Skip to main content Accessibility help
×
Hostname: page-component-848d4c4894-xm8r8 Total loading time: 0 Render date: 2024-07-07T11:50:49.933Z Has data issue: false hasContentIssue false

Part III - The Big Mammal Menagerie: Herbivores, Carnivores and Their Ecosystem Impacts

Published online by Cambridge University Press:  09 September 2021

Norman Owen-Smith
Affiliation:
University of the Witwatersrand, Johannesburg
Get access

Summary

Africa’s large herbivore assemblage is the product of the environments that also nurtured human origins and these animals contributed to evolutionary transitions in our hominin lineage. Around 90 species of large herbivore can be tallied on the continent, around half of them associated with the savanna biome (Figure III.1; the rest live in forests or deserts). All of them are ungulates, representing the orders Artiodactyla (with even-toes) and Perissodactyla (with odd-toes), except for the African elephant (Proboscidea). They span a range in body mass from ~5000-kg elephants down to 5-kg dikdiks.

Type
Chapter
Information
Only in Africa
The Ecology of Human Evolution
, pp. 141 - 246
Publisher: Cambridge University Press
Print publication year: 2021

Access options

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

References

Owen-Smith, RN. (1988) Megaherbivores: The Influence of Very Large Body Size on Ecology. Cambridge University Press, Cambridge.Google Scholar
Owen-Smith, N. (1990) Demography of a large herbivore, the greater kudu Tragelaphus strepsiceros, in relation to rainfall. The Journal of Animal Ecology 59:893913.Google Scholar
Owen-Smith, N; Cooper, SM. (1987) Palatability of woody plants to browsing ruminants in a South African savanna. Ecology 68:319331.CrossRefGoogle Scholar
Owen-Smith, N. (1994) Foraging responses of kudus to seasonal changes in food resources: elasticity in constraints. Ecology 75:10501062.Google Scholar
Owen-Smith, RN. (2002) Adaptive Herbivore Ecology: From Resources to Populations in Variable Environments. Cambridge University Press, Cambridge.Google Scholar
Owen-Smith, N. (2009) Dynamics of Large Herbivore Populations in Changing Environments: Towards Appropriate Models. John Wiley & Sons, Chichester.Google Scholar

Suggested Further Reading

Du Toit, JT; Cumming, DHM. (1999) Functional significance of ungulate diversity in African savannas and the ecological implications of the spread of pastoralism. Biodiversity and Conservation 8:16431661.Google Scholar
Kingdon, J; Hoffmann, M (eds) (2013) The Mammals of Africa. Vols I–VI. A & C Black, London.Google Scholar
Owen-Smith, N. (2002) Adaptive Herbivore Ecology. From Resources to Populations in Variable Environments. Cambridge University Press, Cambridge.CrossRefGoogle Scholar

