, South Africa. Estudios Geológicos, 72, e059.Google Scholar
Pickford, M. & Mein, P. (1988). The discovery of fossiliferous Plio-Pleistocene cave fillings in Ngamiland, Botswana. Comptes Rendus de l’Académie des Sciences de Paris, Série IIa, 307, 16811686.Google Scholar
Pickford, M. & Mein, P. (2011). New Pedetidae (Rodentia: Mammalia) from the Mio-Pliocene of Africa. Estudios Geológicos, 67, 455469.Google Scholar
Pickford, M. & Senut, B. (1997). Cainozoic mammals from coastal Namaqualand. Palaeontologia Africana, 34, 199217.Google Scholar
Pickford, M. & Senut, B. (1998). Orange River Man, an archaic Homo sapiens from Namibia. South African Journal of Science, 94, 312.Google Scholar
Pickford, M. & Senut, B. (1999). Geology and palaeobiology of the Namib Desert, southwestern Africa. Memoirs of the Geological Survey of Namibia, 18, 1155.Google Scholar
Pickford, M. & Senut, B. (2002). The Fossil Record of Namibia. Windhoek: Geological Survey of Namibia.Google Scholar
Pickford, M. & Senut, B. (2003). Miocene paleobiology of the Orange River Valley, Namibia. Memoirs of the Geological Survey of Namibia, 19, 122.Google Scholar
Pickford, M. & Senut, B. (2008). Geology and palaeobiology of the northern Sperrgebiet: general conclusions and summary. Memoirs of the Geological Survey of Namibia, 20, 555573.Google Scholar
Pickford, M. & Senut, B. (2010). Karst geology and palaeobiology of northern Namibia. Memoirs of the Geological Survey of Namibia, 21, 174.Google Scholar
Pickford, M. & Senut, B. (2018). Afrohyrax namibensis (Hyracoidea, Mammalia) from the Early Miocene of Elisabethfeld and Fiskus, Sperrgebiet, Namibia. Communications of the Geological Survey of Namibia, 18, 93112.Google Scholar
Pickford, M. & Tassy, P. (1980). A new species of Zygolophodon (Mammalia, Proboscidea) from the Miocene hominoid localities of Meswa Bridge and Moroto (East Africa). Neues Jahrbuch für Geologie und Paläontologie-Abhandlungen, 4, 235251.Google Scholar
Pickford, M. & Uhen, M. D. (2014). Namaia Pickford et al., 2008, preoccupied by Namaia Green, 1963: proposal of a replacement name. Communications of the Geological Survey of Namibia, 15, 91.Google Scholar
Pickford, M., Senut, B. & Mein, P. (1992). Otavipithecus namibiensis, first Miocene hominoid from southern Africa. Nature, 356, 144148.Google Scholar
Pickford, M., Senut, B., Mein, P. & Conroy, G. C. (1993). Premiers gisements fossilifères post-miocènes dans le Kaokoland, nord-ouest de la Namibie. Comptes Rendus de l’Académie des Sciences de Paris, Série IIa, 317, 719720.Google Scholar
Pickford, M., Mein, P. & Senut, B. (1994). Fossiliferous Neogene karst fillings in Angola, Botswana and Namibia. South African Journal of Science, 90, 227230.Google Scholar
Pickford, M., Senut, B., Mein, P., et al. (1995). The discovery of lower and middle Miocene vertebrates at Auchas, southern Namibia. Comptes Rendus de l’Academie des Sciences de Paris, Série IIa, 322, 901906.Google Scholar
Pickford, M., Senut, B., Mein, P., et al. (1996). Preliminary results of new excavations at Arrisdrift, middle Miocene of southern Namibia. Comptes Rendus Hebdomadaires des Séances de l’Académie des Sciences de Paris, 322, 991996.Google Scholar
Pickford, M., Mein, P., Moyà-Solà, S. & Köhler, M. (1997). Phylogenetic implications of the first African middle Miocene hominoid frontal bone from Otavi, Namibia. Comptes Rendus de l’Académie des Sciences de Paris, Sciences de la Terre et des Planètes, 325, 459466.Google Scholar
Pickford, M., Eisenmann, V. & Senut, B. (1999). Timing of landscape development and calcrete genesis in northern Namaqualand, South Africa. South African Journal of Science, 95, 357359.Google Scholar
Pickford, M., Senut, B., Morales, J., Mein, P. & Sánchez, I. M. (2008). Mammalia, from the Lutetian of Namibia. Memoirs of the Geological Survey of Namibia, 20, 465514.Google Scholar
Pickford, M., Senut, B., Hipondoka, M., et al. (2014). Mio-Plio-Pleistocene geology and palaeobiology of the Etosha Pan, Namibia. Communications of the Geological Survey of Namibia, 15, 1668.Google Scholar
Piveteau, J. (1948). Un félide du Pliocène de Roussillon. Annales de Paléontologie, 43, 97124.Google Scholar
Pleurdeau, D., Imalwa, E., Détroit, F., et al. (2012). ‘Of sheep and men’: earliest direct evidence of caprine domestication in southern Africa at Leopard Cave (Erongo, Namibia). PLoS ONE, 7. DOI: 10.1371/journal.pone.0040340.Google Scholar
Plug, I. (1979a). Appendix: Striped Giraffe Shelter faunal report. In Wadley, L.. Big Elephant Shelter and its role in the prehistory of central South West Africa. Cimbebasia Series B, 3, 7172.Google Scholar
Plug, I. (1979b). The faunal remains from Wildeebeestfontein. South African Archaeological Society Goodwin Series, 3, 130132.Google Scholar
Plug, I. (1981). Some research results on the late Pleistocene and early Holocene deposits of Bushman Rock Shelter, eastern Transvaal. South African Archaeological Bulletin, 36, 1421.Google Scholar
Plug, I. (1985). Appendix 1: the faunal remains from two Iron Age sites, Rooikrans and Rhenosterkloof, central Transvaal. In Hall, S. L.. Later Iron Age sites in the Rooiberg area of the Transvaal. Annals of the Cape Provincial Museums: Human Sciences, 1, 201210.Google Scholar
Plug, I. (1989). Aspects of life in the Kruger National Park during the Early Iron Age. South African Archaeological Society Goodwin Series, 6, 6268.Google Scholar
Plug, I. (1990). The macrofaunal remains from Mhlwazini Cave, a Holocene site in the Natal Drakensberg. Natal Museum of Humanities, 2, 135142.Google Scholar
Plug, I. (1992). The macrofaunal remains of Collingham Shelter, a Late Stone Age site in Natal. Natal Museum Journal of Humanities, 4, 5359.Google Scholar
Plug, I. (1993a). The macrofaunal remains of wild animals from Abbot’s Cave and Lame Sheep Shelter, Seacow Valley, Cape. Koedoe, 36, 1526.Google Scholar
Plug, I. (1993b). The faunal remains from Nanda, an Early Iron Age site in Natal. Natal Journal of Humanities, 5, 99107.Google Scholar
Plug, I. (1993c). KwaThwaleyakhe Shelter: the faunal remains from a Holocene site in the Thukela Basin, Natal. Natal Journal of Humanities, 5, 3745.Google Scholar
Plug, I. (1993d). The macrofaunal and molluscan remains from Tloutle, a Later Stone Age site in Lesotho. Southern African Field Archaeology, 2, 4448.Google Scholar
Plug, I. (1994). Randjies: the faunal remains from a Late Iron Age site, northern Transvaal. Research by the National Cultural History Museum, 3, 119129.Google Scholar
Plug, I. (1996a). The hunter’s choice: faunal remains from Maqonqo Shelter, South Africa. Natal Journal of Humanities, 8, 4152.Google Scholar
Plug, I. (1996b). Seven centuries of Iron Age traditions at Bosutswe, Botswana: a faunal perspective. South African Journal of Science, 92, 9197.Google Scholar
Plug, I. (1997a). Early Iron Age buffalo hunters on the Kadzi River, Zimbabwe. African Archaeological Review, 14, 85105.Google Scholar
Plug, I. (1997b). Late Pleistocene and Holocene hunter-gatherers in the Eastern Highlands of South Africa and Lesotho: a faunal interpretation. Journal of Archaeological Science, 24, 715727.Google Scholar
Plug, I. (1997c). The faunal samples from Thulamela 2231AC2, Kruger National Park, South Africa. Research by the National Cultural History Museum, 6, 7893.Google Scholar
Plug, I. (1997d). Resource exploitation: animal use during the Middle Stone Age at Sibudu Cave, KwaZulu-Natal. South African Journal of Science, 100, 151158.Google Scholar
Plug, I. (1999a). Some early Iron Age communities of the Eastern escarpment and lowveld, South Africa: a faunal perspective. Archaeozoologia, 10, 189199.Google Scholar
Plug, I. (1999b). The fauna from Later Stone Age and contact sites in the Karoo, South Africa. In Becker, C., Manhart, H., Peters, J. & Schibler, J., eds. Historia animalium ex ossibus: Beiträge zur Paläoanatomie, Archäologie, Ethnologie und Geschichte der Tiermedizin. Festschrift für Angela von den Driesch zum 65 Geburstag. Rahden/Westfalen: Marie Leidorf, pp. 343353.Google Scholar
Plug, I. (2000). Overview of Iron Age fauna from the Limpopo Valley. South African Archaeological Society Goodwin Series, 8, 117126.Google Scholar
Plug, I. (2002). Faunal remains from Mzinyashana, a Later Stone Age site in KwaZulu-Natal, South Africa. Southern African Humanities, 14, 5163.Google Scholar
Plug, I. (2004). Resource exploitation: animal use during the Middle Stone Age at Sibudu Cave, KwaZulu-Natal. South African Journal of Science, 100, 151158.Google Scholar
Plug, I. & Badenhorst, S. (2001). The distribution of macromammals in southern Africa over the past 30,000 years. Transvaal Museum Monograph, 21, 1234.Google Scholar
Plug, I. & Badenhorst, S. (2002). Appendix B: bones from Muozi Midden Trench II. In Soper, R.. Nyanga: Ancient Fields, Settlements and Agricultural History in Zimbabwe. Oxford: British Institute in Eastern Africa, pp. 242248.Google Scholar
Plug, I. & Badenhorst, S. (2006). Notes on the fauna from three Late Iron Age mega-sites, Boitsemagano, Molokwane and Mabjanamatshwana, North West Province, South Africa. South African Archaeological Bulletin, 61, 5767.Google Scholar
Plug, I. & Brown, A. (1982). Mgoduyanuka: faunal remains. Annals of the Transvaal Museum, 25, 115121.Google Scholar
Plug, I. & Engela, R. (1992). The macrofaunal remains from recent excavations at Rose Cottage Cave, Orange Free State. South African Archaeological Bulletin, 47, 1625.Google Scholar
Plug, I. & Keyser, A. W. (1993). Haasgat Cave, a Pleistocene site in the central Transvaal: geomorphological, faunal and taphonomic considerations. Annals of the Transvaal Museum, 36, 139145.Google Scholar
Plug, I. & Mitchell, P. (2008). Sehonghong: hunter-gatherer utilization of animal resources in the highlands of Lesotho. Annals of the Transvaal Museum, 45, 3153.Google Scholar
Plug, I. & Peters, J. (1991). Osteomorphological differences in the appendicular skeleton of Antidorcas marsupialis (Zimmerman, 1780) and Antidorcas bondi (Cooke & Wells, 1951) (Mammalia: Bovidae) with notes on the osteometry of Antidorcas bondi. Annals of the Transvaal Museum, 35, 253264.Google Scholar
Plug, I. & Pistorius, J. C. C. (1999). Animal remains from industrial Iron Age communities in Phalaborwa, South Africa. African Archaeological Review, 16, 155184.Google Scholar
Plug, I. & Roodt, F. (1990). The faunal remains from recent excavations at uMgungundlovu. South African Archaeological Bulletin, 45, 4752.Google Scholar
Plug, I. & Sampson, C. G. (1996). European and Bushman impacts on Karoo fauna in the nineteenth century: an archaeological perspective. South African Archaeological Bulletin, 51, 2631.Google Scholar
