Skip to main content Accessibility help
×
Hostname: page-component-8448b6f56d-42gr6 Total loading time: 0 Render date: 2024-04-16T11:45:14.215Z Has data issue: false hasContentIssue false

Part I - The Human–Non-human Primate Interface

Published online by Cambridge University Press:  25 January 2019

Alison M. Behie
Affiliation:
Australian National University, Canberra
Julie A. Teichroeb
Affiliation:
University of Toronto, Scarborough
Nicholas Malone
Affiliation:
University of Auckland
Get access

Summary

Image of the first page of this content. For PDF version, please use the ‘Save PDF’ preceeding this image.'
Type
Chapter
Information
Publisher: Cambridge University Press
Print publication year: 2019

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

References

Adger, W. N., Hughes, T. P., Folke, C., Carpenter, S. R. & Rockström, J. (2005). Social-ecological resilience to coastal disasters. Science, 309(5737), 1036–9.Google Scholar
Campbell-Smith, G., Campbell-Smith, M., Singleton, I. & Linkie, M. (2011). Apes in space: saving an imperilled orangutan population in Sumatra. PLoS ONE, 6(2), e17210.CrossRefGoogle ScholarPubMed
Cheyne, S. M. (2011). Gibbon locomotion research in the field: problems, possibilities, and benefits for conservation. In D’Août, K. & Vereecke, E. E. (eds) Primate Locomotion. New York: Springer, pp. 201–13.Google Scholar
Cheyne, S. M., Thompson, C. J. & Chivers, D. J. (2013). Travel adaptations of Bornean agile gibbons Hylobates albibarbis (Primates: Hylobatidae) in a degraded secondary forest, Indonesia. Journal of Threatened Taxa, 5(5), 3963–8.CrossRefGoogle Scholar
Clutton-Brock, T. H. (1974). Primate social organization and ecology. Nature, 250, 539–42.Google Scholar
Cowlishaw, G. & Dunbar, R. (2000). Primate Conservation Biology. Chicago, IL: University of Chicago Press.CrossRefGoogle Scholar
Dore, K. M., Riley, E. P. & Fuentes, A. (eds) (2017). Ethnoprimatology: A Practical Guide to Research at the Human–Nonhuman Primate Interface. Cambridge: Cambridge University Press.Google Scholar
Eudey, A. A. (2008). The crab-eating macaque (Macaca fascicularis): widespread and rapidly declining. Primate Conservation, 23, 129–32.Google Scholar
Fuentes, A. (2012). Ethnoprimatology and the anthropology of the human–primate interface. Annual Review of Anthropology, 41, 101–17.CrossRefGoogle Scholar
Fuentes, A. & Baynes-Rock, M. (2017). Anthropogenic landscapes, human action and the process of co-construction with other species: making anthromes in the Anthropocene. Land, 6(1), 15.Google Scholar
Geissmann, T. & Nijman, V. (2006). Calling in wild silvery gibbons (Hylobates moloch) in Java (Indonesia): behaviour, phylogeny, and conservation. American Journal of Primatology, 68, 119.Google Scholar
Hockings, K. J. & McLennan, M. R. (2012). From forest to farm: systematic review of cultivar feeding by chimpanzees – management implications for wildlife in anthropogenic landscapes. PLoS ONE, 7(4), e33391.Google Scholar
Hockings, K. J., McLennan, M. R., Carvalho, S., et al. (2015). Apes in the Anthropocene: flexibility and survival. Trends in Ecology & Evolution, 30(4), 215–22.Google Scholar
IUCN. (2017). The IUCN Red List of Threatened Species. Version 2017.1. Available at: www.iucnredlist.org.Google Scholar
Kappeler, M. (1981). The Javan silvery gibbon (Hylobates lar moloch). PhD Thesis, Universitat Basel.Google Scholar
Kappeler, M. (1984a). The gibbon in Java. In Prueschoft, H., Chivers, D., Brockelman, W. Y. & Creel, N. (eds) The Lesser Apes: Evolution, Behaviour, and Biology. Edinburgh: Edinburgh University Press, pp. 1931.Google Scholar