References

Owen-Smith, N; Chafota, J. (2012) Selective feeding by a megaherbivore, the African elephant (Loxodonta africana). Journal of Mammalogy 93:698705.Google Scholar
Barnes, RFW. (1982) Elephant feeding behaviour in Ruaha National Park, Tanzania. African Journal of Ecology 20:123136.CrossRefGoogle Scholar
Illius, AW; O’Connor, TG. (2000) Resource heterogeneity and ungulate population dynamics. Oikos 89:283294.CrossRefGoogle Scholar
Owen-Smith, RN. (1988) Megaherbivores: The Influence of Very Large Body Size on Ecology. Cambridge University Press, Cambridge.Google Scholar
Yoganand, K; Owen‐Smith, N. (2014) Restricted habitat use by an African savanna herbivore through the seasonal cycle: key resources concept expanded. Ecography 37:969982.Google Scholar
Macandza, VA, et al. (2012) Habitat and resource partitioning between abundant and relatively rare grazing ungulates. Journal of Zoology 287:175185.Google Scholar
Macandza, VA. (2003) Forage selection of African buffalo through the late dry season in the Satara region of the Kruger National Park. South African Journal of Wildlife Research 34: 113121.Google Scholar
Staver, AC, et al. (2019) Grazer movements exacerbate grass declines during drought in an African savanna. Journal of Ecology 107:14821491.CrossRefGoogle Scholar
Owen-Smith, N; Cooper, SM. (1987) Palatability of woody plants to browsing ruminants in a South African savanna. Ecology 68:319331.Google Scholar
Owen‐Smith, N; Cooper, SM. (1989) Nutritional ecology of a browsing ruminant, the kudu (Tragelaphus strepsiceros), through the seasonal cycle. Journal of Zoology 219:2943.CrossRefGoogle Scholar
Owen-Smith, N. (1994) Foraging responses of kudus to seasonal changes in food resources: elasticity in constraints. Ecology 75:10501062.Google Scholar
Owen-Smith, RN. (2002) Adaptive Herbivore Ecology: From Resources to Populations in Variable Environments. Cambridge University Press, Cambridge.CrossRefGoogle Scholar
McNaughton, SJ. (1988) Mineral nutrition and spatial concentrations of African ungulates. Nature 334:343345.Google Scholar
McNaughton, SJ, et al. (1997) Promotion of the cycling of diet-enhancing nutrients by African grazers. Science 278:17981800.Google Scholar
Stock, WD, et al. (2010) Herbivore and nutrient control of lawn and bunch grass distributions in a southern African savanna. Plant Ecology 206:1527.CrossRefGoogle Scholar
Grant, CC, et al. (2000) Nitrogen and phosphorus concentration in faeces: an indicator of range quality as a practical adjunct to existing range evaluation methods. African Journal of Range and Forage Science 17:8192.Google Scholar
Grant, CC, et al. (1996) The usefulness of faecal phosphorus and nitrogen in interpreting differences in live-mass gain and the response to P supplementation in grazing cattle in arid regions. Onderstepoort Journal of Veterinary Research 63:121126.Google Scholar
Gordon, IJ; Illius, AW. (1988) Incisor arcade structure and diet selection in ruminants. Functional Ecology 2:1522.Google Scholar
Janis, CM; Ehrhardt, D. (1988) Correlation of relative muzzle width and relative incisor width with dietary preference in ungulates. Zoological Journal of the Linnean Society 92:267284.CrossRefGoogle Scholar
Mendoza, M; Palmqvist, P. (2008) Hypsodonty in ungulates: an adaptation for grass consumption or for foraging in open habitat? Journal of Zoology 274:134142.CrossRefGoogle Scholar
Damuth, J; Janis, CM. (2011) On the relationship between hypsodonty and feeding ecology in ungulate mammals, and its utility in palaeoecology. Biological Reviews 86:733758.Google Scholar
Codron, D, et al. (2008) Functional differentiation of African grazing ruminants: an example of specialized adaptations to very small changes in diet. Biological Journal of the Linnean Society 94:755764.Google Scholar
Schubert, BW, et al. (2006) Microwear evidence for Plio–Pleistocene bovid diets from Makapansgat Limeworks Cave, South Africa. Palaeogeography, Palaeoclimatology, Palaeoecology 241:301319.CrossRefGoogle Scholar
Ungar, PS. (2017) Evolution’s Bite: A Story of Teeth, Diet, and Human Origins. Princeton University Press, Princeton, NJ.Google Scholar
Hofmann, RR; Stewart, DRM. (1972) Grazer or browser: a classification based on the stomach-structure and feeding habits of East African ruminants. Mammalia 36:226240.Google Scholar
Clauss, M, et al. (2008) The morphophysiological adaptations of browsing and grazing mammals. In Gordon, IJ; Prins, HHT (eds) The Ecology of Browsing and Grazing. Springer, Berlin, pp. 4788.Google Scholar
Hofmann, RR. (1989) Evolutionary steps of ecophysiological adaptation and diversification of ruminants: a comparative view of their digestive system. Oecologia 78:443457.CrossRefGoogle ScholarPubMed
Codron, D; Clauss, M. (2010) Rumen physiology constrains diet niche: linking digestive physiology and food selection across wild ruminant species. Canadian Journal of Zoology 88:11291138.CrossRefGoogle Scholar
Langer, P. (1988) The Mammalian Herbivore Stomach: Comparative Anatomy, Function and Evolution. Gustav Fischer, Jena.Google Scholar
Duncan, P, et al. (1990) Comparative nutrient extraction from forages by grazing bovids and equids: a test of the nutritional model of equid/bovid competition and coexistence. Oecologia 84:411418.CrossRefGoogle ScholarPubMed
Meyer, K. (2010) The relationship between forage cell wall content and voluntary food intake in mammalian herbivores. Mammal Review 40:221245.Google Scholar
Clauss, M, et al. (2007) The relationship of food intake and ingesta passage predicts feeding ecology in two different megaherbivore groups. Oikos 116:209216.CrossRefGoogle Scholar
Cooper, SM; Owen-Smith, N. (1985) Condensed tannins deter feeding by browsing ruminants in a South African savanna. Oecologia 67:142146.CrossRefGoogle Scholar
Owen-Smith, N. (1993) Woody plants, browsers and tannins in southern African savannas. South African Journal of Science 89:505510.Google Scholar
Farrell-Gray, CC; Gotelli, NJ. (2005) Allometric exponents support a 3/4‐power scaling law. Ecology 86:20832087.Google Scholar
Bell, RHV. (1971) A grazing ecosystem in the Serengeti. Scientific American 225:8693.Google Scholar
Jarman, P. (1974) The social organisation of antelope in relation to their ecology. Behaviour 48:215267.CrossRefGoogle Scholar
Owen-Smith, N; Cooper, SM. (1985) Comparative consumption of vegetation components by kudus, impalas and goats in relation to their commercial potential as browsers in savanna regions. South African Journal of Science 81:7276.Google Scholar
Cooper, SM; Owen-Smith, N. (1986) Effects of plant spinescence on large mammalian herbivores. Oecologia 68:446455.CrossRefGoogle ScholarPubMed
Owen-Smith, N. (1985) Niche separation among African ungulates. In Vrba, ES (ed.) Species and Speciation. Transvaal Museum Monograph 4. Transvaal Museum, Pretoria, pp. 167171.Google Scholar
Arsenault, R; Owen‐Smith, N. (2008) Resource partitioning by grass height among grazing ungulates does not follow body size relation. Oikos 117:17111717.Google Scholar
Heitkonig, IMA. (1993) Feeding strategy of roan antelope in a low nutrient savanna. PhD thesis, University of the Witwatersrand, Johannesburg.Google Scholar
O’Shaughnessy, R, et al. (2014) Comparative diet and habitat selection of puku and lechwe in northern Botswana. Journal of Mammalogy 95:933942.Google Scholar
Fuller, A, et al. (2016) Towards a mechanistic understanding of the responses of large terrestrial mammals to heat and aridity associated with climate change. Climate Change Responses 3:119.Google Scholar
Finch, VA. (1972) Thermoregulation and heat balance of the East African eland and hartebeest. American Journal of Physiology 222:13741379.Google Scholar
Owen‐Smith, N. (1998) How high ambient temperature affects the daily activity and foraging time of a subtropical ungulate, the greater kudu (Tragelaphus strepsiceros). Journal of Zoology 246:183192.CrossRefGoogle Scholar
Owen‐Smith, N. (2000) Modeling the population dynamics of a subtropical ungulate in a variable environment: rain, cold and predators. Natural Resource Modeling 13:5787.CrossRefGoogle Scholar
Dunkin, RC, et al. (2013) Climate influences thermal balance and water use in African and Asian elephants: physiology can predict drivers of elephant distribution. Journal of Experimental Biology 216:29392952.Google Scholar
Woodall, PF; Skinner, JD. (1993) Dimensions of the intestine, diet and faecal water loss in some African antelope. Journal of Zoology 229:457471.Google Scholar
Kihwele, ES, et al. (2020) Quantifying water requirements of African ungulates through a combination of functional traits. Ecological Monographs 90:e01404.Google Scholar
Cain III, JW, et al. (2012) The costs of drinking: comparative water dependency of sable antelope and zebra. Journal of Zoology 286:5867.Google Scholar
Naidoo, R, et al. (2020) Mapping and assessing the impact of small‐scale ephemeral water sources on wildlife in an African seasonal savannah. Ecological Applications 30:e02203.CrossRefGoogle Scholar
Manser, MB; Brotherton, PNM. (1995) Environmental constraints on the foraging behaviour of a dwarf antelope (Madoqua kirkii). Oecologia 102:404412.Google Scholar
Knight, M. (2013) Oryx gazella. In Kingdon, J; Hoffman, AM (eds) The Mammals of Africa VI Pigs, Hippopotamuses, Chevrotain, Giraffes, Deer and Bovids. A & C Black, London, pp. 572576.Google Scholar
Western, D. (1975) Water availability and its influence on the structure and dynamics of a savannah large mammal community. African Journal of Ecology 13:265286.Google Scholar
Smit, IPJ, et al. (2007) Do artificial waterholes influence the way herbivores use the landscape? Herbivore distribution patterns around rivers and artificial surface water sources in a large African savanna park. Biological Conservation 136:8599.Google Scholar
Gereta, E; Wolanski, E. (1998) Wildlife–water quality interactions in the Serengeti National Park, Tanzania. African Journal of Ecology 36:114.CrossRefGoogle Scholar
Ledger, HP. (1968) Body composition as a basis for a comparative study of some East African mammals. Symposia of the Zoological Society of London 21:289310.Google Scholar
Smith, NS. (1970) Appraisal of condition estimation methods for East African ungulates. East African Wildlife Journal 8:123129.Google Scholar
Sinclair, ARE; Duncan, P. (1972) Indices of condition in tropical ruminants. African Journal of Ecology 10:143149.Google Scholar
Dunham, KM; Murray, MG. (1982) The fat reserves of impala, Aepycevos melampus. African Journal of Ecology 20:8187.Google Scholar
Sinclair, ARE; Arcese, P. (1995) Population consequences of predation‐sensitive foraging: the Serengeti wildebeest. Ecology 76:882891.Google Scholar
Selous, FC. (1881) A Hunter’s Wanderings in Africa. R. Bentley & Son, London.Google Scholar
Curtin, NA, et al. (2018) Remarkable muscles, remarkable locomotion in desert-dwelling wildebeest. Nature 563:393396.Google Scholar
Stabach, JA, et al. (2016) Variation in habitat selection by white‐bearded wildebeest across different degrees of human disturbance. Ecosphere 7:e01428.Google Scholar
Georgiadis, NJ, et al. (2007) Savanna herbivore dynamics in a livestock-dominated landscape. II: Ecological, conservation, and management implications of predator restoration. Biological Conservation 137:473483.Google Scholar
Ng’weno, CC, et al. (2017) Lions influence the decline and habitat shift of hartebeest in a semiarid savanna. Journal of Mammalogy 98:10781087.Google Scholar
Spinage, CA. (2013) Kobus ellipsiprymnus. In Kingdon, J; Hoffmann, M (eds) Mammals of Africa VI Pigs, Hippopotamuses, Chevrotain, Giraffes, Deer and Bovids. A & C Black, London, pp. 461468.Google Scholar
Mills, MGL. (2017) Kalahari Cheetahs: Adaptations to an Arid Region. Oxford University Press, Oxford.Google Scholar
Caro, TM. (1986) The functions of stotting: a review of the hypotheses. Animal Behaviour 34:649662.Google Scholar
Owen‐Smith, N; Goodall, V. (2014) Coping with savanna seasonality: comparative daily activity patterns of African ungulates as revealed by GPS telemetry. Journal of Zoology 293:181191.Google Scholar
Owen‐Smith, N; Traill, LW. (2017) Space use patterns of a large mammalian herbivore distinguished by activity state: fear versus food? Journal of Zoology 303:281290.CrossRefGoogle Scholar
Owen‐Smith, N. (2008) Changing vulnerability to predation related to season and sex in an African ungulate assemblage. Oikos 117:602610.Google Scholar
Owen-Smith, N. (2015) How diverse large herbivores coexist with multiple large carnivores in African savanna ecosystems: demographic, temporal and spatial influences on prey vulnerability. Oikos 124:14171426.Google Scholar
Kleynhans, EJ, et al. (2011) Resource partitioning along multiple niche dimensions in differently sized African savanna grazers. Oikos 120:591600.Google Scholar
Pansu, J, et al. (2019) Trophic ecology of large herbivores in a reassembling African ecosystem. Journal of Ecology 107:13551376.Google Scholar
Owen‐Smith, N. (2013) Contrasts in the large herbivore faunas of the southern continents in the late Pleistocene and the ecological implications for human origins. Journal of Biogeography 40:12151224.Google Scholar