Plug, I. & Voigt, E. A. (1985). Archaeozoological studies of Iron Age communities in southern Africa. Advances in World Archaeology, 4, 189238.Google Scholar
Plug, I., Dippenaar, N. J. & Hanisch, E. O. M. (1979). Evidence of Rattus rattus (house rat) from Pont Drift, an Iron Age site in the northern Transvaal. South African Journal of Science, 75, 82.Google Scholar
Plug, I., Bollong, C. A., Hart, T. J. G. & Sampson, C. G. (1994). Context and direct dating of pre-European livestock in the upper Seecow River valley. Annals of the South African Museum, 104, 3148.Google Scholar
Plug, I., Soper, R. & Chirawu, S. (1997). Pits, tunnels and cattle in Nyanga, Zimbabwe: new light on an old problem. South African Archaeological Bulletin, 52, 8994.Google Scholar
Plug, I., Scott, K. & Fish, W. (2000). Schoemansdal: faunal remains from selected sites in an historic village. Annals of the Transvaal Museum, 37, 125130.Google Scholar
Plug, I., Mitchell, P. & Bailey, G. (2003). Animal remains from Likoaeng, an open-air river site and its place in the post-classic Wilton of Lesotho and eastern Free State, South Africa. South African Journal of Science, 99, 143152.Google Scholar
Pocock, T. N. (1969). Appendix 1: micro-fauna provisionally identified from sieved material recovered thus far from Dump 8 at Sterkfontein. 1967–1969. South African Archaeological Bulletin, 24, 168169.Google Scholar
Pocock, T. N. (1976). Pliocene mammalian microfauna from Langebaanweg: a new fossil genus linking the Otomyinae with the Murinae. South African Journal of Science, 72, 5860.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 mammalian microfauna in southern Africa: a preliminary report including description of two new fossil muroid genera (Mammalia: Rodentia). Palaeontologia Africana, 26, 6991.Google Scholar
Pohle, H. (1928). Die Raubtiere von Oldoway. In Reck, H., ed. Wissenschaftliche Ergebnisse der Oldoway-Expedition 1913, New Series, 3, 4554.Google Scholar
Pomel, A. (1848). Etudes sur les carnassiers insectivores (Extrait): Premiere partie – Insectivores fossiles. Archives des Sciences Physiques et Naturelles, Genève, 9, 159165.Google Scholar
Pomel, A. (1879). Ossements d’Eléphants et d’Hippopotames découvertes dans une station préhistorique de la plaine d’Eglis (Provence d’Oran). Bulletin de la Société Géologique de France, Series 3, 7, 4451.Google Scholar
Pomel, A. (1892). Sur Libytherium maurisium, grand ruminant du terrain pliocène plaisancien d’Algérie. Comptes Rendus Hebdomadaires des Séances de l’Académie des Sciences, 115, 100102.Google Scholar
Pomel, A. (1893). Bubalus antiquus: Carte Géologique de l’Algérie. Alger: Fontana.Google Scholar
Pomel, A. (1895). Les éléphants Quaternaires: Carte Géologique de l’Algérie. Alger: Fontana.Google Scholar
Pomel, A. (1897). Les Equides: Carte Géologique de l’Algérie. Alger: Fontana.Google Scholar
Power, J. H. (1955). Power’s Site, Vaal River. South African Archaeological Bulletin, 10, 96101.Google Scholar
Prat, S. & Gommery, D. (2012). First partial skeleton of Paranthropus robustus from Swartkrans (South Africa). In 2nd Annual Meeting of the European Society for the Study of Human Evolution, Bordeaux, p. 142.Google Scholar
Prat, S., Jashajsvili, T., Gommery, D. & Thackeray, J. F. (2014). A specimen of Paranthropus robustus from Bolt’s Farm, Cradle of Humankind, South Africa? In 4th Annual Meeting European Society for the Study of Human Evolution, Florence, p. 134.Google Scholar
Pwiti, G. (1996). Continuity and change. Studies in African Archaeology, 13, 1180.Google Scholar
Pycraft, W. P. (1925). On the calvaria found at Boskop, Transvaal, in 1913 and its relationship to Cromagnard and Negroid skulls. Journal of the Royal Anthropological Institute of Great Britain and Ireland, 55, 179198.Google Scholar
Quérouil, S., Hutterer, R., Barrière, P., et al. (2001). Phylogeny and evolution of African shrews (Mammalia: Soricidae) inferred from 16s rRNA sequence. Molecular Phylogenetics and Evolution, 20, 185195.Google Scholar
Quiralte, V., Sánchez, I. M., Morales, J. & Pickford, M. (2008). Tragulidae (Artiodactyla, Ruminantia) from the Early Miocene of the Sperrgebiet, Southern Namibia. Memoirs of the Geological Survey of Namibia, 20, 387396.Google Scholar
Qumsiyeh, M. B. (1986). Phylogenetic studies of the rodent family Gerbillidae: 1. Chromosomal evolution in the southern African complex. Journal of Mammalogy, 67, 680692.Google Scholar
Qumsiyeh, M. B., Hamilton, M. J., Dempster, E. R. & Baker, R. J. (1991). Cytogenetics and systematics of the rodent genus Gerbillurus. Journal of Mammalogy, 72, 8996.Google Scholar
Rafinesque, C. S. (1814). Précis des découvertes et travaux somiologiques. Palerme: Royale Typographie Militaire.Google Scholar
Rafinesque, C. S. (1820). I Class. Mastosia – the Sucklers. Annals of Nature, 1, 24.Google Scholar
Ramsay, P. J., Smith, A. M., Lee-Thorp, J. A., et al. (1993). 130 000-year-old fossil elephant found near Durban, South Africa: preliminary report. South African Journal of Science, 89, 165.Google Scholar
Rasmussen, D. T. & Gutiérrez, M. (2010). Hyracoidea. In Werdelin, L. & Sanders, W. J., eds. Cenozoic Mammals of Africa. Berkeley, CA: University of California Press, pp. 123145.Google Scholar
Rasmussen, D. T., Pickford, M., Mein, P., Senut, B. & Conroy, G. C. (1996). Earliest known procaviid hyracoid from the late Miocene of Namibia. Journal of Mammalogy, 77, 745754.Google Scholar
Ratcliffe, J. M. (2002). Myotis welwitschii. Mammalian Species, 701, 13.Google Scholar
Ray, J. C. (1995). Civettictis civetta. Mammalian Species, 488, 17.Google Scholar
Reck, H. (1928). Pelorovis oldowayensis nov. gen. nov. sp. Wissenschaft Ergebnisse der Oldoway-Expedition 1913, New Series, 3, 5667.Google Scholar
Rector, A. L. & Reed, K. E. (2010). Middle and Late Pleistocene faunas of Pinnacle Point and their paleoecological implications. Journal of Human Evolution, 59, 340357.Google Scholar
Reed, K. E., Kitching, J. W., Grine, F. E., Jungers, W. L. & Sokoloff, L. (1993). Proximal femur of Australopithecus africanus from member 4, Makapansgat, South Africa. American Journal of Physical Anthropology, 92, 115.Google Scholar
Reed Cohen, D. (2010). Hunting and herding at Moritsane, a village in southeastern Botswana, c. AD 1165–1275. South African Archaeological Bulletin, 65, 154163.Google Scholar
Renard, A., Lavoie, M., Pitt, J. A. & Larivière, S. (2015). Felis nigripes. Mammalian Species, 47, 7883.Google Scholar
Repenning, C. A. (1965). An extinct shrew from the early Pleistocene of South Africa. Journal of Mammalogy, 46, 189196.Google Scholar
Reynard, J. P., Discamps, E., Badenhorst, S., Van Niekerk, K. & Henshilwood, C. S. (2016a). Subsistence strategies in the southern Cape during the Howiesons Poort: taphonomic and zooarchaeological analyses of Klipdrift Shelter, South Africa. Quaternary International, 404, 219.Google Scholar
Reynard, J. P., Discamps, E., Wurz, S., et al. (2016b). Occupational intensity and environmental changes during the Howiesons Poort at Klipdrift Shelter, southern Cape, South Africa. Palaeogeography, Palaeoclimatology, Palaeoecology, 449, 349364.Google Scholar
Reynolds, S. C. (2012). Nyctereutes terblanchei: the raccoon dog that never was. South African Journal of Science, 108. DOI: 10.4102/sajs.v108i1/2.589.Google Scholar
Reynolds, S. C. (2010a). Morphological evaluation of genetic evidence for a Pleistocene extirpation of eastern African impala. South African Journal of Science, 106. DOI: 10.4102/sajs.v106i11/12.325.Google Scholar
Reynolds, S. C. (2010b). Where the wild things were: spatial and temporal distribution of carnivores in the Cradle of Humankind (Gauteng, South Africa) in relation to the accumulation of mammalian and hominin assemblages Journal of Taphonomy, 8, 233257.Google Scholar
Reynolds, S. C. & Bishop, L. C. (2006). Craniodental variability in modern and fossil plains zebra (Equus burchellii Gray 1824) from East and southern Africa. In Mashkour, M., ed. Equids in Time and Space. Oxford: Oxbow Books, pp. 4960.Google Scholar
Reynolds, S. C. & Kibii, J. M. (2011). Sterkfontein at 75: review of palaeoenvironments, fauna and archaeology from the hominin site of Sterkfontein (Gauteng Province, South Africa). Palaeontologia Africana, 46, 5988.Google Scholar
Reynolds, S. C., Vogel, J. C., Clarke, R. J. & Kuman, K. (2003). Preliminary results of excavations at Lincoln Cave, Sterkfontein, South Africa. South African Journal of Science, 99, 286288.Google Scholar
Reynolds, S. C., Clarke, R. J. & Kuman, K. (2007). The view from the Lincoln Cave: mid- to late Pleistocene fossil deposits from Sterkfontein hominid site, South Africa. Journal of Human Evolution, 53, 160271.Google Scholar
Rhoads, S. N. (1896). Mammals collected by Dr A. Donaldson Smith during his expedition to Lake Rudolf, Africa. Proceedings of the Academy of Natural Sciences of Philadelphia, 48, 517546.Google Scholar
Riccucci, M. & Lanza, B. (2008). Neoromicia Roberts, 1926 (Mammalia, Vespertilionidae): correction of gender and etymology. Italian Journal of Zoology, 19, 175177.Google Scholar
Rightmire, G. P. (1978). Florisbad and human population succession in southern Africa. American Journal of Physical Anthropology, 48, 475486.Google Scholar
Rightmire, G. P. (1979a). Appendix 3: human skeletal remains from Die Kelders, Cape. In Schweitzer, F. R.. Excavations at Die Kelders, Cape Province, South Africa: the Holocene deposits. Annals of the South African Museum, 78, 233.Google Scholar
Rightmire, G. P. (1979b). Implications of Border Cave skeletal remains for Later Pleistocene human evolution. Current Anthropology, 20, 2335.Google Scholar
Rightmire, G. P. & Deacon, H. J. (1991). Comparative studies of Late Pleistocene human remains from Klasies River Mouth, South Africa. Journal of Human Evolution, 20, 131156.Google Scholar
Rightmire, G. P., Deacon, H. J., Schwartz, J. H. & Tattersall, I. (2006). Human foot bones from Klasies River main site, South Africa. Journal of Human Evolution, 50, 96103.Google Scholar
Robbins, C. B. (1978). Taxonomic identification of Scotophilus nigrita (Schreber) (Chiroptera: Vespertilionidae). Journal of Mammalogy, 59, 212213.Google Scholar
Robbins, C. B., De Vree, F. & Cakenberghe, V. (1985). A review of the systematics of the African bat genus Scotophilus (Vespertilionidae). Annales du Musée Royal de l’Afrique Centrale Série in-8vo Sciences Zoologiques, 237, 5384.Google Scholar
Robbins, L. H. (1986). Recent archaeological research in southeastern Botswana: the Thamaga site. Botswana Notes and Records, 18, 113.Google Scholar
Robbins, L. H. (1990). Excavation at the White Paintings Rock-Shelter Tsodilo Hills. Nyame Akuma, 34, 24.Google Scholar