Kappeler, M. (1984b). Vocal bouts and territorial maintenance in the Moloch gibbon. In Prueschoft, H. D., Chivers, D. J., Brockelman, W. Y. & Creel, N. (eds) The Lesser Apes: Evolution, Behaviour, and Biology. Edinburgh: Edinburgh University Press, pp. 376–89.Google Scholar
Kendal, J., Tehrani, J. J. & Odling-Smee, J. (2011). Human niche construction in interdisciplinary focus. Philosophical Transactions of the Royal Society B, 366, 785–92.CrossRefGoogle ScholarPubMed
Kim, S., Lappan, S. & Choe, J. C. (2011). Diet and ranging behaviour of the endangered Javan gibbon (Hylobates moloch) in a submontane tropical rainforest. American Journal of Primatology, 73, 270–80.Google Scholar
Laland, K. N., Odling-Smee, J. & Feldman, M. W. (2001). Niche construction, ecological inheritance, and cycles of contingency in evolution. In Oyama, S., Griffiths, P. E. & Gray, R. D. (eds) Cycles of Contingency: Developmental Systems and Evolution. Cambridge, MA: MIT Press, pp. 117–26.Google Scholar
Lappan, S. & Whittaker, D. (eds) (2009). The Gibbons: New Perspectives on Small Ape Socioecology and Population Biology. New York: Springer Science & Business Media.Google Scholar
Maldonado, A. & Peck, M. (2013). The role of primate conservation to fight the illegal trade in primates: the case of the owl monkeys in the Colombian–Peruvian Amazon. Folia Primatologica, 84, 299300.Google Scholar
Malone, N. (2007). The socioecology of the Critically Endangered Javan gibbon (Hylobates moloch): assessing the impact of anthropogenic disturbance on primate social systems. PhD Thesis, University of Oregon.Google Scholar
Malone, N. & Oktavinalis, H. (2006). The socioecology of the silvery gibbon (Hylobates moloch) in the Cagar Alam Leuweung Sancang (CALS), West Java, Indonesia. American Journal of Physical Anthropology, 129(S42), 124.Google Scholar
Malone, N., Fuentes, A., Purnama, A. R. & Wedana Adi Putra, I. M. (2004). Displaced Hylobatids: biological, cultural, and economic aspects of the primate trade in Java and Bali, Indonesia. Tropical Biodiversity, 40, 41–9.Google Scholar
Malone, N., Fuentes, A. & White, F. J. (2012a). Variation in the social systems of extant hominoids: comparative insight into the social behaviour of early hominins. International Journal of Primatology, 33(6), 1251–77.Google Scholar
Malone, N., Wade, A., Wedana Adi Putra, I. M., Reisland, M. & Selby, M. (2012b). Calibrating a conservation strategy for silvery gibbons (Hylobates moloch). American Journal of Physical Anthropology, 147(S54), 202.Google Scholar
Malone, N., Selby, M. & Longo, S. B. (2014a). Political and ecological dimensions of silvery gibbon conservation efforts: an endangered ape in (and on) the verge. International Journal of Sociology, 44(1), 3453.CrossRefGoogle Scholar
Malone, N., Wade, A. H., Fuentes, A., et al. (2014b). Ethnoprimatology: critical interdisciplinarity and multispecies approaches in anthropology. Critique of Anthropology, 38, 829.Google Scholar
Malone, N., Palmer, A. & Wade, A. H. (2017). Incorporating the ethnographic perspective: the value, process and responsibility of working with human participants. In Dore, K. M., Riley, E. P. & Fuentes, A. (eds) Ethnoprimatology: A Practical Guide to Research on the Human–Nonhuman Primate Interface. Cambridge: Cambridge University Press, pp. 176–89.Google Scholar
Marsh, L. K. (2003). The nature of fragmentation. In Marsh, L. K. (ed.) Primates in Fragments: Ecology and Conservation. New York: Kluwer Academic/Plenum Publishers, pp. 110.Google Scholar
Megantara, E. N. (1995). Distribusi, habitat dan populasi Owa (Hylobates moloch: Cabrera 1930) di Cagar Alam Leuweung Sancang, Jawa Barat. Bandung: Universitas Padjadjaran.Google Scholar