Suggested Further Reading

Mills, MGL; Biggs, H. (1993) Prey apportionment and related ecological relationships between large carnivores in the Kruger National Park. Symposium of the Zoological Society of London 65:253268.Google Scholar
Radloff, FGT; du Toit, JT. (2004) Large predators and their prey in a southern African savanna: a predator’s size determines its prey size range. Journal of Animal Ecology 73:410423.CrossRefGoogle Scholar
Schaller, GB. (1972) The Serengeti Lion. A Study of Predator–Prey Relations. University of Chicago Press, Chicago.Google Scholar

References

DeVault, TL, et al. (2003) Scavenging by vertebrates: behavioral, ecological, and evolutionary perspectives on an important energy transfer pathway in terrestrial ecosystems. Oikos 102:225234.Google Scholar
Mills, MGL. (1990) Kalahari Hyaenas. Blackburn Press, Caldwell, NJ.Google Scholar
Hayward, MW; Kerley, GIH. (2005) Prey preferences of the lion (Panthera leo). Journal of Zoology 267:309322.Google Scholar
Owen-Smith, N; Mills, MGL. (2008) Shifting prey selection generates contrasting herbivore dynamics within a large‐mammal predator–prey web. Ecology 89:11201133.Google Scholar
Funston, PJ, et al. (1998) Hunting by male lions: ecological influences and socioecological implications. Animal Behaviour 56:13331345.Google Scholar
Scheel, D. (1993) Profitability, encounter rates, and prey choice of African lions. Behavioral Ecology 4:9097.Google Scholar
Joubert, D. (2006) Hunting behaviour of lions (Panthera leo) on elephants (Loxodonta africana) in the Chobe National Park, Botswana. African Journal of Ecology 44:279281.Google Scholar
Grange, S, et al. (2004) What limits the Serengeti zebra population? Oecologia 140:523532.Google Scholar
Owen-Smith, N; Mills, MGL. (2006) Manifold interactive influences on the population dynamics of a multispecies ungulate assemblage. Ecological Monographs 76:7392.Google Scholar
Kruuk, H. (1972) The Spotted Hyaena: A Study of Predation and Social Behavior. Chicago University Press, Chicago, IL.Google Scholar
De Boer, WF, et al. (2010) Spatial distribution of lion kills determined by the water dependency of prey species. Journal of Mammalogy 91:12801286.Google Scholar
Davidson, Z, et al. (2012) Environmental determinants of habitat and kill site selection in a large carnivore: scale matters. Journal of Mammalogy 93:677685.Google Scholar
Owen‐Smith, N. (2008) Changing vulnerability to predation related to season and sex in an African ungulate assemblage. Oikos 117:602610.Google Scholar
Ogutu, JO; Owen‐Smith, N. (2005) Oscillations in large mammal populations: are they related to predation or rainfall? African Journal of Ecology 43:332339.Google Scholar
Hayward, MW, et al. (2006) Prey preferences of the leopard (Panthera pardus). Journal of Zoology 270:298313.Google Scholar
Mills, MGL. (2017) Kalahari Cheetahs: Adaptations To An Arid Region. Oxford University Press, Oxford.CrossRefGoogle Scholar
Cozzi, G, et al. (2012) Fear of the dark or dinner by moonlight? Reduced temporal partitioning among Africa’s large carnivores. Ecology 93:25902599.Google Scholar
Hayward, MW, et al. (2006) Prey preferences of the cheetah (Acinonyx jubatus) (Felidae: Carnivora): morphological limitations or the need to capture rapidly consumable prey before kleptoparasites arrive? Journal of Zoology 270:615627.Google Scholar
Hayward, MW, et al. (2006) Prey preferences of the African wild dog Lycaon pictus (Canidae: Carnivora): ecological requirements for conservation. Journal of Mammalogy 87:11221131.Google Scholar
Pole, A, et al. (2003) African wild dogs test the ‘survival of the fittest’ paradigm. Proceedings of the Royal Society of London Series B: Biological Sciences 270:S57.Google Scholar
Hayward, MW. (2006) Prey preferences of the spotted hyaena (Crocuta crocuta) and degree of dietary overlap with the lion (Panthera leo). Journal of Zoology 270:606614.Google Scholar
Cooper, SM, et al. (1999) A seasonal feast: long‐term analysis of feeding behaviour in the spotted hyaena (Crocuta crocuta). African Journal of Ecology 37:149160.CrossRefGoogle Scholar
Henschel, JR; Skinner, JD. (1990) The diet of the spotted hyaenas Crocuta crocuta in Kruger National Park. African Journal of Ecology 28:6982.Google Scholar
Périquet, S, et al. (2015) Spotted hyaenas switch their foraging strategy as a response to changes in intraguild interactions with lions. Journal of Zoology 297:245254.CrossRefGoogle Scholar
Kruuk, H. (1976) Feeding and social behaviour of the striped hyaena (Hyaena vulgaris Desmarest). African Journal of Ecology 14:91111.CrossRefGoogle Scholar
Scheel, D; Packer, C. (1995) Variation in predation by lions: tracking a movable feast. In Sinclair, ARE, et al. (eds) Serengeti II: Dynamics, Management, and Conservation of an Ecosystem. University of Chicago Press, Chicago, IL, p. 299.Google Scholar
West, PM; Packer, C. (2013) Panthera leo. In Kingdon, J; Hoffmann, M (eds) Mammals of Africa V. Bloomsbury, London, pp. 149159.Google Scholar
Hunter, L, et al. (2013) Panthera pardus. In Kingdon, J; Hoffmann, M (eds) Mammals of Africa V. Bloomsbury, London, pp. 159167.Google Scholar
McNutt, JW; Woodroffe, R. (2013) Lycaon pictus. In Kingdon, J; Hoffmann, M (eds) Mammals of Africa V. Bloomsbury, London, pp. 5158.Google Scholar
Fitzgibbon, CD; Fanshawe, JH. (1989) The condition and age of Thomson’s gazelles killed by cheetahs and wild dogs. Journal of Zoology 218:99107.Google Scholar

Suggested Further Reading

Owen-Smith, N. (1988) Megaherbivores. The Influence of Very Large Body Size on Ecology. Cambridge University Press, Cambridge.Google Scholar
Owen-Smith, N. (2010) Dynamics of Large Herbivore Populations in Changing Environments. Wiley-Blackwell, Oxford.CrossRefGoogle Scholar