Robbins, L. H., Murphy, M. L., Stevens, N. J., et al. (1996). Paleoenvironment and archaeology of Drotsky’s Cave: western Kalahari Desert, Botswana. Journal of Archaeological Science, 23, 722.Google Scholar
Robbins, L. H., Murphy, M. L., Brook, G. A., et al. (2000). Archaeology, palaeoenvironment and chronology of the Tsodilo Hills White Paintings Rock Shelter, northwest Kalahari Desert, Botswana. Journal of Archaeological Science, 27, 10851113.Google Scholar
Robbins, L. H., Campbell, A. C., Murphy, M. L., et al. (2008). Recent archaeological research at Toteng, Botswana: early domesticated livestock in the Kalahari. Journal of African Archaeology, 6, 131149.Google Scholar
Robbins, L. H., Campbell, A. C., Murphy, M. L., et al. (2009). Mogapelwa: archaeology, palaeoenvironment and oral traditions at Lake Ngami, Botswana. South African Archaeological Bulletin, 64, 1332.Google Scholar
Roberts, A. (1913). Supplement to list of mammals in the Transvaal Museum. Annals of the Transvaal Museum, 4, 108109.Google Scholar
Roberts, A. (1914). Supplementary list of African mammals in the collection of the Transvaal Museum, with descriptions of some new species. Annals of the Transvaal Museum, 5, 180186.Google Scholar
Roberts, A. (1919). Descriptions of some new mammals. Annals of the Transvaal Museum, 6, 112115.Google Scholar
Roberts, A. (1924). Some additions to the list of South African mammals. Annals of the Transvaal Museum, 10, 5976.Google Scholar
Roberts, A. (1926). Some new S. African mammals and some changes in nomenclature. Annals of the Transvaal Museum, 11, 245263.Google Scholar
Roberts, A. (1929). New forms of African mammals. Annals of the Transvaal Museum, 13, 82121.Google Scholar
Roberts, A. (1931). New forms of South African mammals. Annals of the Transvaal Museum, 14, 221236.Google Scholar
Roberts, A. (1932). Preliminary description of fifty-seven new forms of South African mammals. Annals of the Transvaal Museum, 15, 119.Google Scholar
Roberts, A. (1933). Eleven new forms of South African mammals. Annals of the Transvaal Museum, 15, 265270.Google Scholar
Roberts, A. (1938). Description of new forms of mammals. Annals of the Transvaal Museum, 19, 231245.Google Scholar
Roberts, A. (1946). Descriptions of numerous new subspecies of mammals. Annals of the Transvaal Museum, 20, 303328.Google Scholar
Roberts, A. (1951). The Mammals of South Africa. Johannesburg: The Trustees of the Mammals of South Africa Book Fund.Google Scholar
Robertshaw, P. T. (1977). Excavations at Paternoster, South-Western Cape. South African Archaeological Bulletin, 32, 6373.Google Scholar
Robertshaw, P. T. (1978). Archaeological investigations at Langebaan Lagoon, Cape Province. Palaeoecology of Africa, 10–11, 139148.Google Scholar
Robertshaw, P. T. (1979). Excavations at Duiker Eiland, Vredenburg District, Cape Province. Annals of the Cape Provincial Museums (Human Sciences), 1, 126.Google Scholar
Robey, T. (1980). Mpambanyoni: a Late Iron Age site on the Natal south coast. Annals of the Natal Museum, 24, 147164.Google Scholar
Robinson, J. T. (1970). Two new early hominid vertebrae from Swartkrans. Nature, 225, 12171219.Google Scholar
Robinson, T. J. & Dippenaar, N. J. (1983). The status of Lepus saxatilis, L. whytei and L. crawshayi in southern Africa. Acta Zoologica Fennica, 174, 3539.Google Scholar
Robinson, T. J. & Dippenaar, N. J. (1987). Morphometrics of the South African Leporidae. II: Lepus Linnaeus, 1758 and Bunolagus Thomas, 1929. Annals of the Transvaal Museum, 34, 379404.Google Scholar
Robinson, T. J. & Matthee, C. A. (2005). Phylogeny and evolutionary origins of the Leporidae: a review of cytogenetics, molecular analyses and a supermatrix analysis. Mammal Review, 35, 231247.Google Scholar
Robinson, T. J. & Skinner, J. D. (1983). Karyology of the riverine rabbit, Bunolagus monticularis and its taxonomic implications. Journal of Mammalogy, 64, 678681.Google Scholar
Roger, O. (1902). Wirbeltierreste aus dem Obermiocän der bayerischschwäbischen Hochebene. Bericht des Naturwissenschaftlichen Vereines für Schwaben und Neuburg, 1902, 122.Google Scholar
Rookmaker, L. C. (1989). The Zoological Exploration of Southern Africa 1650–1790. Rotterdam: Balkema.Google Scholar
Rookmaker, L. C. & Meester, J. (1988). Case 2605: Euryotis brantsii A. Smith, 1834 (currently Parotomys brantsii; Mammalia, Rodentia) – proposed conservation of the specific name. Bulletin of Zoological Nomenclature, 45, 4344.Google Scholar
Rossouw, L. (2006). Florisian mammal fossils from erosional gullies along the Modder River at Mitasrust Farm, central Free State, South Africa. Navorsinge van die Nasionale Museum Bloemfontein, 22, 145162.Google Scholar
Rovinsky, D. S., Herries, A. I. R., Menter, C. G. & 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, 18.2.34A, 121.Google Scholar
Rüppell, E. (1835). Neue Wirbelthiere zu der Fauna von Abyssinien gehörig. Frankfurt am Main: Schmerber.Google Scholar
Rüppell, E. (1842a). Über Säugthiere aus der Ordnung der Nager. Abhandlungen aus dem Gebiete der beschreibenden Naturgeschichte Museum Senckenbergianum, 1833–1845, 91101.Google Scholar
Rüppell, E. (1842b). Beschreibung mehrerer neuer Säugethiere, in der zoologischen Sammlung der Senckenbergische naturforschenden Gesellschaft befindlich. Museum Senckenbergianum: Abhandlungen aus dem Gebiete der beschreibenden Naturgeschichte, 3, 129144.Google Scholar
Russo, I. M. (2009). Patterns and processes underlying genetic diversity in the Namaqua rock mouse Micaelamys namquensis Smith, 1834 (Rodentia: Muridae) from southern Africa. Unpublished PhD thesis, University of Pretoria.Google Scholar
Sach, V. J. & Heizmann, E. P. J. (2001). Stratigraphy and mammal faunas of the Brackwassermolasse in the surroundings of Ulm (Southwest Germany). Stuttgarter Beiträge zur Naturkunde Serie B (Geologie und Paläontologie), 310, 195.Google Scholar
Sadr, K. (2007). Early first millennium pastoralists on Kasteelberg? The UB/UCT excavation at KBA. South African Archaeological Bulletin, 62, 154161.Google Scholar
Sadr, K. & Plug, I. (2001). Faunal remains in the transition from hunting to herding in Southeastern Botswana. South African Archaeological Bulletin, 56, 7682.Google Scholar
Sallam, H. M. & Seiffert, E. R. (2016). New phiomorph rodents from the latest Eocene of Egypt and the impact of Bayesian ‘clock’-based phylogenetic methods on estimates of basal hystricognath relationships and biochronology. PeerJ, 4, 153.Google Scholar
Sandelowsky, B. H. (1979). Kapako and Vungu Vungu: Iron Age sites on the Kavango River. South African Archaeological Society Goodwin Series, 3, 5261.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., Gheerbrandt, E., Harris, J. M., Saegusa, H. & Delmer, C. (2010a). Proboscidea. In Werdelin, L. & Sanders, W. J., eds. Cenozoic Mammals of Africa. Berkeley, CA: University of California Press, pp. 161251.Google Scholar
Sanders, W. J., Rasmussen, D. T. & Kappelman, J. (2010b). Embripothoda. In Werdelin, L. & Sanders, W. J., eds. Cenozoic Mammals of Africa. Berkeley, CA: University of California Press, pp. 115122.Google Scholar
Sardella, R. & Werdelin, L. (2007). Amphimachairodus (Felidae, Mammalia) from Sahabi (Latest Miocene–Earliest Pliocene, Libya), with a review of African Miocene Machairodontinae. Rivista Italiana di Paleontologia e Stratigrafia, 113, 6777.Google Scholar
Savage, R. J. G. (1965). The Miocene Carnivora of East Africa. Fossil Mammals of Africa, 19, 239316.Google Scholar
Schepers, G. W. H. (1941). The mandible of the Transvaal human skeleton from Springbok Flats. Annals of the Transvaal Museum, 20, 253271.Google Scholar
Schinz, H. R. (1821–1825). Das Thierreich eingetheilt nach dem Bau der Thiere als Grundlage ihrer Naturgeschichte und der vergleichenden Anatomie von dem Herrn Ritter von Cuvier. Stuttgart: Cotta’schen.Google Scholar
Schoonraad, M. & Beaumont, P. B. (1968). The North Brabant Shelter, north western Transvaal. South African Journal of Science, 64, 319331.Google Scholar
Schreber, J. C. D. (1775–1792). Die Säugethiere in Abbildungen nach der Natur, mit Beschreibungen. Erlangen: Walther.Google Scholar
Schwann, H. (1906). A list of the mammals obtained by Messrs R.B. Woosnam and R.E. Dent in Bechuanaland. Proceedings of the Zoological Society of London, 1906, 101111.Google Scholar
Schwartz, G. T. (1997). Re-evaluation of the Plio-Pleistocene hyraxes (Mammalia, Procaviidae) from South Africa. Neues Jahrbuch für Geologie und Paläontologie, Abhandlungen, 206, 365383.Google Scholar
Schwarz, E. (1932). Neue diluviale Antilopen aus Ostafrika. Zentralblatt für Mineralogie, Geologie und Paläontologie B, 1932, 14.Google Scholar
Schweitzer, F. R. (1974). Archaeological evidence for sheep at the Cape. South African Archaeological Bulletin, 29, 7582.Google Scholar
Schweitzer, F. R. (1979). Excavations at Die Kelders, Cape Province, South Africa: the Holocene deposits. Annals of the South African Museum, 78, 101233.Google Scholar
Schweitzer, F. R. & Scott, K. (1973). Early occurrence of domestic sheep in sub-Saharan Africa. Nature, 241, 547.Google Scholar
Sclater, P. L. & Thomas, O. (1894). The Book of Antelopes. London: Porter.Google Scholar
Scott, K. & Klein, R. G. (1981). A hyena-accumulated bone assemblage from Late Holocene deposits at Deelpan, Orange Free State. Annals of the South African Museum, 86, 217227.Google Scholar
Scott, L. & Brink, J. S. (1992). Quaternary palaeoenvironments of pans in central South Africa: palynological and palaeontological evidence. South African Geographer, 19, 2234.Google Scholar
Scott, W. B. (1907). A collection of fossil mammal bones from the coast of Zululand. Geological Survey of Natal and Zululand Third and Final Report.Google Scholar
Sealy, J., Maggs, T., Jerardino, A. & Kaplan, J. (2004). Excavations at Melkbosstrand: variability among herder sites on Table Bay, South Africa. South African Archaeological Bulletin, 59, 1728.Google Scholar
Seeley, H. G. (1891). On Bubalis bainii (Seeley). Geological Magazine, New series 3, 8, 199202.Google Scholar
Sénégas, F. (1996). Introduction à l’étude des faunes de rongeurs du Plio-Pleistocène du sud de l’Afrique: analyse d’un échantillon de brêches fossilifères de gisements du Transvaal (S. Afr.) et du site de Friesenberg (Namibie). Montpellier: Diplôme d’Etudes Aprofondies (DEA) Paléontologie, Montpellier II.Google Scholar
Sénégas, F. (2001). Interpretation of the dental pattern of the South African fossil Euryotomys (Rodentia, Murinae) and the origin of otomyine dental morphology. In Denys, C., Granjon, L. & Poulet, A., eds. African Small Mammals. Paris: I.R.D., pp. 151160.Google Scholar
Sénégas, F. (2004). A new species of Petromus (Rodentia, Hystricognatha, Petromuridae) from the Early Pliocene of South Africa and its paleoenvironmental implications. Journal of Vertebrate Paleontology, 24, 757763.Google Scholar