Meijerink, A. M. J. (1977). A hydrological reconnaissance survey of the Serayn river basin, Central Java. I. T. C. Journal, 4, 646–73.Google Scholar
National Research Council (1981). Techniques for the Study of Primate Population Ecology. Washington, DC: National Academy Press.Google Scholar
Nijman, V. (2004). Conservation of the Javan gibbon (Hylobates moloch): population estimates, local extinctions, and conservation priorities. The Raffles Bulletin of Zoology, 52(1), 271–80.Google Scholar
Nijman, V. (2005a). In Full Swing: An Assessment of the Trade in Gibbons and Orang-utans on Java and Bali, Indonesia. Petaling Jaya: TRAFFIC South-east Asia.Google Scholar
Nijman, V. (2005b). Hanging in the Balance: An Assessment of Trade in Orang-utans and Gibbons in Kalimantan, Indonesia. Petaling Jaya: TRAFFIC South-east Asia.Google Scholar
Nijman, V. (2010). An overview of international wildlife trade from Southeast Asia. Biodiversity and Conservation, 19, 1101–14.Google Scholar
Odling-Smee, J. (2007). Niche inheritance: a possible basis for classifying multiple inheritance systems in evolution. Biological Theory, 2(3), 276–89.Google Scholar
Odling-Smee, J., Laland, K. N. & Feldman, M. W. (2003). Niche Construction: The Neglected Process in Evolution. Princeton, NJ: Princeton University Press.Google Scholar
Pakpahan, H. (2003). Owa Jawa (Hylobates moloch: Cabrera 1930) menghilang dari Cagar Alam Leuweung Sancang. Habitat, 3(1), 24.Google Scholar
Piersma, T. & Drent, J. (2003). Phenotypic flexibility and the evolution of organismal design. Trends in Ecology and Evolution, 18, 228–33.Google Scholar
Ramos, A. R. (1990). Ethnology Brazilian style. Cultural Anthropology, 5(4), 452–72.Google Scholar
Reichard, U. H., Ganpanakngan, M. & Barelli, C. (2012). White-handed gibbons of Khao Yai: social flexibility, complex reproductive strategies, and a slow life history. In Kappeler, P. & Watts, D. (eds) Long-Term Field Studies of Primates. Berlin: Springer, pp. 237–58.Google Scholar
Reisland, M. A. (2013). Conservation in a sacred forest: an integrated approach to assessing the management of a community-based conservation site. PhD Thesis, University of Wisconsin–Madison.Google Scholar
Reisland, M. A. & Lambert, J. E. (2016). Sympatric apes in sacred forests: shared space and habitat use by humans and endangered Javan gibbons (Hylobates moloch). PLoS ONE, 11(1), e0146891.CrossRefGoogle ScholarPubMed
Riley, E. P. & Fuentes, A. (2010). Conserving social-ecological systems in Indonesia: human–nonhuman primate interconnections in Bali and Sulawesi. American Journal of Primatology, 73(1), 6274.Google Scholar
Smiet, A. C. (1990). Forest ecology on Java: conversion and usage in a historical perspective. Journal of Tropical Forest Science, 2, 286302.Google Scholar
Sterck, E. H. M., Watts, D. P. & van Schaik, C. P. (1997). The evolution of female social relationships in nonhuman primates. Behavioural Ecology and Sociobiology, 41, 291309.Google Scholar
Struhsaker, T. T. (1997). Ecology of an African Rainforest: Logging in Kibale and the Conflict Between Conservation and Exploitation. Gainesville, FL: University of Florida Press.Google Scholar
Struhsaker, T. T. (1999). Primate communities in Africa: the consequences of long-term evolution or the artifact of recent hunting? In Fleagle, J. G., Janson, C. & Reed, K. E. (eds) Primate Communities. Cambridge: Cambridge University Press, pp. 289–94.Google Scholar
Strum, S. C. (2010). The development of primate raiding: implications for management and conservation. International Journal of Primatology, 31, 133–56.Google Scholar
Supriatna, J. (2006). Conservation programs for the endangered Javan gibbon (Hylobates moloch). Primate Conservation, 21, 155–62.Google Scholar