References

Mills, MGL, et al. (1995) The relationship bewteen rainfall, lion predation and population trends in African herbivores. Wildlife Research 22:7587.Google Scholar
Marshal, JP, et al. (2011) The role of El Niño–Southern Oscillation in the dynamics of a savanna large herbivore population. Oikos 120:11751182.Google Scholar
Dublin, HT; Ogutu, JO. (2015) Population regulation of African buffalo in the Mara–Serengeti ecosystem. Wildlife Research 42:382393.Google Scholar
Ottichilo, WK, et al. (2000) Population trends of large non‐migratory wild herbivores and livestock in the Masai Mara ecosystem, Kenya, between 1977 and 1997. African Journal of Ecology 38:202216.Google Scholar
Estes, RD, et al. (2006) Downward trends in Ngorongoro Crater ungulate populations 1986–2005: conservation concerns and the need for ecological research. Biological Conservation 131:106120.Google Scholar
Staver, AC, et al. (2019) Grazer movements exacerbate grass declines during drought in an African savanna. Journal of Ecology 107:14821491.Google Scholar
Ogutu, JO; Owen‐Smith, N. (2003) ENSO, rainfall and temperature influences on extreme population declines among African savanna ungulates. Ecology Letters 6:412419.Google Scholar
Owen-Smith, N; Mills, MGL. (2006) Manifold interactive influences on the population dynamics of a multispecies ungulate assemblage. Ecological Monographs 76:7392.CrossRefGoogle Scholar
Grange, S, et al. (2004) What limits the Serengeti zebra population? Oecologia 140:523532.Google Scholar
Georgiadis, N, et al. (2003) The influence of rainfall on zebra population dynamics: implications for management. Journal of Applied Ecology 40:125136.Google Scholar
Owen-Smith, N. (1990) Demography of a large herbivore, the greater kudu Tragelaphus strepsiceros, in relation to rainfall. The Journal of Animal Ecology 59:893913.Google Scholar
Mduma, SAR, et al. (1999) Food regulates the Serengeti wildebeest: a 40‐year record. Journal of Animal Ecology 68:11011122.Google Scholar
Owen-Smith, N (1981) The white rhinoceros overpopulation problem, and a proposed solution. In Jewell, PA et al. (eds) Problems in Management of Locally Abundant Wild Mammals. Academic Press, New York, NY, pp. 129150.Google Scholar
Moss, CJ. (2001) The demography of an African elephant (Loxodonta africana) population in Amboseli, Kenya. Journal of Zoology 255:145156.CrossRefGoogle Scholar
Corfield, TF. (1973) Elephant mortality in Tsavo National Park, Kenya. African Journal of Ecology 11:339368.Google Scholar
Smit, IPJ; Bond, WJ. (2020) Observations on the natural history of a savanna drought. African Journal of Range & Forage Science 37:119136.Google Scholar
Young, TP. (1994) Natural die‐offs of large mammals: implications for conservation. Conservation Biology 8:410418.Google Scholar
Spinage, CA; Matlhare, JM. (1992) Is the Kalahari cornucopia fact or fiction? A predictive model. Journal of Applied Ecology 29:605610.Google Scholar
Walker, BH, et al. (1987) To cull or not to cull: lessons from a southern African drought. Journal of Applied Ecology 24:381401.Google Scholar
Owen-Smith, N. (2019) Ramifying effects of the risk of predation on African multi-predator, multi-prey large-mammal assemblages and the conservation implications. Biological Conservation 232:5158.CrossRefGoogle Scholar
Owen-Smith, N. (2015) How diverse large herbivores coexist with multiple large carnivores in African savanna ecosystems: demographic, temporal and spatial influences on prey vulnerability. Oikos 124:14171426.Google Scholar
Yoganand, K; Owen‐Smith, N. (2014) Restricted habitat use by an African savanna herbivore through the seasonal cycle: key resources concept expanded. Ecography 37:969982.Google Scholar
Owen‐Smith, N; Traill, LW. (2017) Space use patterns of a large mammalian herbivore distinguished by activity state: fear versus food? Journal of Zoology 303:281290.Google Scholar
Scheel, D; Packer, C (1995) Variation in predation by lions: tracking a movable feast. In Sinclair, ARE et al. (eds) Serengeti II: Dynamics, Management, and Conservation of an Ecosystem. University of Chicago Press, Chicago, IL, p. 299.Google Scholar
Sinclair, ARE, et al. (2003) Patterns of predation in a diverse predator–prey system. Nature 425:288290.Google Scholar
Harrington, R, et al. (1999) Establishing the causes of the roan antelope decline in the Kruger National Park, South Africa. Biological Conservation 90:6978.CrossRefGoogle Scholar
Owen-Smith, N; Mills, MGL. (2008) Shifting prey selection generates contrasting herbivore dynamics within a large‐mammal predator–prey web. Ecology 89:11201133.Google Scholar
Fryxell, JM, et al. (1988) Why are migratory ungulates so abundant? The American Naturalist 131:781798.Google Scholar
Fryxell, JM; Sinclair, ARE. (1988) Seasonal migration by white‐eared kob in relation to resources. African Journal of Ecology 26:1731.Google Scholar
Cooper, SM, et al. (1999) A seasonal feast: long‐term analysis of feeding behaviour in the spotted hyaena (Crocuta crocuta). African Journal of Ecology 37:149160.Google Scholar
Harris, G, et al. (2009) Global decline in aggregated migrations of large terrestrial mammals. Endangered Species Research 7:5576.Google Scholar
Naidoo, R, et al. (2016) A newly discovered wildlife migration in Namibia and Botswana is the longest in Africa. Oryx 50:138146.Google Scholar
Morjan, MD, et al. (2018) Armed conflict and development in South Sudan threatens some of Africa’s longest and largest ungulate migrations. Biodiversity and Conservation 27:365380.Google Scholar
Naidoo, R, et al. (2012) Home on the range: factors explaining partial migration of African buffalo in a tropical environment. PLoS One 7:e36527.Google Scholar
Hillman, J. (1988) Home range and movement of the common eland (Taurotragus oryx Pallas 1766) in Kenya. African Journal of Ecology 26:135148.Google Scholar
Fryxell, JM, et al. (2005) Landscape scale, heterogeneity, and the viability of Serengeti grazers. Ecology Letters 8:328335.Google Scholar
Cagnacci, F, et al. (2016) How many routes lead to migration? Comparison of methods to assess and characterize migratory movements. Journal of Animal Ecology 85:5468.Google Scholar
Kreulen, D. (1975) Wildebeest habitat selection on the Serengeti plains, Tanzania, in relation to calcium and lactation: a preliminary report. African Journal of Ecology 13:297304.Google Scholar
Murray, MG. (1995) Specific nutrient requirements and migration of wildbeest. In Sinclair, ARE, et al. (eds) Serengeti II: Dynamics, Management, and Conservation of an Ecosystem. University of Chicago Press, Chicago, IL, p. 231.Google Scholar
Ogutu, J, et al. (2013) Changing wildlife populations in Nairobi National Park and adjoining Athi-Kaputiei Plains: collapse of the migratory wildebeest. The Open Conservation Biology Journal 7:1126.Google Scholar
Ogutu, JO, et al. (2012) Dynamics of ungulates in relation to climatic and land use changes in an insularized African savanna ecosystem. Biodiversity and Conservation 21:10331053.Google Scholar
Owen-Smith, RN. (1988) Megaherbivores: The Influence of Very Large Body Size on Ecology. Cambridge University Press, Cambridge.Google Scholar
Gaillard, J-M, et al. (2015) Does tooth wear influence ageing? A comparative study across large herbivores. Experimental Gerontology 71:4855.Google Scholar
Owen‐Smith, N; Ogutu, JO. (2013) Controls over reproductive phenology among ungulates: allometry and tropical–temperate contrasts. Ecography 36:256263.Google Scholar
Bigalke, RG. (1970) Observations on springbok populations. African Zoology 5:5970.Google Scholar
Mills, MGL. (2017) Kalahari Cheetahs: Adaptations to An Arid Region. Oxford University Press, Oxford.Google Scholar
Ryan, SJ, et al. (2007) Ecological cues, gestation length, and birth timing in African buffalo (Syncerus caffer). Behavioral Ecology 18:635644.Google Scholar
Smuts, GL. (1976) Population characteristics of Burchell’s zebra (Equus burchelli antiquorum. H. Smith, 1841) in the Kruger National Park. South African Journal of Wildlife Research 6:99112.Google Scholar
Whyte, IJ; Joubert, CSJ. (1988) Blue wildebeest population trends in the Kruger National Park and the effects of fencing. South African Journal of Wildlife Research 18:7887.Google Scholar
Sinclair, ARE, et al. (2000) What determines phenology and synchrony of ungulate breeding in Serengeti? Ecology 81:21002111.Google Scholar
Ogutu, JO, et al. (2015) How rainfall variation influences reproductive patterns of African savanna ungulates in an equatorial region where photoperiod variation is absent. PloS One 10:e0133744.Google Scholar
Estes, RD. (1976) The significance of breeding synchrony in the wildebeest. African Journal of Ecology 14:135152.Google Scholar
Augustine, DJ; McNaughton, SJ. (2004) Regulation of shrub dynamics by native browsing ungulates on East African rangeland. Journal of Applied Ecology 41:4558.Google Scholar
Owen-Smith, RN. (1983) Dispersal and the dynamics of large herbivores in enclosed areas: implications for management. In Owen-Smith, N (ed.) Management of Large Mammals in African Conservation Areas. Haum Educational Publishers, Pretoria, pp. 127143.Google Scholar
Owen‐Smith, N, et al. (2012) Shrinking sable antelope numbers in Kruger National Park: what is suppressing population recovery? Animal Conservation 15:195204.Google Scholar
Caughley, G, et al. (1980) Does dingo predation control the densities of kangaroos and emus? Wildlife Research 7:112.Google Scholar