Sénégas, F. & Avery, D. M. (1998). New evidence for the murine origin of the Otomyinae (Mammalia, Rodentia) and the age of Bolt’s Farm. South African Journal of Science, 94, 503507.Google Scholar
Sénégas, F. & Michaux, J. (2000). Boltimys broomi gen nov.; sp. nov. (Rodentia, Mammalia), nouveau Muridae d’affinité incertaine du Pliocène inférieur d’Afrique du Sud. Comptes Rendus de l’ Académie des Sciences de Paris, Sciences de la Terre et des Planètes, 330, 521525.Google Scholar
Sénégas, F., Paradis, E. & Michaux, J. (2005). Homogeneity of fossil assemblages extracted from mine dumps: an analysis of Plio-Pleistocene fauna from South African caves. Lethaia, 38, 315322.Google Scholar
Senut, B. (1996). Plio-Pleistocene Cercopithecoidea from the Koanaka Hills (Ngamiland, Botswana). Comptes Rendus de l’Academie des Sciences de Paris, Série IIa, 322, 423428.Google Scholar
Senut, B. (2003). The Macroscelididae from the Miocene of the Orange River, Namibia. Memoirs of the Geological Survey of Namibia, 19, 119141.Google Scholar
Senut, B. (2008). Macroscelididae from the lower Miocene of the northern Sperrgebiet, Namibia. Memoirs of the Geological Survey of Namibia, 20, 185225.Google Scholar
Senut, B. & Georgalis, G. (2014). Brevirhynchocyon gen. nov., a new name for the genus Brachyrhynchocyon Senut, 2008 (Mammalia, Macroscelidea) preoccupied by Brachyrhynchocyon Loomis, 1936 (Mammalia, Carnivora). Communications of the Geological Survey of Namibia, 15, 69.Google Scholar
Senut, B., Pickford, M., Mein, P., Conroy, G. & Van Couvering, J. (1992). Discovery of 12 new Late Cainozoic fossiliferous sites in paleokarst of the Otavi Mountains, Namibia. Comptes Rendus de l’Académie des Sciences de Paris, Série IIa, 314, 727733.Google Scholar
Senut, B., Pickford, M., De Wit, M., et al. (1996). Biochronology of sediments at Bosluis Pan, Northern Cape Province, South Africa. South African Journal of Science, 92, 249251.Google Scholar
Senut, B., Pickford, M. & Wessels, D. (1997). Panafrican distribution of Lower Miocene Hominoidea. Comptes Rendus de l’ Académie des Sciences de Paris, Sciences de la Terre et des Planètes, 325, 741746.Google Scholar
Senut, B., Pickford, M., Braga, M., Marais, D. & Coppens, Y. (2000). Découverte d’un Homo sapiens archaïque à Oranjemund, Namibie. Comptes Rendus de l’ Académie des Sciences de Paris, Sciences de la Terre et des Planètes, 330, 813819.Google Scholar
Shackley, M. (1980). An Acheulean industry with Elephas recki fauna from Namib IV, South West Africa (Namibia). Nature, 284, 340341.Google Scholar
Shackley, M. (1985). Palaeolithic archaeology of the central Namib Desert. Cimbebasia, 6, 184.Google Scholar
Shaw, G. (1800). General Zoology, or Systematic Natural History. London: Kearsley.Google Scholar
Shaw, J. C. M. (1937). Evidence concerning a large fossil hyrax. Journal of Dental Research, 16, 3740.Google Scholar
Shaw, J. C. M. (1938). The teeth of the South African fossil pig (Notochoerus capensis syn. meadowsi) and their geological significance. Transactions of the Royal Society of South Africa, 26, 2537.Google Scholar
Shaw, J. C. M. & Cooke, H. B. S. (1940). New fossil pig remains from the Vaal River gravels. Transactions of the Royal Society of South Africa, 28, 293299.Google Scholar
Shee, J. C. (1963). Appendix A: skeletal remains from Gokomere. In Robinson, K. R.. Further excavations in the Iron Age deposits at the Tunnel Site, Gokomere Hill, Southern Rhodesia. South African Archaeological Bulletin, 18, 170.Google Scholar
Shenjere-Nyabezi, P. (2017). Ndongo: a Zimbabwe culture site in the Middle Save. In Manyanga, M. & Chirikure, S., eds. Archives, Objects, Places and Landscapes. Cameroon: Bamenda, pp. 137167.Google Scholar
Shenjere-Nyabezi, P. & Pwiti, G. (2015). Late Stone Age economies of the Lower Save Valley, south eastern Zimbabwe: an archaeozoological perspective. Zimbabwea, 11, 2635.Google Scholar
Shenjere-Nyabezi, P., Pwiti, G. & Manyanga, M. (2013). Making the most out of rubbish: trends in archaeozoological studies in post-independence Zimbabwe. In Manyanga, M. & Katsamudanga, S., eds. Archaeology in the Post-Independence Era. Harare: Sapes Books, pp. 117142.Google Scholar
Shortridge, G. C. (1942). Field notes on the first and second expeditions of the Cape museums’ mammal survey of the Cape Province: descriptions of some new subgenera and subspecies. Annals of the South African Museum, 36, 27100.Google Scholar
Shortridge, G. C. & Carter, D. (1938). A new genus and new species and subspecies of mammals from Little Namaqualand and the north-west Cape Province; and a new subspecies of Gerbillus paeba from the eastern Cape Province. Annals of the South African Museum, 32, 281291.Google Scholar
Shoshani, J., Goldman, C. A. & Thewissen, J. G. M. (1988). Orycteropus afer. Mammalian Species, 300, 18.Google Scholar
Shrubsall, F. C. (1911). A note on craniology. In Péringuey, L.. The Stone Ages of South Africa as represented in the collections of the South African Museum. Annals of the South African Museum, 8, 202206.Google Scholar
Sigvallius, B. (1988). The faunal remains from Manyikeni. Studies in African Archaeology, 2, 2334.Google Scholar
Sinclair, P. J. J. (1991). Excavations at Chivowa Hill, south central Zimbabwe. Zimbabwea, 3, 2350.Google Scholar
Singer, R. (1954). The Saldanha skull from Hopefield, South Africa. American Journal of Physical Anthropology, 12, 345362.Google Scholar
Singer, R. (1956). Man and animals in South Africa. Journal of the Palaeontological Society of India, 1, 122130.Google Scholar
Singer, R. & Boné, E. L. (1960). Modern giraffes and the fossil giraffids of Africa. Annals of the South African Museum, 45, 375548.Google Scholar
Singer, R. & Fuller, A. O. (1962). The geology and description of a fossiliferous deposit near Zwartklip in False Bay. Transactions of the Royal Society of South Africa, 36, 205211.Google Scholar
Singer, R. & Keen, E. N. (1955). Fossil Suiformes from Hopefield. Annals of the South African Museum, 42, 169179.Google Scholar
Singleton, M. (2000). The phylogenetic affinities of Otavipithecus namibiensis. Journal of Human Evolution, 38, 537573.Google Scholar
Singleton, M., Gilbert, C. C., Frost, S. R. & Seitelman, B. C. (2016). Comparative morphometric analysis of a juvenile papionin (Primates: Cercopithecidae) from Kromdraai A. Annals of the Ditsong National Museum of Natural History, 6, 117.Google Scholar
Skead, C. J. (2011). Historical Incidence of the Larger Land Mammals in the Broader Western and Northern Cape, 2nd edition. Port Elizabeth: Centre for African Conservation Ecology.Google Scholar
Skinner, J. D. & Chimimba, C. T. (2005). The Mammals of the Southern African Subregion, 3rd edition. Cambridge: Cambridge University Press.Google Scholar
Skurski, D. A. & Waterman, J. M. (2005). Xerus inauris. Mammalian Species, 781, 14.Google Scholar
Smit, A. A. & Van der Bank, H. F. H. (2001). Isozyme and allozyme markers distinguishing two morphologically similar, medically important Mastomys species (Rodentia: Muridae). BMC Genetics, 2. DOI: 10.1186/1471-2156-2-15.Google Scholar
Smith, A. (1827). Description of two quadrupeds inhabiting the South of Africa, about the Cape of Good Hope. Transactions of the Linnean Society of London, 15, 460470.Google Scholar
Smith, A. (1829). Contributions to the natural history of South Africa. Zoological Journal, 4, 433444.Google Scholar
Smith, A. (1830). [New species of Macroscelides and Erinaceus]. Proceedings of the Committee of Science and Correspondence of the Zoological Society of London, 1, 11.Google Scholar
Smith, A. (1831). Contributions to the natural history of South Africa, etc. No. 1. South African Quarterly Journal, 1(5), 1024.Google Scholar
Smith, A. (1833a). African Zoology, continued. South African Quarterly Journal, Series 2, 1, 4964.Google Scholar
Smith, A. (1833b). African Zoology, continued. South African Quarterly Journal, Series 2, 1, 8196.Google Scholar
Smith, A. (1834a). African Zoology, continued. South African Quarterly Journal, Series 2, 2, 145160.Google Scholar
Smith, A. (1834b). African Zoology, continued. South African Quarterly Journal, Series 2, 2, 169192.Google Scholar
Smith, A. (1834c). African Zoology, continued. South African Quarterly Journal, Series 2, 3, 209224.Google Scholar
Smith, A. (1836 [1834d]). Report of the Expedition for Exploring Central Africa from the Cape of Good Hope, June 23rd, 1834. Cape Town: Government Gazette.Google Scholar
Smith, A. (1838–1849). Illustrations of the Zoology of South Africa. London: Smith, Elder.Google Scholar
Smith, A. B. (1981). An archaeological investigation of Holocene deposits at Rooiels Cave, southwestern Cape. South African Archaeological Bulletin, 36, 7583.Google Scholar
Smith, A. B. (1995). Archaeological observations along the Orange River and its hinterland. In Smith, A. B., ed. Einiqualand: Studies of the Orange River Frontier. Cape Town: University of Cape Town Press, pp. 265300.Google Scholar
Smith, A. B. & Jacobson, L. (1995). Excavations at Geduld and the appearance of early domestic stock in Namibia. South African Archaeological Bulletin, 50, 320.Google Scholar
Smith, A. & Mütti, B. (2013). The past 3000 years of human habitation and coastal resource exploitation on the Vredenburg Peninsula. In Jerardino, A., Malan, A. & Braun, D., eds. The Archaeology of the West Coast of South Africa. Oxford: Archaeopress, pp. 6884.Google Scholar
Smith, A. B. & Ripp, M. R. (1978). An archaeological reconnaissance of the Doorn/Tanqua Karoo. South African Archaeological Bulletin, 33, 118133.Google Scholar
Smith, A. B., Sadr, K., Gribble, J. & Yates, R. (1991). Excavations in the south-western Cape, South Africa and the archaeological identity of prehistoric hunter-gatherers within the last 2000 years. South African Archaeological Bulletin, 46, 7191.Google Scholar
Smith, A. B., Yates, R., Miller, D., Jacobson, L. & Evans, G. (1995). Excavations at Geduld and the appearance of early domestic stock in Namibia. South African Archaeological Bulletin, 50, 320.Google Scholar
Smith, C. H. (1827). Synopsis of the species of the Class Mammalia. In Griffiths, E., ed. The Animal Kingdom Arranged in Conformity with its Organization by the Baron Cuvier with Additional Descriptions, Vol. 5. London: Whittaker.Google Scholar
Smith, H. F. & Grine, F. E. (2008). Cladistic analysis of early Homo crania from Swartkrans and Sterkfontein, South Africa. Journal of Human Evolution, 54, 684704.Google Scholar
Smith, P., Nshimirimana, R., De Beer, F., et al. (2012). Canteen Kopje: a new look at an old skull. South African Journal of Science, 108. DOI: 10.4102/sajs.v108i1/2.738.Google Scholar