Supriatna, J., Mootnick, A. & Andayani, N. (2010). Javan gibbon (Hylobates moloch): population and conservation. In Gursky-Doyen, S. & Supriatna, J. (eds) Indonesian Primates. New York: Springer, pp. 5772.CrossRefGoogle Scholar
Thierry, B. (2008). Primate socioecology, the lost dream of ecological determinism. Evolutionary Anthropology, 17, 93–6.CrossRefGoogle Scholar
USAID (2004) USAID Strategic Plan for Indonesia 2004–2008: Strengthening a Moderate, Stable, and Productive Indonesia. Washington DC: USAID.Google Scholar
van Schaik, C. P. (1989). The ecology of social relationships amongst female primates. In Standen, V. & Foley, R. A. (eds) Comparative Socioecology. Oxford: Blackwell Scientific Publications, pp. 195218.Google Scholar
van Schaik, C. P. & van Hooff, J. A. R. A. M. (1983). On the ultimate causes of primate social systems. Behaviour, 85, 91117.Google Scholar
Wedana Adi Putra, I. M. & Jeffery, S. (2016). Reinforcing the Javan silvery gibbon population in the Mount Tilu Nature Reserve, West Java, Indonesia. Presentation poster at the XXVIth Congress of the International Primatological Society Chicago, IL, USA, August 21–27, 2016.Google Scholar
Wedana Adi Putra, I. M., Atmoko, S. S., Oktavinalis, H. & Setiawan, A. (2010). Survey on the abundance and distribution of Javan silvery gibbons and endemic langur species outside of national park areas in West Java and Central Java. The Aspinall Foundation Indonesia Program.Google Scholar
Wessing, R. (1993). A change in the forest: myth and history in West Java. Journal of Southeast Asian Studies, 24(1), 117.Google Scholar
Wessing, R. (2006). Symbolic animals in the land between the waters: markers of place and transition. Asian Folklore Studies, 65(2), 205–39.Google Scholar
West-Eberhardt, M. J. (2003). Developmental Plasticity. Oxford: Oxford University Press.Google Scholar
Whitten, T., Soeriaatmaddja, R. E. & Afiff, S. A. (1996). The Ecology of Java and Bali. North Clarendon, VT: Periplus Editions.Google Scholar
Wrangham, R. W. (1979). On the evolution of ape social systems. Social Science Information, 18, 335–69.Google Scholar
Wrangham, R. W. (1980). An ecological model of female-bonded primate groups. Behaviour, 75, 262300.Google Scholar

References

Agostinelli, C. & Lund, U. (2013). R package ‘circular’: circular statistics. Available at: https://r-forge.r-project.org/projects/circular.Google Scholar
Alberts, S. C. & Altmann, J. (1995). Balancing costs and opportunities: dispersal in male baboons. The American Naturalist, 145, 279306.Google Scholar
Altmann, J., Altmann, S. A., Hausfater, G. & McCuskey, S. (1977). Life history of yellow baboons: physical development, reproductive parameters, and infant mortality. Primates, 18, 315–30.Google Scholar
Beissinger, S. R. (2002). Population viability analysis. In Beissinger, S. R. & McCullough, D. R. (eds) Population Viability Analysis: Past, Present, Future. Chicago, IL: University of Chicago Press, pp. 517.Google Scholar
Borries, C., Koenig, A. & Winkler, P. (2001). Variation of life history traits and mating patterns in female langur monkeys (Semnopithecus entellus). Behavioral Ecology and Sociobiology, 50, 391402.Google Scholar
Brook, B. W., Cannon, J. R., Lacy, R. C., Mirande, C. & Frankham, R. (1999). Comparison of the population viability analysis packages GAPPS, INMAT, RAMAS and VORTEX for the whooping crane (Grus americana). Animal Conservation, 2, 2331.Google Scholar
Brook, B. W., O’Grady, J. J., Chapman, A. P., et al. (2000). Predictive accuracy of population viability analysis in conservation biology. Nature, 404, 385–7.Google Scholar