Suggested Further Reading

Owen-Smith, N. (1988) Megaherbivores. The Influence of Very Large Body Size on Ecology. Cambridge University Press, Cambridge.Google Scholar

References

Briske, DD, et al. (2020) Strategies for global rangeland stewardship: assessment through the lens of the equilibrium–non‐equilibrium debate. Journal of Applied Ecology 57:10561067.Google Scholar
Bell, RHV. (1982) The effect of soil nutrient availability on community structure in African ecosystems. In Huntley, BJ; Walker, BH (eds) Ecology of Tropical Savannas. Springer, Berlin, pp. 193216.Google Scholar
Fritz, H; Duncan, P. (1994) On the carrying capacity for large ungulates of African savanna ecosystems. Proceedings of the Royal Society of London Series B: Biological Sciences 256:7782.Google Scholar
Rutherford, MC. (1980) Annual plant production–precipitation relations in arid and semi-arid regions. South African Journal of Science 76:5357.Google Scholar
Deshmukh, IK. (1984) A common relationship between precipitation and grassland peak biomass for east and southern Africa. African Journal of Ecology 22:181186.Google Scholar
Deshmukh, I. (1986) Primary production of a grassland in Nairobi National Park, Kenya. Journal of Applied Ecology 23:115123.Google Scholar
Cox, GW; Waithaka, JM. (1989) Estimating aboveground net production and grazing harvest by wildlife on tropical grassland range. Oikos 54:6066.Google Scholar
Strugnell, RG; Pigott, CD. (1978) Biomass, shoot-production and grazing of two grasslands in the Rwenzori National Park, Uganda. The Journal of Ecology 66:7396.Google Scholar
Anderson, TM, et al. (2007) Rainfall and soils modify plant community response to grazing in Serengeti National Park. Ecology 88:11911201.Google Scholar
Waldram, MS, et al. (2008) Ecological engineering by a mega-grazer: white rhino impacts on a South African savanna. Ecosystems 11:101112.Google Scholar
Arsenault, R; Owen-Smith, N. (2011) Competition and coexistence among short-grass grazers in the Hluhluwe-iMfolozi Park, South Africa. Canadian Journal of Zoology 89:900907.Google Scholar
Cromsigt, JPGM; te Beest, M. (2014) Restoration of a megaherbivore: landscape‐level impacts of white rhinoceros in Kruger National Park, South Africa. Journal of Ecology 102:566575.Google Scholar
Cromsigt, J, et al. (2017) The functional ecology of grazing lawns – how grazers, termites, people, and fire shape HiP’s savanna grassland mosaic. In Cromsigt, JPGM, et al. (eds) Conserving Africa’s Mega-diversity in the Anthropocene: The Hluhluwe-iMfolozi Park Story. Cambridge University Press, Cambridge, pp. 135160.Google Scholar
Olivier, RCD; Laurie, WA. (1974) Habitat utilization by hippopotamus in the Mara River. African Journal of Ecology 12:249271.Google Scholar
Lock, JM. (1972) The effects of hippopotamus grazing on grasslands. The Journal of Ecology 60:445467.Google Scholar
Kanga, EM, et al. (2013) Hippopotamus and livestock grazing: influences on riparian vegetation and facilitation of other herbivores in the Mara Region of Kenya. Landscape and Ecological Engineering 9:4758.Google Scholar
Verweij, R, et al. (2006) Grazing lawns contribute to the subsistence of mesoherbivores on dystrophic savannas. Oikos 114:108116.Google Scholar
McNaughton, SJ. (1983) Serengeti grassland ecology – the role of composite environmental-factors and contingency in community organization. Ecological Monographs 53:291320.Google Scholar
Yoganand, K; Owen‐Smith, N. (2014) Restricted habitat use by an African savanna herbivore through the seasonal cycle: key resources concept expanded. Ecography 37:969982.Google Scholar
McNaughton, SJ. (1985) Ecology of a grazing ecosystem: the Serengeti. Ecological Monographs 55:259294.Google Scholar
Anderson, TM, et al. (2010) Landscape‐scale analyses suggest both nutrient and antipredator advantages to Serengeti herbivore hotspots. Ecology 91:15191529.Google Scholar
Archibald, S, et al. (2017) Interactions between fire and ecosystem processes. In Cromsigt, JPGM, et al. (eds) Conserving Africa’s Mega-Diversity in the Anthropocene: The Hluhluwe-iMfolozi Park Story. Cambridge University Press, Cambridge, pp. 233261.Google Scholar
Sinclair, A, et al. (2008) Historical and future changes to the Serengeti ecosystem. In Sinclair, ARE, et al. (eds) Serengeti III: Human Impacts on Ecosystem Dynamics. University of Chicago Press, Chicago, pp. 746.Google Scholar
Eby, S, et al. (2015) Fire in the Serengeti ecosystem: history, drivers, and consequences. In Sinclair, ARE, et al. (eds) Serengeti IV: Sustaining Biodiversity in a Coupled Human–Natural System. University of Chicago Press, Chicago, pp. 73103.Google Scholar
Roques, KG, et al. (2001) Dynamics of shrub encroachment in an African savanna: relative influences of fire, herbivory, rainfall and density dependence. Journal of Applied Ecology 38:268280.Google Scholar
O’Connor, TG, et al. (2014) Bush encroachment in southern Africa: changes and causes. African Journal of Range & Forage Science 31:6788.Google Scholar
Moe, SR, et al. (2009) What controls woodland regeneration after elephants have killed the big trees? Journal of Applied Ecology 46:223230.Google Scholar
Voysey, MD, et al. (2021) The role of browsers in maintaining the openness of savanna grazing lawns. Journal of Ecology 109:913926.Google Scholar
Pellew, RA. (1984) The feeding ecology of a selective browser, the giraffe (Giraffa camelopardalis tippelskirchi). Journal of Zoology 202:5781.Google Scholar
Augustine, DJ; McNaughton, SJ. (2004) Regulation of shrub dynamics by native browsing ungulates on East African rangeland. Journal of Applied Ecology 41:4558.Google Scholar
Staver, AC; Bond, WJ. (2014) Is there a ‘browse trap’? Dynamics of herbivore impacts on trees and grasses in an African savanna. Journal of Ecology 102:595602.Google Scholar
Prins, HHT; van der Jeugd, HP. (1993) Herbivore population crashes and woodland structure in East Africa. Journal of Ecology 81:305314.Google Scholar
Cooper, SM; Owen-Smith, N. (1985) Condensed tannins deter feeding by browsing ruminants in a South African savanna. Oecologia 67:142146.Google Scholar
du Toit, JT; Owen-Smith, N. (1989) Body size, population metabolism, and habitat specialization among large African herbivores. The American Naturalist 133:736740.Google Scholar
Owen‐Smith, N, et al. (2019) Megabrowser impacts on woody vegetation in savannas. In Scogings, PF; Sankaran, M (eds) Savanna Woody Plants and Large Herbivores. Wiley, Oxford, pp. 585611.Google Scholar
Chafota, J; Owen-Smith, N. (2009) Episodic severe damage to canopy trees by elephants: interactions with fire, frost and rain. Journal of Tropical Ecology 25:341345.Google Scholar
Cromsigt, JPGM; Kuijper, DPJ. (2011) Revisiting the browsing lawn concept: evolutionary interactions or pruning herbivores? Perspectives in Plant Ecology, Evolution and Systematics 13:207215.Google Scholar