Smuts, J. (1832). Dissertation Zoologica, Ennumerationem Mammalium Capensium. London: Cyfveer.Google Scholar
Sparrman, A. (1779). Bos caffer, et nytt Species af Buffel, från Caput Bonæ Spei. Kungl. Svenska Vetenskapsakademiens Handlingar, Series 1, 40, 7984.Google Scholar
Sparrman, A. (1783). Resa Goda-Hopps-Udden, Soedra Pol-kretsen Och Omkring Jordklotet, Samt till Hottentott-och Caffer-Landen, Aren 1772–76. Stockholm: Nordstroem.Google Scholar
Sparrman, A. (1784). Mus pumilio, en ny Râtta från det fódra af Africa, uptåkt och belrifven. Kungl. Svenska Vetenskapsakademiens Handlingar, Series 2, 5, 236237.Google Scholar
Speke, J. H. (1863). Journal of the Discovery of the Source of the Nile. Edinburgh: Blackwood & Sons.Google Scholar
Steele, T. E. & 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
Steele, T. E., Mackay, A., Orton, J. & Schwortz, S. (2012). Varsche Rivier 003, a new Middle Stone Age site in southern Namaqualand, South Africa. South African Archaeological Bulletin, 67, 108119.Google Scholar
Steininger, C., Berger, L. R. & Kuhn, B. F. (2008). A partial skull of Paranthropus robustus from Cooper’s Cave, South Africa. South African Journal of Science, 104, 143146.Google Scholar
Steyn, M. & Broekhuizen, T. (1993). Report on human skeletal remains from Rooiberg (Transvaal). Southern African Field Archaeology, 2, 5355.Google Scholar
Steyn, M. & Nienaber, W. C. (2000). Iron Age human skeletal remains from the Limpopo Valley and Soutpansberg area. South African Archaeological Society Goodwin Series, 8, 112116.Google Scholar
Steyn, M., Meyer, A. & Loots, M. (1998). Report on isolated human remains from K2, South Africa. Southern African Field Archaeology, 7, 5358.Google Scholar
Steyn, M., Nienaber, W. C., Loots, M. & Meiring, J. H. (1999). An infant grave from K2 (Greefswald). South African Archaeological Bulletin, 54, 104106.Google Scholar
Steyn, M., Binneman, J. & Loots, M. (2007). The Kouga mummified human remains. South African Archaeological Bulletin, 62, 38.Google Scholar
Storr, G. C. C. (1780). Prodromus Methodi Mammalium. Tübingen: Reissianis.Google Scholar
Stromer, E. (1921). Erste Mitteilung űber Tertiare Wirbeltier-Reste aus Deutsch-Sudwestafrika. Sitzungsberichte der Bayern Akademie der Wissenschaften zu München, 1921 (II), 331340.Google Scholar
Stromer, E. (1923). Ergebnisse der Bearbeitung mitteltertiarer Wirbeltier-Reste aus Deutsch-Sudwestafrika. Sitzungsberichte der Bayern Akademie der Wissenschaften München, 1923 (II), 253270.Google Scholar
Stromer, E. (1926). Reste land- und süsswasser-bewohnender Wirbeltiere aus den Diamenten-feldern Deutsch-Südwestafrikas. In Kaiser, E., ed. Die Diamentenwuste südwestafrikas. Vol. 2. Berlin: Dietrich Reimer (Ernst Volsen), pp. 107153.Google Scholar
Stromer, E. (1931a). Reste Susswasser- und Land-bewohnender Wirbeltiere aus den Diamantfeldern Klein-Namaqualandes (Sudwestafrika). Sitzungsberichte der Bayern Akademie der Wissenschaften München. Mathematisch-Naturwissenschaftliche Abteilung, 1931, 1747.Google Scholar
Stromer, E. (1931b). Palaeothentoides africanus nov. gen., nov. spec., ein erstes Beuteltier aus Afrika. Sitzungsberichte der Bayern Akademie der Wissenschaften München. Mathematisch-Naturwissenschaftliche Abteilung, 1931, 177190.Google Scholar
Stynder, D. D. (1997). The use of faunal evidence to reconstruct site history at Hoedjiespunt 1 (HDP1), Western Cape. Unpublished MA thesis, University of Cape Town.Google Scholar
Stynder, D., Moggi-Cecchi, J., Berger, L. R. & Parkington, J. E. (2001). Human mandibular incisors from the late Middle Pleistocene locality of Hoedjiespunt 1, South Africa. Journal of Human Evolution, 41, 369383.Google Scholar
Sundevall, C. J. (1842). Öfversigt af slagtet Manis. Kungl. Svenska Vetenskapsakademiens Handlingar Stockholm, Series 3, 30, 245282.Google Scholar
Sundevall, C. J. (1845). Methodisk öfversigt af Idislande djuren, Linnés Pecora. Fam. 5 Antilopina cont. Kungl. Svenska Vetenskapsakademiens Handlingar, Series 3, 33, 267330.Google Scholar
Sundevall, C. J. (1846). Nya Mammalia, frän Sydafrika: Öfversigt af Kongl. Vetenskaps-Akademiens Förhandlingar, 3, 118123.Google Scholar
Susman, R. L. (1993). Hominid postcranial remains from Swartkrans. In Brain, C. K., ed. Swartkrans. A cave’s chronicle of early man. Transvaal Museum Monograph, 8, 117136.Google Scholar
Susman, R. L. & De Ruiter, D. (2004). New hominin first metatarsal (SK 1813) from Swartkrans. Journal of Human Evolution, 47, 171181.Google Scholar
Susman, R. L., De Ruiter, D. & Brain, C. K. (2001). Recently identified postcranial remains of Paranthropus and an early Homo from Swartkrans Cave, South Africa. Journal of Human Evolution, 41, 607629.Google Scholar
Swan, L. (2002). Excavations at Copper Queen Mine, Northwestern Zimbabwe. South African Archaeological Bulletin, 57, 6479.Google Scholar
Swanepoel, E. (2003). An analysis of the faunal remains of Kemp’s Caves and an investigation into possible computerized classification of bones. Unpublished MSc thesis, University of Pretoria.Google Scholar
Sydow, W. (1969). The discovery of a Boskop skull at Otjiseva, near Windhoek, South West Africa. South African Journal of Science, 65, 7782.Google Scholar
Sykes, W. H. (1831). [Cercopithecus albogularis]. Proceedings of the Committee of Science and Correspondence of the Zoological Society of London, 1, 106107.Google Scholar
Tassy, P. (1986). Nouveau Elephantoidea (Mammalia) dans le Miocene de Kenya. Cahiers de Paléontologie, 10, 1230.Google Scholar
Taylor, M. E. (1972). Ichneumia albicauda. Mammalian Species, 12, 14.Google Scholar
Taylor, M. E. (1975). Herpestes sanguineus. Mammalian Species, 65, 15.Google Scholar
Taylor, P. J. (2000). Patterns of chromosomal variation in southern African rodents. Journal of Mammalogy, 81, 317331.Google Scholar
Taylor, P. J. & Meester, J. (1993). Cynictis penicillata. Mammalian Species, 432, 17.Google Scholar
Taylor, P. J., Campbell, G. K., Meester, J., Willan, K. & Van Dyk, D. (1989). Genetic variation in the African rodent subfamily Otomyinae (Muridae): 1. Allozyme divergence among four species. South African Journal of Science, 85, 257262.Google Scholar
Taylor, P. J., Meester, J. & Kearney, T. (1993). The taxonomic status of Saunders’ vlei rat, Otomys saundersiae Roberts (Rodentia: Muridae: Otomyinae). Journal of African Zoology, 107, 571596.Google Scholar
Taylor, P. J., Rautenbach, I. L., Gordon, D., Sink, K. & Lotter, P. (1995). Diagnostic morphometrics and southern African distribution of two sibling species of tree rat, Thallomys paedulcus and Thallomys nigricauda (Rodentia: Muridae). Durban Museum Novitates, 20, 4962.Google Scholar
Taylor, P. J., Denys, C. & Mukerjee, M. (2004). Phylogeny of the African murid tribe Otomyini (Rodentia), based on morphological and allozyme evidence. Zoologica Scripta, 33, 389402.Google Scholar
Taylor, P. J., Maree, S., Van Sandwyk, J., Baxter, R. M. & Rambau, R. V. (2009). When is a species not a species? Uncoupled phenotypic, karyotypic and genotypic divergence in two species of South African laminate-toothed rats (Murinae: Otomyini). Journal of Zoology, 277, 317332.Google Scholar
Taylor, P. J., Stoffberg, S., Monadjem, A., et al. (2012). Four new bat species (Rhinolophus hildebrandtii complex) reflect Plio-Pleistocene divergence of dwarfs and giants across an Afromontane Archipelago. PLoS ONE, 7. DOI: 10.1371/journal.pone.0041744.Google Scholar
Temminck, C. J. (1820). Sur le genre hyène, et description d’une espèce nouvelle, découverte en Afrique. Annales Générales des Sciences Physiques, 3, 46.Google Scholar
Temminck, C. J. (1827). Monographies de Mammalogie. Paris: Dufour & d’Ocagne.Google Scholar
Temminck, C. J. (1832). In J. Smuts. Dissertatio Zoologica, Ennumerationem Mammalium Capensium. Leiden: Cyfveer.Google Scholar
Thackeray, A. I., Thackeray, J. F. & Beaumont, P. B. (1983). Excavations at the Blinkklipkop specularite mine near Postmasburg, northern Cape. South African Archaeological Bulletin, 38, 1725.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. (1984). Man, animals and extinctions: the analysis of Holocene faunal remains from Wonderwerk Cave, South Africa. Unpublished PhD thesis, Yale University.Google Scholar
Thackeray, J. F. (2010). Ancient DNA from fossil equids: a milestone in palaeogenetics. South African Journal of Science, 106. DOI: 10.4102/sajs.v106i1/2.111.Google Scholar
Thackeray, J. F. (2015). Faunal remains from Holocene deposits: Excavation 1, Wonderwerk Cave, South Africa. African Archaeological Review, 32, 729750.Google Scholar
Thackeray, J. F. & Brink, J. S. (2004). Damaliscus niro horns from Wonderwerk Cave and other Pleistocene sites: morphological and chronological considerations. Palaeontologia Africana, 40, 8993.Google Scholar
Thackeray, J. F. & Myer, S. (2004). Parapapio broomi and Parapapio jonesi from Sterkfontein: males and females of one species? Annals of the Transvaal Museum, 41, 7982.Google Scholar
Thackeray, J. F. & Watson, V. (1994). A preliminary account of faunal remains from Plover’s Lake. South African Journal of Science, 90, 231233.Google Scholar
Thackeray, J. F., De Ruiter, D. J., Berger, L. R. & Van der Merwe, N. J. (2001). Fossil hominins from Kromdraai: a revised list of specimens discovered since 1938. Annals of the Transvaal Museum, 38, 4356.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, 8389.Google Scholar
Thomas, H. (1981). Les bovides miocènes de la formation de Ngorara du Bassin de Baringo (Rift Valley, Kenya): 2. Proceedings of the Koninklijke Nederlandse Akademie van Wetenschappen Series B Physical Sciences, 84, 357375.Google Scholar
Thomas, O. (1883). On Mustela albinucha, Gray. Annals and Magazine of Natural History, Series 5, 11, 370371.Google Scholar
Thomas, O. (1892). On the mammals of Nyasaland: (first notice). Proceedings of the Zoological Society of London, 1892, 546554.Google Scholar
Thomas, O. (1894). Description of a new species of reed-rat (Aulacodus) from East Africa, with remarks on the milk-dentition of the genus. Annals and Magazine of Natural History, Series 6, 13, 202204.Google Scholar
Thomas, O. (1895a). Description of five new African shrews. Annals and Magazine of Natural History, Series 6, 16, 5155.Google Scholar
Thomas, O. (1895b). On African mole-rats of the genera Georychus and Myoscalops. Annals and Magazine of Natural History, Series 6, 16, 238241.Google Scholar
Thomas, O. (1896). On the mammals of Nyasaland: fourth notice. Proceedings of the Zoological Society of London, 1896, 788798.Google Scholar
Thomas, O. (1897a). Small mammals collected by Mr Alexander Whyte during his expedition to the Nyika Plateau and the Masuku Mountains, N. Nyasa. Proceedings of the Zoological Society of London, 1897, 430436.Google Scholar