Campbell, C. J., Fuentes, A., Mackinnon, K. C., Bearder, S. K. & Stumpf, R. M. (2011). Primates in Perspective, 2nd edn. New York: Oxford University Press.Google Scholar
Charpentier, M. J. E., Widdig, A. & Alberts, S. C. (2007). Inbreeding depression in non-human primates: a historical review of methods used and empirical data. American Journal of Primatology, 69, 1370–86.Google Scholar
Colmenares, F. & Gomendio, M. (1988). Changes in female reproductive condition following male takeovers in a colony of hamadryas and hybrid baboons. Folia Primatologica, 50, 157–74.Google Scholar
Crockett, C. M. (1984). Emigration by female red howler monkeys and the case for female competition. In Small, M. F. (ed.) Female Primates: Studies by Women Primatologists. New York: Alan R. Liss., Inc., pp. 159–73.Google Scholar
Deschner, T. & Boesch, C. (2007). Can the patterns of sexual swelling cycles in female Taï chimpanzees be explained by the cost-of-sexual-attraction hypothesis? International Journal of Primatology, 28, 389406.Google Scholar
Dinesen, L., Lehmberg, T., Rahner, M. C. & Fjeldsa, J. (2001). Conservation priorities for the forests of the Udzungwa Mountains, Tanzania, based on primates, duikers and birds. Biological Conservation, 99, 223–36.CrossRefGoogle Scholar
Dittus, W. P. J. (1988). Group fission among wild toque macaques as a consequence of female resource competition and environmental stress. Animal Behaviour, 36, 1626–45.CrossRefGoogle Scholar
Ehardt, C. L. (2001). The endemic primates of the Udzungwa Mountains, Tanzania. African Primates, 4, 1526.Google Scholar
Ehardt, C. L. & Butynski, T. M. (2006a). The recently described highland mangabey Lophocebus kipunji (Cercopithecoidea, Cercopithicinae): current knowledge and conservation assessment. Primate Conservation, 21, 8188.Google Scholar
Ehardt, C. L. & Butynski, T. M. (2006b). Sanje mangabey, Cercocebus sanjei (Mittermeier, 1986). In Mittermeier, R. A., Valladares-Padua, C., Rylands, A. B., et al. (eds) Primates in Peril: The World’s 25 Most Endangered Primates 2004–2006. Washington, DC: IUCN Species Survival Commission.Google Scholar
Ehardt, C. L., Struhsaker, T. T. & Butynski, T. M. (1999). Conservation of the endangered primates of the Udzungwa Mountains, Tanzania: surveys, habitat assessment, and long-term monitoring. Unpublished report to Margot Marsh Biodiversity Foundation, Conservation International.Google Scholar
Ehardt, C. L., Butynski, T. M. & Struhsaker, T. T. (2001). Conservation of the endangered endemic primates of the Udzungwa Mountains, Tanzania, phase II: population survey and census, demography, and socioecology. Unpublished report to Margot Marsh Biodiversity Foundation, Conservation International.Google Scholar
Ehardt, C. L., Jones, T. P. & Butynski, T. M. (2005). Protective status, ecology and strategies for improving conservation of Cercocebus sanjei in the Udzungwa Mountains, Tanzania. International Journal of Primatology, 26, 557–83.Google Scholar
Ehardt, C. L., Butynski, T. M. & Struhsaker, T. T. (2008). Cercocebus sanjei. IUCN 2008. IUCN Red List of Threatened Species. Version 2012.2. Available at: www.iucnredlist.org (accessed 27 January 2013).Google Scholar
Ellison, P. T. (2003). Energetics and reproductive effort. American Journal of Human Biology, 15, 342–51.Google Scholar
Fernández, D. (2017). Consequences of a male takeover on mating skew in wild Sanje mangabeys. American Journal of Primatology, 79, e22532.Google Scholar
Fernández, D., Doran-Sheehy, D., Borries, C. & Brown, J. L. (2014). Reproductive characteristics of wild Sanje mangabeys (Cercocebus sanjei). American Journal of Primatology, 76, 1163–74.CrossRefGoogle ScholarPubMed