du Toit, JT; Olff, H. (2014) Generalities in grazing and browsing ecology: using across-guild comparisons to control contingencies. Oecologia 174:10751083.Google Scholar
Asner, GP; Levick, SR. (2012) Landscape‐scale effects of herbivores on treefall in African savannas. Ecology Letters 15:12111217.Google Scholar
Pellegrini, AFA, et al. (2017) Woody plant biomass and carbon exchange depend on elephant–fire interactions across a productivity gradient in African savanna. Journal of Ecology 105:111121.Google Scholar
Sinclair, ARE, et al. (2008) Historical and future changes to the Serengeti ecosystem. In Sinclair, ARE, et al. (eds) Serengeti III: Human Impacts on Ecosystem Dynamics. University of Chicago Press, Chicago, pp. 746.Google Scholar
Morrison, TA, et al. (2016) Elephant damage, not fire or rainfall, explains mortality of overstorey trees in Serengeti. Journal of Ecology 104:409418.Google Scholar
Trollope, WSW, et al. (1998) Long-term changes in the woody vegetation of the Kruger National Park, with special reference to the effects of elephants and fire. Koedoe 41:103112.Google Scholar
Laws, RM, et al. (1975) Elephants and Their Habitats. Clarendon Press, Oxford.Google Scholar
Dublin, HT, et al. (1990) Elephants and fire as causes of multiple stable states in the Serengeti–Mara woodlands. The Journal of Animal Ecology 59:11471164.Google Scholar
Dublin, HT. (1991) Dynamics of the Serengeti–Mara woodlands: an historical perspective. Forest & Conservation History 35:169178.Google Scholar
Agnew, ADQ. (1968) Observations on the changing vegetation of Tsavo National Park (East). African Journal of Ecology 6:7580.Google Scholar
Leuthold, W. (1977) Changes in tree populations of Tsavo East National Park, Kenya. African Journal of Ecology 15:6169.Google Scholar
Leuthold, W. (1996) Recovery of woody vegetation in Tsavo National Park, Kenya, 1970–94. African Journal of Ecology 34:101112.Google Scholar
Mosugelo, DK, et al. (2002) Vegetation changes during a 36‐year period in northern Chobe National Park, Botswana. African Journal of Ecology 40:232240.Google Scholar
Skarpe, C, et al. (2014) Plant–herbivore interactions. In Skarpe, C, et al. (eds) Elephants and Savanna Woodland Ecosystems. Wiley, Oxford, pp. 189206.Google Scholar
Teren, G, et al. (2018) Elephant‐mediated compositional changes in riparian canopy trees over more than two decades in northern Botswana. Journal of Vegetation Science 29:585595.Google Scholar
Makhabu, SW. (2005) Resource partitioning within a browsing guild in a key habitat, the Chobe Riverfront, Botswana. Journal of Tropical Ecology 21:641649.Google Scholar
Chamaillé‐Jammes, S, et al. (2007) Managing heterogeneity in elephant distribution: interactions between elephant population density and surface‐water availability. Journal of Applied Ecology 44:625633.Google Scholar
Sianga, K, et al. (2017) Spatial refuges buffer landscapes against homogenisation and degradation by large herbivore populations and facilitate vegetation heterogeneity. Koedoe 59:a1434.Google Scholar
Ben-Shahar, R. (1998) Changes in structure of savanna woodlands in northern Botswana following the impacts of elephants and fire. Plant Ecology 136:189194.Google Scholar
Rutina, LP, et al. (2005) Elephant Loxodonta africana driven woodland conversion to shrubland improves dry-season browse availability for impalas Aepyceros melampus. Wildlife Biology 11:207213.Google Scholar
Stevens, N, et al. (2017) Savanna woody encroachment is widespread across three continents. Global Change Biology 23:235244.Google Scholar
Gosling, CM, et al. (2012) Effects of erosion from mounds of different termite genera on distinct functional grassland types in an African savannah. Ecosystems 15:128139.Google Scholar
Goudie, AS. (1988) The geomorphological role of earthworms and termites in the tropics. In Viles, H (ed.) Biogeomorphology. Blackwell, Oxford, pp. 4382.Google Scholar
Davies, AB, et al. (2014) Spatial variability and abiotic determinants of termite mounds throughout a savanna catchment. Ecography 37:852862.Google Scholar
Mobæk, R, et al. (2005) Termitaria are focal feeding sites for large ungulates in Lake Mburo National Park, Uganda. Journal of Zoology 267:97102.Google Scholar
Davies, AB, et al. (2016) Termite mounds differ in their importance for herbivores across savanna types, seasons and spatial scales. Oikos 125:726734.Google Scholar
Holdo, RM; McDowell, LR. (2004) Termite mounds as nutrient-rich food patches for elephants. Biotropica 36:231239.Google Scholar
Loveridge, JP; Moe, SR. (2004) Termitaria as browsing hotspots for African megaherbivores in miombo woodland. Journal of Tropical Ecology 20:337343.Google Scholar
Ferrar, P. (1982) Termites of a South African savanna. IV. Subterranean populations, mass determinations and biomass estimations. Oecologia 52:147151.Google Scholar
Deshmukh, I. (1989) How important are termites in the production ecology of African savannas? Sociobiology 15:155168.Google Scholar
Borer, ET, et al. (2019) More salt, please: global patterns, responses and impacts of foliar sodium in grasslands. Ecology Letters 22:11361144.Google Scholar
Seagle, SW; McNaughton, SJ. (1992) Spatial variation in forage nutrient concentrations and the distribution of Serengeti grazing ungulates. Landscape Ecology 7:229241.Google Scholar
Griffith, DM, et al. (2017) Ungulate grazing drives higher ramet turnover in sodium‐adapted Serengeti grasses. Journal of Vegetation Science 28:815823.Google Scholar
Stock, WD, et al. (2010) Herbivore and nutrient control of lawn and bunch grass distributions in a southern African savanna. Plant Ecology 206:1527.Google Scholar
Weir, JS. (1972) Spatial distribution of elephants in an African National Park in relation to environmental sodium. Oikos 23:113.Google Scholar
Bond, WJ. (2005) Large parts of the world are brown or black: a different view on the ‘Green World’ hypothesis. Journal of Vegetation Science 16:261266.Google Scholar
Zimov, SA, et al. (2012) Mammoth steppe: a high-productivity phenomenon. Quaternary Science Reviews 57:2645.Google Scholar
Owen-Smith, N. (1987) Pleistocene extinctions: the pivotal role of megaherbivores. Paleobiology 13:351362.Google Scholar
Mueller-Dombois, D. (1972) Crown distortion and elephant distribution in the woody vegetations of Ruhuna National Park, Ceylon. Ecology 53:208226.Google Scholar
Karanth, KU; Sunquist, ME. (1992) Population structure, density and biomass of large herbivores in the tropical forests of Nagarahole, India. Journal of Tropical Ecology 8:2135.Google Scholar
Cardoso, AW, et al. (2020) The role of forest elephants in shaping tropical forest–savanna coexistence. Ecosystems 23:602616.Google Scholar