Thomas, O. (1897b). On a new dormouse from Mashonaland. Annals and Magazine of Natural History, Series 6, 19, 388389.Google Scholar
Thomas, O. (1897c). On the mammals obtained by Mr A. Whyte in Nyasaland and presented to the British Museum by Sir H.H. Johnston, K.C.B.; being a fifth contribution to the mammal fauna of Nyasaland. Proceedings of the Zoological Society of London, 1897, 925939.Google Scholar
Thomas, O. (1902). On some new forms of Otomys. Annals and Magazine of Natural History, Series 7, 10, 311314.Google Scholar
Thomas, O. (1903). On a remarkable new hare from Cape Colony. Annals and Magazine of Natural History, Series 7, 11, 7879.Google Scholar
Thomas, O. (1904a). Three new bats, African and Asiatic. Annals and Magazine of Natural History, Series 7, 13, 384388.Google Scholar
Thomas, O. (1904b). On mammals from northern Angola collected by Dr W.J. Ansorge. Annals and Magazine of Natural History, Series 7, 13, 405421.Google Scholar
Thomas, O. (1909). New African mammals. Annals and Magazine of Natural History, Series 8, 4, 542553.Google Scholar
Thomas, O. (1910). New African mammals. Annals and Magazine of Natural History, Series 8, 5, 8392.Google Scholar
Thomas, O. (1915a). New African rodents and insectivores, mostly collected by Dr C. Christy for the Congo Museum. Annals and Magazine of Natural History, Series 8, 16, 146152.Google Scholar
Thomas, O. (1915b). List of mammals (excluding Ungulata) collected on the Upper Congo by Dr Christy for the Congo Museum, Tervuren. Annals and Magazine of Natural History, Series 8, 16, 465481.Google Scholar
Thomas, O. (1916). On the rats usually included in the genus Arvicanthis. Annals and Magazine of Natural History, Series 8, 18, 6770.Google Scholar
Thomas, O. (1917). Notes on Georychus and its allies. Annals and Magazine of Natural History, Series 8, 20, 441444.Google Scholar
Thomas, O. (1918a). New species of Gerbillus and Taterillus. Annals and Magazine of Natural History, Series 9, 2, 146151.Google Scholar
Thomas, O. (1918b). A revised classification of the Otomyinae, with descriptions of new genera and species. Annals and Magazine of Natural History, Series 9, 2, 203211.Google Scholar
Thomas, O. (1920). The generic position of Mus migricauda, Thos. and woosnami, Schwann. Annals and Magazine of Natural History, Series 9, 5, 140142.Google Scholar
Thomas, O. (1926). The generic position of certain African Muridae hitherto referred to Aethomys and Praomys. Annals and Magazine of Natural History, Series 9, 17, 174179.Google Scholar
Thomas, O. (1929). On mammals from the Kaokoveld, South West Africa, obtained during Capt Shortridge’s fifth Percy Sladen and Kaffrarian Museum Expedition. Proceedings of the Zoological Society of London, 1929, 99111.Google Scholar
Thomas, O. & Hinton, M. A. C. (1925). On mammals collected in 1923 by Capt G.C. Shortridge during the Percy Sladen and Kaffrarian Museum Expedition to South West Africa. Proceedings of the Zoological Society of London, 1925, 221246.Google Scholar
Thomas, O. & Schwann, H. (1904a). [A collection of mammals from Namaqualand presented to the British Museums by Mr C.D. Rudd]. Abstract of the Proceedings of the Zoological Society of London, 2, 56.Google Scholar
Thomas, O. & Schwann, H. (1904b). On a collection of mammals from British Namaqualand presented to the National Museum by Mr C.D. Rudd. Proceedings of the Zoological Society of London, 1904, 171183.Google Scholar
Thomas, O. & Schwann, H. (1905). [Mammals from Zululand]. Abstract of the Proceedings of the Zoological Society of London, 18, 23.Google Scholar
Thomas, O. & Schwann, H. (1906a). The Rudd exploration of South Africa: V. List of mammals obtained by Mr Grant in NE Transvaal. Proceedings of the Zoological Society of London, 1906, 575591.Google Scholar
Thomas, O. & Schwann, H. (1906b). [Mammals from the Zoutpansberg]. Abstract of the Proceedings of the Zoological Society of London, 33, 10.Google Scholar
Thomas, O. & Wroughton, R. C. (1908). The Rudd exploration of South Africa: X. List of mammals obtained by Mr Grant near Tette, Zambesia. Proceedings of the Zoological Society of London, 1908, 535552.Google Scholar
Thorp, C. (1979). Cattle from the Early Iron Age of Zimbabwe-Rhodesia. South African Journal of Science, 95, 461.Google Scholar
Thorp, C. (1984). A cultural interpretation of the faunal remains from Khami Hill ruin. In Hall, M., Avery, G., Avery, D. M., Wilson, M. L. & Humphreys, A. J. B., eds. Frontiers: Southern African Archaeology Today. Oxford: Archaeopress, pp. 266276.Google Scholar
Thorp, C. (2009). Excavations at Hlamba Mlonga Hill, Malilangwe Trust, south-eastern Zimbabwe. Journal of African Archaeology, 7, 191218.Google Scholar
Thorp, C. (2010). Induna Cave, southeastern Zimbabwe: a ‘contact period’ assemblage in a changing social landscape. Southern African Humanities, 22, 113147.Google Scholar
Thunberg, C. P. (1788–1789). Resa uti Europa Africa, Asia, forrattad aren 1770–1779. Uppsala: Edman.Google Scholar
Thunberg, C. P. (1811). Mammalia Capensia, recensita et illustrata. Mémoires de l’Académie impériale des sciences de St. Pétersbourg, 3, 299323.Google Scholar
Thunberg, C. P. (1820). Beskrifning och techning på ett nytt species, Hyaena Brunnea. Kungl. Svenska Vetenskapsakademiens Handlingar, Series 3, 8, 5965.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
Todd, N. E. (2005). Reanalysis of African Elephas recki: implications for time, space and taxonomy. Quaternary International, 126–128, 6572.Google Scholar
Todd, N. E. (2010). New phylogenetic analysis of the Family Elephantidae based on cranial–dental morphology. The Anatomical Record, 293, 7490.Google Scholar
Toerien, M. J. (1952). The fossil hyaenas of the Makapansgat Valley. South African Journal of Science, 48, 293300.Google Scholar
Toerien, M. J. (1955). A sabre-tooth cat from the Makapansgat valley. Palaeontologia Africana, 3, 4346.Google Scholar
Tomes, R. F. (1861). Notes on a collection of bats made by Mr Andersson in the Damara country, south-western Africa, with notices of some other African species. Proceedings of the Zoological Society of London, 1861, 3140.Google Scholar
Toussaint, M., Macho, G. A., Tobias, P. V., Partridge, T. C. & Hughes, A. R. (2003). The third partial skeleton of a late Pliocene hominin (Stw 431) from Sterkfontein, South Africa. South African Journal of Science, 99, 215223.Google Scholar
Trinkaus, E., Ruff, C. B. & Conroy, G. C. (1999). The anomalous archaic Homo femur from Berg Aukas, Namibia: a biomechanical assessment. American Journal of Physical Anthropology, 110, 379391.Google Scholar
Troussaert, E.-L. (1881). Catalogue des mammifères vivants et fossiles. Bulletin de la Société d’Etudes Scientifiques d’Angers, 10, 58212.Google Scholar
Turner, A. (1984). The interpretation of variation in fossil specimens of spotted hyaena (Crocuta crocuta Erxleben, 1777) from Sterkfontein Valley sites (Mammalia: Carnivora). Annals of the Transvaal Museum, 33, 399418.Google Scholar
Turner, A. (1986). Miscellaneous carnivore remains from Plio-Pleistocene deposits in the Sterkfontein Valley (Mammalia: Carnivora). Annals of the Transvaal Museum, 34, 203226.Google Scholar
Turner, A. (1987a). New fossil carnivore remains from the Sterkfontein hominid site (Mammalia: Carnivora). Annals of the Transvaal Museum, 34, 319347.Google Scholar
Turner, A. (1987b). Megantereon cultridens (Cuvier) (Mammalia, Felidae, Machairodontinae) from Plio-Pleistocene deposits in Africa and Eurasia, with comments on dispersal and the possibility of a New World origin. Journal of Paleontology, 61, 12561268.Google Scholar
Turner, A. (1988). On the claimed occurrence of the hyaenid genus Hyaenictis at Swartkrans (Mammalia: Carnivora). Annals of the Transvaal Museum, 34, 523533.Google Scholar
Turner, A. (1993). New fossil carnivore remains from Swartkrans. In Brain, C. K., ed. Swartkrans. A cave’s chronicle of early man. Transvaal Museum Monograph, 8, 151165.Google Scholar
Turner, A. (1997). Further remains of Carnivora (Mammalia) from the Sterkfontein Hominid Site. Palaeontologia Africana, 34, 115126.Google Scholar
Turner, A. & Antón, M. (1996). The giant hyaena Pachycrocuta brevirostris (Mammalia, Carnivora, Hyaenidae). Geobios, 29, 455468.Google Scholar
Turner, A., Bishop, L. C., Denys, C. & McKee, J. K. (1999). Appendix: a locality-based listing of African Plio-Pleistocene mammals. In Bromage, T. G. & Schrenk, F., eds. African Biogeography, Climate Change, & Human Evolution. New York: Oxford University Press, pp. 369399.Google Scholar
Turner, G. (1984). Vertebrate remains from Lekkerwater. South African Archaeological Bulletin, 39, 106108.Google Scholar
Turner, G. (1986). Faunal remains from Jubilee Shelter. South African Archaeological Bulletin, 41, 6368.Google Scholar
Turner, G. (1987a). Hunters and herders of the Okavango delta, Botswana. Botswana Notes and Records, 19, 2540.Google Scholar
Turner, G. (1987b). Early Iron Age herders in northwestern Botswana: the faunal evidence. Botswana Notes and Records, 19, 723.Google Scholar
Underhill, D. (2011). A history of Stone Age archaeological study in South Africa. South African Archaeological Bulletin, 66, 314.Google Scholar
Val, A. & 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. M., et al. (2011). 3D techniques and fossil identification: an elephant shrew hemi-mandible from the Malapa site. South African Journal of Science, 107. DOI: 10.4102/sajs.v107i11/12.583 .Google Scholar
Val, A., Dirks, P. H. G., Backwell, L., D’Errico, F. & Berger, L. R. (2014). Taphonomic analysis of the faunal assemblage associated with the hominins (Australopithecus sediba) from the Early Pleistocene cave deposits of Malapa, South Africa. PLoS ONE, 10. DOI: 10.1371/journal.pone.0126904.Google Scholar
Van Aardt, A. C., Bousman, C. B., Brink, J. S., et al. (2016). First chronological, palaeoenvironmental and archaeological data from the Baden-Baden fossil spring complex in the western Free State, South Africa. Palaeoecology of Africa, 33, 117152.Google Scholar
Van der Horst, C. J. (1944). Remarks on the systematics of Elephantulus. Journal of Mammalogy, 25, 7782.Google Scholar
Van der Merwe, M. (2007). Discriminating between Thryonomys swinderianus and Thryonomys gregorianus. African Zoology, 42, 165171.Google Scholar
Van der Ryst, M. M. (2006). Seeking shelter: Later Stone Age hunters, gatherers and fishers of Olieboomspoort in the western Waterberg, south of the Limpopo. Unpublished PhD thesis, University of the Witwatersrand.Google Scholar
Van Doornum, B. (2007). Tshisiku Shelter and the Shashe-Limpopo confluence area hunter-gatherer sequence. Southern African Humanities, 19, 1767.Google Scholar