Fernández, D., Doran-Sheehy, D., Borries, C. & Ehardt, C. L. (2017). Exaggerated sexual swellings and the probability of conception in wild Sanje mangabeys (Cercocebus sanjei). International Journal of Primatology, 38, 513–32.Google Scholar
Gilmore, D. & Cook, B. (1981). Environmental Factors in Mammal Reproduction. London: Palgrave Macmillan.Google Scholar
Hagell, S., Whipple, A. V. & Chambers, C. L. (2013). Population genetic patterns among social groups of the endangered Central American spider monkey (Ateles geoffroyi) in a human-dominated landscape. Ecology and Evolution, 3, 1388–99.Google Scholar
Harrison, P. (2006). Socio-Economic Study of Forest-Adjacent Communities from Nyanganje Forest to Udzungwa Scarp: A Potential Wildlife Corridor. Dar es Salaam: World Wide Fund for Nature. Available at: www.cepf.net/Documents/wwf_5fudzungwas.pdf (accessed 2 September 2017).Google Scholar
Homewood, K. (1975). Monkey on a riverbank. Natural History, 84, 6873.Google Scholar
Homewood, K. & Rodgers, W. A. (1981). A previously undescribed mangabey from Southern Tanzania. International Journal of Primatology, 2, 4755.Google Scholar
Isbell, L. A. & Van Vuren, D. (1996). Differential costs of locational and social dispersal and their consequences for female group-living primates. Behaviour, 133, 136.CrossRefGoogle Scholar
Jack, K. & Isbell, L. (2009). Dispersal in primates: advancing an individualized approach. Behaviour, 146, 429–36.Google Scholar
Jones, T., Ehardt, C. L., Butynski, T. M., et al. (2005). The highland mangabey Lophocebus kipunji: a new species of African monkey. Science, 308, 1161–4.Google Scholar
Jones, T., Mselewa, L. S. & Mtui, A. (2006). Sanje mangabey Cercocebus sanjei kills an African crowned eagle Stephanoaetus coronatus. Folia Primatologica, 77, 359–63.Google Scholar
Kikula, I. S., Mnzava, E. Z. & Mung’ong’o, C. G. (2003). Shortcomings of Linkages Between Environmental Conservation Initiatives and Poverty Alleviation in Tanzania. Dar es Salaam: Research on Poverty Alleviation. Available at: www.repoa.or.tz/documents/03.2_-_Kikula_Mnzava_Mungongo_.pdf (accessed 2 September 2017).Google Scholar
Kinnaird, M. F. (1990). Pregnancy, gestation and parturition in free-ranging Tana River crested mangabeys (Cercocebus galeritus galeritus). American Journal of Primatology, 22, 285–9.Google Scholar
Kinnaird, M. F. & O’Brien, T. G. (1991). Viable populations for an endangered forest primate, the Tana River crested mangabey (Cercocebus galeritus galeritus). Conservation Biology, 5, 203–13.Google Scholar
Lacy, R. C. (1993). VORTEX: a computer simulation model for population viability analysis. Wildlife Research, 20, 4565.Google Scholar
Lacy, R. C. (1997). Importance of genetic variation to the viability of mammalian populations. Journal of Mammalogy, 78, 320–35.Google Scholar
Lovett, J. C. (1993). Eastern Arc moist forest flora. In Lovett, J. C. & Wasser, S. K. (eds) Biogeography and Ecology of the Rain Forest of Eastern Africa. New York: Cambridge University Press, pp. 3355.CrossRefGoogle Scholar
Martin, R. D. (2007). The evolution of human reproduction: a primatological perspective. Yearbook of Physical Anthropology, 134, 5984.Google Scholar
Mbora, D. N. M. & McPeek, M. A. (2015). How monkeys see a forest: genetic variation and population genetic structure of two forest primates. Conservation Genetics, 16, 559–69.CrossRefGoogle Scholar
McCabe, G. M. & Emery Thompson, M. (2013). Reproductive seasonality in wild Sanje mangabeys (Cercocebus sanjei), Tanzania: relationship between the capital breeding strategy and infant survival. Behaviour, 150, 1399–429.Google Scholar