Suggested Further Reading

Bobe, R. (2011) Fossil mammals and paleoenvironments in the Omo–Turkana Basin. Evolutionary Anthropology 20:254263.Google Scholar
Elliot, MC; Berger, LR. (2018) A Handbook to the Cradle of Humankind. Reach Publishers, Wandsbeck.Google Scholar
Werdelin, L; Sanders, WJ (eds) (2010) The Cenozoic Mammals of Africa. University of California Press, Berkeley.Google Scholar

References

Abbate, E, et al. (2014) The East Africa Oligocene intertrappean beds: regional distribution, depositional environments and Afro/Arabian mammal dispersals. Journal of African Earth Sciences 99:463489.Google Scholar
Janis, CM. (1993) Tertiary mammal evolution in the context of changing climates, vegetation, and tectonic events. Annual Review of Ecology and Systematics 24:467500.Google Scholar
Sanders, WJ, et al. (2010) Proboscidea. In Werdelin, L; Sanders, WJ (eds) Cenozoic Mammals of Africa. University of California Press, Berkeley, pp. 161252.Google Scholar
Leakey, M, et al. (2011) Faunal change in the Turkana Basin during the late Oligocene and Miocene. Evolutionary Anthropology: Issues, News, and Reviews 20:238253.Google Scholar
Bibi, F. (2009) The fossil record and evolution of Bovidae. Palaeontologia Electronica 12:111.Google Scholar
Bibi, F. (2013) A multi-calibrated mitochondrial phylogeny of extant Bovidae (Artiodactyla, Ruminantia) and the importance of the fossil record to systematics. BMC Evolutionary Biology 13:166.Google Scholar
Gentry, AW (2010) Bovidae. In Werdelin, L; Sanders, WJ (eds) Cenozoic Mammals of Africa. University of California Press, Berkeley, pp. 741796.Google Scholar
Willows-Munro, S, et al. (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
Marcot, JD.(2007) Molecular phylogeny of terrestrial artiodactyls. In Prothero, DR; Foss, SE (eds) The Evolution of Artiodactyls. Johns Hopkins University Press, Baltimore, pp. 418.Google Scholar
Vrba, ES. (1997) New fossils of Alcelaphini and Caprinae (Bovidae: Mammalia) from Awash, Ethiopia, and phylogenetic analysis of Alcelaphini. Palaeontologica Africana 34:127198.Google Scholar
Leakey, MG, et al. (1996) Lothagam: a record of faunal change in the Late Miocene of East Africa. Journal of Vertebrate Paleontology 16:556570.Google Scholar
Bibi, F. (2011) Mio–Pliocene faunal exchanges and African biogeography: the record of fossil bovids. PLoS One 6:e16688.Google Scholar
Bishop, LC. (1999) Suid paleoecology and habitat preferences at African Pliocene and Pleistocene hominid localities. In Bromage, TG; Schrenki, F (eds) African Biogeography, Climate Change, and Human Evolution. Oxford University Press, Oxford, pp. 216225.Google Scholar
Gentry, AW. (2000) The ruminant radiation. In Vrba, ES; Schaller, GB (eds) Antelopes, Deer, and Relatives: Fossil Record, Behavioral Ecology, Systematics, and Conservation. Yale University Press, New Haven, pp. 1125.Google Scholar
Todd, NE. (2010) New phylogenetic analysis of the family Elephantidae based on cranial–dental morphology. The Anatomical Record: Advances in Integrative Anatomy and Evolutionary Biology 293:7490.Google Scholar
Cerling, TE, et al. (2010) Stable carbon and oxygen isotopes in East African mammals: modern and fossil. In Werdelin, L; Sanders, WJ (eds) Cenozoic Mammals of Africa. University of California Press, Berkeley, pp. 941952.Google Scholar
Cerling, TE, et al. (2015) Dietary changes of large herbivores in the Turkana Basin, Kenya from 4 to 1 Ma. Proceedings of the National Academy of Sciences 112:1146711472.Google Scholar
Uno, KT, et al. (2011) Late Miocene to Pliocene carbon isotope record of differential diet change among East African herbivores. Proceedings of the National Academy of Sciences 108:65096514.Google Scholar
Geraads, D; Bobe, R. (2020) Ruminants (Giraffidae and Bovidae) from Kanapoi. Journal of Human Evolution 140:102383.Google Scholar
Schubert, BW, et al. (2006) Microwear evidence for Plio–Pleistocene bovid diets from Makapansgat Limeworks Cave, South Africa. Palaeogeography, Palaeoclimatology, Palaeoecology 241:301319.Google Scholar
Zazzo, A, et al. (2000) Herbivore paleodiet and paleoenvironmental changes in Chad during the Pliocene using stable isotope ratios of tooth enamel carbonate. Paleobiology 26:294309.Google Scholar
Franz-Odendaal, TA, et al. (2002) New evidence for the lack of C4 grassland expansions during the early Pliocene at Langebaanweg, South Africa. Paleobiology 28:378388.Google Scholar
Stynder, DD. (2011) Fossil bovid diets indicate a scarcity of grass in the Langebaanweg E Quarry (South Africa) late Miocene/early Pliocene environment. Paleobiology 37:126139.Google Scholar
Lee-Thorp, JA, et al. (2007) Tracking changing environments using stable carbon isotopes in fossil tooth enamel: an example from the South African hominin sites. Journal of Human Evolution 53:595601.Google Scholar
Bobe, R. (2011) Fossil mammals and paleoenvironments in the Omo–Turkana Basin. Evolutionary Anthropology: Issues, News, and Reviews 20:254263.Google Scholar
Andrews, P; Humphrey, L. (1999) African Miocene environments and the transition to early hominines. In Bromage, TG; Schrenk, F (eds) African Biogeography, Climate Change, and Human Evolution. Oxford University Press, Oxford, pp. 282300.Google Scholar
Sanders, WJ. (2020) Proboscidea from Kanapoi, Kenya. Journal of Human Evolution 140:102547.Google Scholar
Su, DF; Harrison, T. (2007) The paleoecology of the Upper Laetolil Beds at Laetoli. In Bobe, R, et al. (eds) Hominin Environments in the East African Pliocene: An Assessment of the Faunal Evidence. Springer, Dordrecht, pp. 279313.Google Scholar
Bobe, R, et al. (2002) Faunal change, environmental variability and late Pliocene hominin evolution. Journal of Human Evolution 42:475497.Google Scholar
Geraads, D, et al. (2013) New ruminants (Mammalia) from the Pliocene of Kanapoi, Kenya, and a revision of previous collections, with a note on the Suidae. Journal of African Earth Sciences 85:5361.Google Scholar
Pickford, M. (2004) Southern Africa: a cradle of evolution. South African Journal of Science 100:205214.Google Scholar
Lorenzen, ED, et al. (2012) Comparative phylogeography of African savannah ungulates 1. Molecular Ecology 21:36563670.Google Scholar
Bibi, F, et al. (2018) Paleoecology of the Serengeti during the Oldowan-Acheulean transition at Olduvai Gorge, Tanzania: the mammal and fish evidence. Journal of Human Evolution 120:4875.Google Scholar
Patterson, DB, et al. (2017) Ecosystem evolution and hominin paleobiology at East Turkana, northern Kenya between 2.0 and 1.4 Ma. Palaeogeography, Palaeoclimatology, Palaeoecology 481:113.Google Scholar