Van Ewyk, J. F. (1987). The prehistory of an Iron Age site on Skutwater. Unpublished MA thesis, University of Pretoria.Google Scholar
Van Hoepen, E. C. N. (1930a). Vrystaatse fossiele perde. Paleontologiese Navorsinge van die Nasionale Museum Bloemfontein, 2, 111.Google Scholar
Van Hoepen, E. C. N. (1930b). Fossiele perde van Cornelia, O.V.S. Paleontologiese Navorsinge van die Nasionale Museum Bloemfontein, 2, 1334.Google Scholar
Van Hoepen, E. C. N. (1932a). Die stamlyn van die Sebras. Paleontologiese Navorsinge van die Nasionale Museum Bloemfontein, 2, 2537.Google Scholar
Van Hoepen, E. C. N. (1932b). Voorlopige beskrywing van Vrystaatse soogdiere. Paleontologiese 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. Paleontologiese Navorsinge van die Nasionale Museum Bloemfontein, 2, 103106.Google Scholar
Van Hoepen, E. C. N. & Van Hoepen, H. E. (1932). Vrystaatse wilde varke. Paleontologiese Navorsinge van die Nasionale Museum Bloemfontein, 2, 3962.Google Scholar
Van Neer, W. & Breunig, P. (1999). Contribution to the archaeozoology of the Brandberg. Cimbebasia, 15, 127140.Google Scholar
Van Noten, F. L. (1974). Excavations at the Gordon’s Bay Shell Midden, South-Western Cape. South African Archaeological Bulletin, 29, 122142.Google Scholar
Van Pletzen, L. (2000). The large mammal fauna from Klasies River. Unpublished MA thesis, University of Stellenbosch.Google Scholar
Van Schalkwyk, J., Pelser, A. & Teichert, F. (1999). Archaeological investigation of a Late Iron Age Tswana settlement on the farm Hoekfontein 432JQ, Odi District, North West Province. National Cultural History Museum Unpublished Report No. 98KH21.Google Scholar
Van Schalkwyk, J. A. (2000). Excavation of a Late Iron Age site in the Makgabeng, Northern Province. Southern African Field Archaeology, 9, 7582.Google Scholar
Van Staaden, M. J. (1994). Suricata suricatta. Mammalian Species, 483, 18.Google Scholar
Van Waarden, C. (1987). Matanga, a Late Zimbabwe cattle post. South African Archaeological Bulletin, 42, 107124.Google Scholar
Van Zyl, W. J., Badenhorst, S., Taljaard, E., Denbow, J. R. & Wilmsen, E. N. (2013). The archaeofauna from Xaro on the Okavango Delta in northern Botswana. Annals of the Ditsong National Museum of Natural History, 3, 4958.Google Scholar
Van Zyl, W. J., Badenhorst, S. & Brink, J. (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, 3973.Google Scholar
Verheyen, W. N., Huiselmans, J. L. J., Dierckx, T., et al. (2003). A craniometric and genetic approach to the systematics of the genus Dasymys Peters, 1875, selection of a neotype and description of three new taxa (Rodentia, Muridae, Africa). Bulletin de L’Institut Royal des Sciences Naturelles de Belgique, Biologie, 73, 2771.Google Scholar
Veyrunes, F., Britton-Davidian, J., Robinson, T. J., et al. (2005). Molecular phylogeny of the African pygmy mice, subgenus Nannomys (Rodentia, Murinae, Mus): implications for chromosomal evolution. Molecular Phylogenetics and Evolution, 36, 358369.Google Scholar
Viret, J. (1939). Monographie paléontologique de la faune de vertébrés des Sables de Montpellier. III. Carnivora, Fissipedia. Travaux du Laboratoire de Géologie de la Faculté des Sciences de Lyon, 37, 526.Google Scholar
Viriot, L., Pelaez-Campomanes, P., Vignaud, P., et al. (2011). A new Xerinae (Rodentia, Sciuridae) from the Late Miocene of Toros-Menalia (Chad). Journal of Vertebrate Paleontology, 31, 844848.Google Scholar
Visser, D. S. & Robinson, T. J. (1986). Cytosystematics of the South African Aethomys (Rodentia: Muridae). South African Journal of Zoology, 21, 264268.Google Scholar
Vogelsang, R., Richter, J., Jacobs, Z., et al. (2010). New excavations of Middle Stone Age deposits at Apollo 11 Rockshelter, Namibia: stratigraphy, archaeology, chronology and past environments. Journal of African Archaeology, 8, 185218.Google Scholar
Voigt, E. A. (1979). The faunal remains from Icon. South African Archaeological Society Goodwin Series, 3, 8085.Google Scholar
Voigt, E. A. (1980a). Appendix. the faunal sample from Msuluzi Confluence. In Maggs, T.. Msuluzi Confluence: a seventh century Early Iron Age site on the Tugela River. Annals of the Natal Museum, 24, 140145.Google Scholar
Voigt, E. A. (1980b). Reconstructing Iron Age economies of the Northern Transvaal: a preliminary report. South African Archaeological Bulletin, 35, 3945.Google Scholar
Voigt, E. A. (1980c). Appendix 1: mammalian remains from Mzonjani. In Maggs, T.. Mzonjani and the beginning of the Iron Age in Natal. Annals of the Natal Museum, 24, 9495.Google Scholar
Voigt, E. A. (1982). Appendix I: faunal report on the Lydenberg Heads Site, 2539 AB4. In Evers, T. M.. Excavations at the Lydenburg Heads Site, Eastern Transvaal, South Africa. South African Archaeological Bulletin, 37, 3132.Google Scholar
Voigt, E. A. (1983). Mapungubwe: an Archaeozoological Interpretation of an Iron Age Community. Pretoria: Transvaal Museum.Google Scholar
Voigt, E. A. (1984). The faunal remains from Magogo and Mhlopeni: small stock herding in the Early Iron Age of Natal. Annals of the Natal Museum, 26, 141163.Google Scholar
Voigt, E. A. & Peters, J. H. (1994a). Appendix 2: the faunal assemblage from Wosi in the Thukela Valley. In van Schalkwyk, L.. Wosi: an Early Iron Age village in the lower Thukela Basin, Natal. Natal Museum Journal of Humanities, 6, 105117.Google Scholar
Voigt, E. A. & Peters, J. H. (1994b). Appendix: the faunal assemblage from the Early Iron Age site of Mamba 1 in the Thukela Valley. In van Schalkwyk, L.. Mamba confluence: a preliminary report on an Early Iron Age industrial centre in the lower Thukela Basin, Natal. Natal Journal of Humanities, 6, 145152.Google Scholar
Voigt, E. A. & Plug, I. (1981). Early Iron Age Herders of the Limpopo Valley. Pretoria: Transvaal Museum.Google Scholar
Voigt, E. A. & Plug, I. (1984). Happy Rest: the earliest Iron Age from the Soutpansberg. South African Journal of Science, 80, 221227.Google Scholar
Voigt, E. A. & Von den Driesch, A. (1984). Preliminary report on the faunal assemblage from Ndondondwane, Natal. Annals of the Natal Museum, 26, 95104.Google Scholar
Voigt, E. A., Plug, I. & Sampson, C. G. (1995). European livestock from rock shelters in the upper Seacow valley. Southern African Field Archaeology, 4, 3749.Google Scholar
Von den Driesch, A. & Deacon, H. J. (1985). Sheep remains from Boomplaas, South Africa. South African Archaeological Bulletin, 40, 3944.Google Scholar
Von Dueben, W. (1846). Om Nesotragus moschatus n. sp. In Sundevall, C. J.. Nya Mammalia, från Sydafrika. Öfversigt af Kongl. Vetenskaps-akademiens forhandlingar, Series 3, 1846, 221222.Google Scholar
Von Mayer, A. (1998). A reassessment of Cercopithecoides in southern Africa. Unpublished MSc thesis, University of the Witwatersrand.Google Scholar
Von Richter, W. (1974). Connochaetes gnou. Mammalian Species, 50, 16.Google Scholar
Vrba, E. S. (1970). Evaluation of springbok-like fossils: measurement and statistical treatment of the teeth of the springbok Antidorcas marsupialis Zimmerman (Artiodactyla: Bovidae). Annals of the Transvaal Museum, 26, 285299.Google Scholar
Vrba, E. S. (1971). A new fossil alcelaphine (Artiodactyla: Bovidae) from Swartkrans. Annals of the Transvaal Museum, 27, 5982.Google Scholar
Vrba, E. S. (1973). Two new species of Antidorcas Sundevall at Swartkrans (Mammalia: Bovidae). Annals of the Transvaal Museum, 28, 287352.Google Scholar
Vrba, E. S. (1974a). Chronological and ecological implications of the fossil Bovidae at the Sterkfontein australopithecine site. Nature, 250, 1923.Google Scholar
Vrba, E. S. (1974b). The fossil Bovidae of Sterkfontein, Swartkrans and Kromdraai. Unpublished PhD thesis, University of Cape Town.Google Scholar
Vrba, E. S. (1976). The Fossil Bovidae of Sterkfontein, Swartkrans and Kromdraai. Memoir of the Transvaal Museum, 21.Google Scholar
Vrba, E. S. (1977). New species of Parmularius Hopwood and Damaliscus Sclater (Alcelaphini, Bovidae, Mammalia) from Makapansgat. Palaeontologia Africana, 20, 137151.Google Scholar
Vrba, E. S. (1978). Problematical alcelaphine fossils from the Kromdraai faunal site (Mammalia, Bovidae). Annals of the Transvaal Museum, 31, 2128.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. S. (1987a). A revision of the Bovini (Bovidae) and a preliminary revised checklist of Bovidae from Makapansgat (Transvaal: South Africa). Palaeontologia Africana, 26, 3346.Google Scholar
Vrba, E. S. (1987b). New species and a new genus of Hippotragini (Bovidae) from Makapansgat Limeworks. Palaeontologia Africana, 26, 4758.Google Scholar
Vrba, E. S. (1997). New fossils of Alcelaphini and Caprinae (Bovidae: Mammalia) from Awash, Ethiopia and phylogenetic analysis of Alcelaphini. Palaeontologia Africana, 34, 127198.Google Scholar
Vrba, E. S. & Panagos, D. C. (1978). A new limestone cave breccia from Vlakplaats near Pretoria. Annals of the Transvaal Museum, 31, 177183.Google Scholar
Wadley, L. (1979). Big Elephant Shelter and its role in the Holocene prehistory of central South West Africa. Cimbebasia B, 3, 176.Google Scholar
Wadley, L. (2001). Who lived in Mauermanshoek Shelter, Korannaberg, South Africa? African Archaeological Review, 18, 153179.Google Scholar
Wadley, L. (2015). Those marvellous millennia: the Middle Stone Age of Southern Africa. Azania, 50, 155226.Google Scholar
Wadley, L. & Laue, G. (2000). Adullam Cave, eastern Free State, South Africa: test excavations at a multiple-occupation Oakhurst Industry site. Natal Journal of Humanities, 12, 113.Google Scholar
Wadley, L. & McLaren, G. (1998). Tandjesberg Shelter, eastern Free State, South Africa. Natal Museum Journal of Humanities, 10, 1932.Google Scholar
Wadley, L. & Turner, G. (1987). Hope Hill Shelter: a Later Stone Age site in the southern Transvaal. South African Journal of Science, 83, 98105.Google Scholar
Wadley, L., Plug, I. & Clark, J. L. (2008). The contribution of Sibudu fauna to an understanding of KwaZulu-Natal environments at ~60 ka, ~50 ka and ~37 ka. In Badenhorst, S., Mitchell, P. & Driver, J. C., eds. Animals and People: Archaeozoological Papers in Honour of Ina Plug. Oxford: Archaeopress, pp. 3445.Google Scholar
Wagler, J. G. (1832). Mittheilungen über werkwürbige Thiere. Isis von Oken, 25, 275.Google Scholar
Wagner, J. A. (1839). Ueber die Berwandtschafts = Berhältnisse der Pharaonstratte. Gelehrte Anzeigen Königlich Bayerische Akademie der Wissenschaften zu München, 9, cols 425–432.Google Scholar
Wagner, J. A. (1841). Gruppirung der Gattungen der Nager. Gelehrte Anzeigen Königlich Bayerische Akademie der Wissenschaften zu München, 12, cols 433–440.Google Scholar