McCabe, G. M. & Fernández, D. (in press). Seasonal patterns of infant mortality in wild sanje mangabeys, Cercocebus sanjei. In Kalbitzer, U. & Jack, K. M. (eds) Primate Life Histories, Sex Roles, and Adaptability: Essays in Honour of Linda M. Fedigan. New York: Springer.Google Scholar
McCabe, G., Fernández, D. & Ehardt, C. (2013). Ecology of reproduction in Sanje mangabeys (Cercocebus sanjei): dietary strategies and energy balance during the high fruit period. American Journal of Primatology, 75, 1196–208.Google Scholar
Medley, K. E. (1993). Primate conservation along the Tana River, Kenya: an examination of the forest habitat. Conservation Biology, 7, 109–21.CrossRefGoogle Scholar
Melnick, D. J. & Hoelzer, G. A. (1996). The population genetic consequences of macaque social organization and behaviour. In Fa, J. E. & Lindburg, D. G. (eds) Evolution and Ecology of Macaque Societies. Cambridge: Cambridge University Press, pp. 413–43.Google Scholar
Melnick, D. J. & Kidd, K. K. (1983). The genetic consequences of social group fission in a wild population of rhesus monkeys (Macaca mulatta). Behavioural Ecology and Sociobiology, 3, 229–36.Google Scholar
Moore, J. (1984). Female transfer in primates. International Journal of Primatology, 5, 537–89.Google Scholar
Palombit, R. A. (2015). Infanticide as sexual conflict: coevolution of male strategies and female counterstrategies. Cold Spring Harbor Perspectives in Biology, 7, a017640.CrossRefGoogle ScholarPubMed
Perry, S. E., Manson, J. H., Muniz, L., Gros-Louis, J. & Vigilant, L. (2008). Kin-biased social behaviour in wild adult female white-faced capuchins, Cebus capucinus. Animal Behaviour, 76, 187–99.Google Scholar
Pope, T. R. (1992). The influence of dispersal patterns and mating systems on genetic differentiation within and between populations of red howler monkeys (Alouatta seniculus). Evolution, 46, 1112–28.Google Scholar
Pusey, A. E. (1979). Intercommunity transfer of chimpanzees in Gombe National Park. In Hamburg, D. A. & McCown, E. R. (eds) The Great Apes. Menlo Park, CA: Benjamin/Cummings, pp. 465579.Google Scholar
Pusey, A., Williams, J. & Goodall, J. (1997). The influence of dominance rank on the reproductive success of female chimpanzees. Science, 277, 828–31.CrossRefGoogle ScholarPubMed
R Core Team (2017). R: A Language and Environment for Statistical Computing. Vienna: R Foundation for Statistical Computing.Google Scholar
Range, F., Foerderer, T., Storrer-Meystre, Y., Benetton, C. & Fruteau, C. (2007). The structure of social relationships among sooty mangabeys in Taï. In McGraw, W. S., Zuberbuehler, W. S. & Noe, R. (eds) Monkeys of the Taï Forest: An African Primate Community. New York: Cambridge University Press, pp. 109–30.Google Scholar
Rao, J. S. (1972). Some variants of chi-square for testing uniformity on the circle. Zeitschrift fr Wahrscheinlichkeitstheorie und Verwandte Gebiete, 22, 3344.Google Scholar
Robbins, M. M., Gray, M., Kagoda, E. & Robbins, A. M. (2009). Population dynamics of the Bwindi mountain gorillas. Biological Conservation, 142, 2886–95.Google Scholar
Rovero, F., Mtui, A., Kitegile, A., Nielsen, M. & Jones, T. (2010). Udzungwa Scarp Forest Reserve in crisis. Available at: www.cepf.net/Documents/UdzungwaSFR_Report_FINAL_High_Res.pdf (accessed 2 September 2017).Google Scholar
Rovero, F., Mtui, A. S., Kitegile, A. S. & Nielsen, M. R. (2012). Hunting or habitat degradation? Decline of primate populations in Udzungwa Mountains, Tanzania: an analysis of threats. Biological Conservation, 146, 8996.Google Scholar
Rovero, F., Mtui, A., Kitegile, A., et al. (2015). Primates decline rapidly in unprotected forests: evidence from a monitoring program with data constraints. PLoS ONE, 10, e0118330.Google Scholar