O’Brien, K, et al. (2020) Ungulate turnover in the Koobi Fora Formation: Spatial and temporal variation in the Early Pleistocene. Journal of African Earth Sciences 161:103658.Google Scholar
Brain, CK. (1983) The Hunters or the Hunted? An Introduction to African Cave Taphonomy. University of Chicago Press, Chicago.Google Scholar
Elliott, MC; Berger, LR. (2018) A Handbook to the Cradle of Humankind. Reach Publishers, Wandsbeck.Google Scholar
McKee, JK. (1999) The autocatalytic nature of hominid evolution in African Plio–Pleistocene environments. In Bromage, TG; Schrenk, F (eds) African Biogeography, Climate Change, and Human Evolution. Oxford University Press, Oxford, pp. 5775.Google Scholar
Bobe, R. (2006) The evolution of arid ecosystems in eastern Africa. Journal of Arid Environments 66:564584.Google Scholar
Codron, D, et al. (2008) The evolution of ecological specialization in southern African ungulates: competition or physical environmental turnover? Oikos 117:344353.Google Scholar
Arctander, P, et al. (1999) Phylogeography of three closely related African bovids (tribe Alcelaphini). Molecular Biology and Evolution 16:17241739.Google Scholar
Faith, JT, et al. (2015) Paleoenvironmental context of the Middle Stone Age record from Karungu, Lake Victoria Basin, Kenya, and its implications for human and faunal dispersals in East Africa. Journal of Human Evolution 83:2845.Google Scholar
Klein, RG. (1984) The large mammals of southern Africa: late Pliocene to Recent. In Klein, RG (ed.) Southern African Prehistory and Paleoenvironments. Balkema, Rotterdam, pp. 107146.Google Scholar
Stynder, DD. (2009) The diets of ungulates from the hominid fossil-bearing site of Elandsfontein, Western Cape, South Africa. Quaternary Research 71:6270.CrossRefGoogle Scholar
Lehmann, SB, et al. (2016) Stable isotopic composition of fossil mammal teeth and environmental change in southwestern South Africa during the Pliocene and Pleistocene. Palaeogeography, Palaeoclimatology, Palaeoecology 457:396408.Google Scholar
Cerling, TE, et al. (1999) Browsing and grazing in elephants: the isotope record of modern and fossil proboscideans. Oecologia 120:364374.Google Scholar
Manthi, FK, et al. (2020) Late Middle Pleistocene elephants from Natodomeri, Kenya and the disappearance of Elephas (Proboscidea, Mammalia) in Africa. Journal of Mammalian Evolution 27:483-495.Google Scholar
Faith, JT. (2011) Ungulate community richness, grazer extinctions, and human subsistence behavior in southern Africa’s Cape Floral Region. Palaeogeography, Palaeoclimatology, Palaeoecology 306:219227.Google Scholar
Faith, JT, 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 361:8493.Google Scholar
Faith, JT. (2014) Late Pleistocene and Holocene mammal extinctions on continental Africa. Earth Science Reviews 128:105121.Google Scholar
Brink, JS. (2016) Faunal evidence for Mid- and Late Quaternary environmental change in southern Africa. In Knight, J; Grab, SW (eds) Quaternary Environmental Change in Southern Africa: Physical and Human Dimensions. Cambridge University Press, Cambridge, pp. 286307.Google Scholar
Faith, JT, et al. (2018) Plio–Pleistocene decline of African megaherbivores: No evidence for ancient hominin impacts. Science 362:938941.Google Scholar
Codron, D, et al. (2008) Functional differentiation of African grazing ruminants: an example of specialized adaptations to very small changes in diet. Biological Journal of the Linnean Society 94:755764.Google Scholar
Leakey, M. (1983) Africa’s Vanishing Art. The Rock Paintings of Tanzania. Doubleday, New York.Google Scholar
Hooijer, DA. (1969) Pleistocene East African rhinoceroses. In Leakey, LSB (ed.) Fossil Vertebrates of Africa. Vol. 1. Academic Press, New York, pp. 7198.Google Scholar
Delegorgue, A. (1990) Adulphe Delegorgue’s Travels in Southern Africa. University of Natal Press, Durban.Google Scholar
Werdelin, L; Peigné, S. (2010) Carnivora. In Werdelin, L; Sanders, WJ (eds) Cenozoic Mammals of Africa. University of California Press, Berkeley, pp. 603657.Google Scholar
Werdelin, L; Lewis, ME. (2005) Plio–Pleistocene Carnivora of eastern Africa: species richness and turnover patterns. Zoological Journal of the Linnean Society 144:121144.Google Scholar
Lewis, ME; Werdelin, L. (2007). Patterns of change in the Plio–Pleistocene carnivorans of eastern Africa. In Bobe, RA, et al. (eds) Hominin Environments in the East African Pliocene: An Assessment of the Faunal Evidence. Springer, Dordrecht, pp. 77105.Google Scholar
Turner, A; Antón, M. (2004) Evolving Eden: An Illustrated Guide to the Evolution of the African Large-mammal Fauna. Columbia University Press, New York.Google Scholar
Kuhn, BF, et al. (2016) The carnivore guild circa 1.98 million years: biodiversity and implications for the palaeoenvironment at Malapa, South Africa. Palaeobiodiversity and Palaeoenvironments 96:611616.Google Scholar
Prüfer, K, et al. (2014) The complete genome sequence of a Neanderthal from the Altai Mountains. Nature 505:4349.CrossRefGoogle ScholarPubMed
Groves, C; Grubb, P. (2011) Ungulate Taxonomy. Johns Hopkin University Press, Baltimore.Google Scholar
Rohland, N, et al. (2010) Genomic DNA sequences from mastodon and woolly mammoth reveal deep speciation of forest and savanna elephants. PLoS Biology 8:e1000564.Google Scholar
Webb, SD. (1978) A history of savanna vertebrates in the New World. Part II: South America and the Great Interchange. Annual Review of Ecology and Systematics 9:393426.Google Scholar
Sánchez, B, et al. (2004) Feeding ecology, dispersal, and extinction of South American Pleistocene gomphotheres (Gomphotheriidae, Proboscidea). Paleobiology 30:146161.2.0.CO;2>CrossRefGoogle Scholar
Owen‐Smith, N. (2013) Contrasts in the large herbivore faunas of the southern continents in the late Pleistocene and the ecological implications for human origins. Journal of Biogeography 40:12151224.Google Scholar
Rowan, J, et al. (2015) Taxonomy and paleoecology of fossil Bovidae (Mammalia, Artiodactyla) from the Kibish Formation, southern Ethiopia: implications for dietary change, biogeography, and the structure of the living bovid faunas of East Africa. Palaeogeography, Palaeoclimatology, Palaeoecology 420:210222.Google Scholar
Shrader, AM, et al. (2006) How a mega-grazer copes with the dry season: food and nutrient intake rates by white rhinoceros in the wild. Functional Ecology 20:376384.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
×