Wagner, J. A. (1843). Malacothrix. Die Didmaus. In Die Säugthiere in Abbildungen nach der Natur von Dr Johann Christian Daniel von Schreber, Supplement 3. Erlangen: Expedition des Schreber’schen säugthier- und des Esper’schen Schmetterlingswerkes, pp. 496499.Google Scholar
Wagner, J. A. (1845). Diagnosen einiger neuen Arten von Nagern und Handflüglern. Archiv für Naturgeschichte, 11, 145149.Google Scholar
Wagner, J. A. (1855). Antilope strepsicerinae. Schrauben-Antipolen. In Die Säugthiere in Abbildungen nach der Natur, mit Beschreibungen von Dr Johann Christian Daniel von Schreber, Vol. 5. Leipzig: Weigel, pp. 438461Google Scholar
Walker, N. (1994). The Late Stone Age of Botswana: some recent excavations. Botswana Notes and Records, 26, 135.Google Scholar
Walker, N. J. (1983). The significance of an early date for pottery and sheep in Zimbabwe. South African Archaeological Bulletin, 38, 8892.Google Scholar
Waterhouse, G. R. (1837). [A species of mouse from the Cape of Good Hope]. Proceedings of the Zoological Society of London, 2, 104105.Google Scholar
Watson, E. J. & Watson, V. (1990). Of commoners and kings: faunal remains from Ondini. South African Archaeological Bulletin, 45, 3346.Google Scholar
Watson, V. (1993a). Composition of the Swartkrans bone accumulations, in terms of skeletal parts and animals represented. In Brain, C. K., ed. Swartkrans: A cave’s chronicle of early man. Transvaal Museum Monograph 8, 3573.Google Scholar
Watson, V. (1993b). Glimpses from Gondolin: a faunal analysis of a fossil site near Broederstroom, Transvaal, South Africa. Palaeontologia Africana, 30, 3542.Google Scholar
Watson, V. & Plug, I. (1995). Oreotragus major Wells and Oreotragus oreotragus (Zimmerman) (Mammalia: Bovidae): two species? Annals of the Transvaal Museum, 36, 183191.Google Scholar
Webb, G. L. (1965). Notes on some chalicothere remains from Makapansgat. Palaeontologia Africana, 9, 4968.Google Scholar
Webb, P. I. & Skinner, J. D. (1995). The dormice (Myoxidae) of southern Africa. Hystrix, New Series, 6, 287293.Google Scholar
Webley, L. (1992a). Early evidence for sheep from Spoeg River Cave, Namaqualand. Southern African Field Archaeology, 1, 313.Google Scholar
Webley, L. E. (1992b). The history and archaeology of pastoralist and hunter-gatherer settlement in the north-western Cape, South Africa. Unpublished PhD thesis, University of Cape Town.Google Scholar
Webley, L. (2001a). Excavations at /hei-/khomas (Vaalhoek) in the Richtersveld, Northern Cape. Southern African Field Archaeology, 10, 4674.Google Scholar
Webley, L. (2001b). The re-excavation of Spoegrivier Cave on the West Coast of South Africa. Annals of the Eastern Cape Museums, 2, 1949.Google Scholar
Webley, L., Archer, F. & Brink, J. (1993). Die Toon: a Late Holocene site in the Richtersveld National Park, northern Cape. Koedoe, 36, 19.Google Scholar
Weithofer, K. A. (1889). Ueber die tertiären Landsäugethiere Italiens. Jahrbuch der Kaiserlich Königlichen Geologischen Reichsanstalt, 39, 5582.Google Scholar
Welbourne, R. (1973). Identification of animal remains from the Broederstroom 24/73 Early Iron Age site. South African Journal of Science, 69, 325.Google Scholar
Welbourne, R. G. (1974). Appendix C: animal remains from the Harmony 24 Iron Age sites. In T. M. Evers. Three Iron Age industrial sites in the Eastern Transvaal Lowveld. Unpublished MA thesis, University of the Witwatersrand, pp. 100–104.Google Scholar
Welbourne, R. G. (1975). Tautswe Iron Age Site: its yield of bones. Botswana Notes and Records, 7, 116.Google Scholar
Welbourne, R. G. (1979). Animals remains from Harmony Salt Factory, Northern Transvaal. South African Archaeological Society Goodwin Series, 3, 108.Google Scholar
Welbourne, R. G. (1985). Faunal analysis. In Sinclair, P. J. J.. Excavations at Chivowa Hill, south central Zimbabwe. Zimbabwea, 9, 3844.Google Scholar
Welbourne, R. G. (1988). Appendix 1: an analysis of the animal remains recovered from Rose Cottage Cave. In Kohary, J., ed. Rose Cottage Cave. Johannesburg: Archaeological Research Unit, University of the Witwatersrand, pp. 133140.Google Scholar
Wells, C. R. (2006). A sample integrity analysis of faunal remains from the RSp layer at Sibudu Cave. Southern African Humanities, 18, 261277.Google Scholar
Wells, L. H. (1940). A fossil horse from Koffiefontein, O.F.S. Transactions of the Royal Society of South Africa, 28, 301306.Google Scholar
Wells, L. H. (1951). Large fossil klipspringer from Potgietersrust. South African Journal of Science, 47, 167168.Google Scholar
Wells, L. H. (1959a). The nomenclature of South African fossil equids. South African Journal of Science, 55, 6466.Google Scholar
Wells, L. H. (1959b). The Quaternary giant hartebeests of South Africa. South African Journal of Science, 55, 123128.Google Scholar
Wells, L. H. (1959c). Mammalian fossils from Barkly West in the Natal Museums, Pietermaritzburg. South African Journal of Science, 55, 146.Google Scholar
Wells, L. H. (1960). Mammalian remains from Late Stone Age sites in the George-Knysna area. South African Journal of Science, 56, 306.Google Scholar
Wells, L. H. (1964). A large extinct antelope skull from the Younger Gravels at Sydney-on-Vaal, C.P. South African Journal of Science, 60, 8891.Google Scholar
Wells, L. H. (1965). Antelopes in the Pleistocene of southern Africa. Zoologica Africana, 1, 115120.Google Scholar
Wells, L. H. (1970a). The fauna of the Aloes Bone Deposit: a preliminary note. South African Archaeological Bulletin, 25, 2223.Google Scholar
Wells, L. H. (1970b). A Late Pleistocene faunal assemblage from Driefontein, Cradock District, C.P. South African Journal of Science, 66, 5961.Google Scholar
Wells, L. H. (1988). Pre-occupation microfauna of the Cave of Hearths. In Mason, R., ed. Cave of Hearths, Makapansgat, Transvaal. Johannesburg: Archaeological Research Unit, University of the Witwatersrand, pp. 549550.Google Scholar
Wells, L. H. & Cooke, H. B. S. (1955). Fossil remains from Chelmer, near Bulawayo, S. Rhodesia: a further note. South African Journal of Science, 52, 49.Google Scholar
Wells, L. H. & Cooke, H. B. S. (1957). Fossil Bovidae from the Limeworks Quarry, Makapansgat, Potgietersrus. Palaeontologia Africana, 4, 155.Google Scholar
Wells, L. H. & Gear, J. H. (1931). Skeletal material from early graves in the Riet river valley. South African Journal of Science, 28, 435443.Google Scholar
Wells, L. H., Cooke, H. B. S. & 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
Werdelin, L. & Cote, S. M. (2010). Prionogalidae (Mammalia, Incertae Sedis). In Werdelin, L. & Sanders, W. J., eds. Cenozoic Mammals of Africa. Berkeley, CA: University of California Press, pp. 561562.Google Scholar
Werdelin, L. & Lewis, M. E. (2001). A revision of the genus Dinofelis (Mammalia, Felidae). Zoological Journal of the Linnean Society, 32, 147258.Google Scholar
Werdelin, L. & Peigné, S. (2010). Carnivora. In Werdelin, L. & Sanders, W. J., eds. Cenozoic Mammals of Africa. Berkeley, CA: University of California Press, pp. 603657.Google Scholar
Werdelin, L. & Sanders, H. L. (2010). Cenozoic Mammals of Africa. Berkeley, CA: University of California Press.Google Scholar
Werdelin, L. & Sardella, R. (2006). The Homotherium from Langebaanweg, South Africa and the origin of Homotherium. Palaeontographica Abt A, 277, 123130.Google Scholar
Werdelin, L. & Solounias, N. (1991). The Hyaenidae: taxonomy, systematics and evolution. Fossils and Strata, 30, 1104.Google Scholar
Werdelin, L., Turner, A. & Solounias, N. (1994). Studies of fossil hyaenids: the genera Hyaenictis Gaudry and Chasmaporthetes Hay, with a reconsideration of the Hyaenidae of Langebaanweg, South Africa. Zoological Journal of the Linnean Society, 111, 197217.Google Scholar
Weston, E. & Boisserie, J.-R. (2010). Hippopotamidae. In Werdelin, L. & Sanders, W. J., eds. Cenozoic Mammals of Africa. Berkeley, CA: University of California Press, pp. 853871.Google Scholar
Whitelaw, G. (1994). KwaGandaganda: settlement patterns in the Natal Early Iron Age. Natal Journal of Humanities, 6, 164.Google Scholar
Whitworth, T. (1958). Miocene ruminants of East Africa. Fossil Mammals of Africa, 15, 150.Google Scholar
Wilkinson, A. F. (1976). The lower Miocene Suidae of Africa. Fossil Vertebrates of Africa, 4, 173282.Google Scholar
Williams, B. A., Ross, C. A., Frost, S. R., et al. (2012). Fossil Papio cranium from !Ncumtsa (Koanaka) Hills, western Ngamiland, Botswana. American Journal of Physical Anthropology, 149, 117.Google Scholar
Willows-Munro, S. & Matthee, C. A. (2009). The evolution of the southern African members of the shrew genus Myosorex: understanding the origin and diversification of a morphologically cryptic group. Molecular Phylogenetics and Evolution, 51, 394398.Google Scholar
Willows-Munro, S., Robinson, T. J. & Matthee, C. A. (2005). Utility of nuclear DNA intron markers at lower taxonomic levels: phylogenetic resolution among nine Tragelaphus spp. Molecular Phylogenetics and Evolution, 35, 624636.Google Scholar
Wilmsen, E. N. (1989). The antecedents of contemporary pastoralism in western Ngamiland. Botswana Notes and Records, 20, 2939.Google Scholar
Wilson, D. E. & Reeder, D. M. (2005). Mammal Species of the World, 3rd edition. Baltimore, MD: Johns Hopkins University Press.Google Scholar
Wilson, M. L. (1988). Forest Hall Shelter: an early excavation on the southern Cape coast. South African Archaeological Bulletin, 43, 5355.Google Scholar
Wood, A. E. (1968). Early Cenozoic Mammalian faunas, Fayum Province, Egypt. Part II: the African Oligocene Rodentia. Bulletin of the Peabody Museum of Natural History, Yale University, 28, 23105.Google Scholar
Woodward, A. S. (1921). A new cave man from Rhodesia, South Africa. Nature, 108, 371372.Google Scholar
Wroughton, R. C. (1905). Notes on the various forms of Arvicanthis pumilio Sparrm. Annals and Magazine of Natural History, Series 7, 16, 629639.Google Scholar
Wroughton, R. C. (1906). Notes on the genus Otomys. Annals and Magazine of Natural History, Series 7, 18, 264278.Google Scholar
Wroughton, R. C. (1909). New species of Dendromus & Tatera. Annals and Magazine of Natural History, Series 8, 3, 246249.Google Scholar
Yellen, J. E., Brooks, A. S., Stuckenrath, R. & Welbourne, R. (1987). A Terminal Pleistocene assemblage from Drotsky’s Cave, western Ngamiland, Botswana. Botswana Notes and Records, 19, 16.Google Scholar
Zdansky, O. (1924). Jungtertiäre Carnivoren Chinas. Palaeontologia Sinica Series C, 2, 1149.Google Scholar
Zimmermann, J. C. (1779–1783). Geographische Geschichte des Menschen und der Allgemein Verbreiteten Vierfüssigen Thiere. Leipzig: Weygandschen.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.

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.

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.

Available formats
×