Sadleir, R. M. F. S. (1969). The Ecology of Reproduction in Wild and Domestic Animals. London: Methuen.Google Scholar
Scheijen, C. (2014). Human–elephant conflict along the eastern boundary of the Udzungwa Mountains National Park, Tanzania. Available at: www.samhao.nl/webopac/MetaDataEditDownload.csp?file=2:141504:1 (accessed 2 September 2017).Google Scholar
Sjögren-Gulve, P. & Ebenhard, T. (2000). The Use of Population Viability Analyses in Conservation Planning. Copenhagen: Munksgaard International Publishers.Google Scholar
Skorupa, J. P. (1986). Responses of rainforest primates to selective logging in Kibale Forest, Uganda. In Benirschke, K. (ed.) Primates: The Road to Self-Sustaining Populations. New York: Springer Verlag, pp. 5770.Google Scholar
Sterck, E. H. M. (1997). Determinants of female dispersal in Thomas langurs. American Journal of Primatology, 42, 179–98.Google Scholar
Strier, K. B. & Ives, A. R. (2012). Demography in the recovery of an endangered primate population. PLoS ONE, 7, e44407.Google Scholar
Struhsaker, T. T. (1999). Primate communities in Africa: the consequence of longterm evolution or the artifact of recent hunting? In Fleagle, J. G., Janson, C. & Reed, K. E. (eds) Primate Communities. Cambridge: Cambridge University Press, pp. 289–94.Google Scholar
Therneau, T. (2015). A package for survival analysis in S. Available at: https://cran.r-project.org/web/packages/survival/index.html.Google Scholar
Watts, D. P. (1990). Ecology of gorillas and its relation to female transfer in mountain gorillas. International Journal of Primatology, 11, 2145.Google Scholar
Whitlock, M. C. & McCauley, D. E. (1990). Some population genetic consequences of colony formation and extinction: genetic correlations within founding groups. Evolution, 7, 1717–24.Google Scholar
Wieczkowski, J. & Butynski, T. M. (2013). Cercocebus galeritus. In Kingdon, J., Happold, D. & Butynski, T. M. (eds) The Mammals of Africa. Primates, Vol. II. London: Bloomsbury Publishing, pp. 167–70.Google Scholar
Wright, S. (1943). Isolation by distance. Genetics, 28, 114–38.Google Scholar
Zinner, D. & Deschner, T. (2000). Sexual swellings in female hamadryas baboons after male take-overs: ‘deceptive’ swellings as a possible female counter-strategy against infanticide. American Journal of Primatology, 52, 157–68.Google Scholar

References

Bergl, R. A., Warren, Y., Nicholas, A., et al. (2012). Remote sensing analysis reveals habitat, dispersal corridors and expanded distribution for the Critically Endangered Cross River gorilla Gorilla gorilla diehliOryx46(2), 278289.Google Scholar
Bergl, R. A., Dunn, A., Fowler, A., et al. (2016). Gorilla gorilla spp. diehli. The IUCN Red List of Threatened Species 2016. Available at: www.iucnredlist.org/details/39998/0 (accessed 28 December 2016).Google Scholar
Brownlow, A. R., Plumptre, A. J., Reynolds, V. & Ward, R. (2001). Sources of variation in the nesting behavior of chimpanzees (Pan troglodytes schweinfurthii) in the Budongo Forest, Uganda. American Journal of Primatology, 55, 4955.Google Scholar
Ceballos, G., Ehrlich, P. R., Barnosky, A. D., et al. (2015). Accelerated modern human-induced species losses: entering the sixth mass extinction. Scientific Advances, 1, 15.Google Scholar
De Vere, R. A., Warren, Y., Nicholas, A., MacKenzie, M. E. & Higham, J. P. (2011). Nest site ecology of the Cross River gorilla at the Kagwene Gorilla Sanctuary, Cameroon, with special reference to anthropogenic influence. American Journal of Primatology, 73, 253–61.Google Scholar
Doran-Sheehy, D. M., Greer, D., Mongo, P. & Schwindt, D. (2004). Impact of ecological and social factors on ranging in western gorillas. American Journal of Primatology, 58, 91116.Google Scholar
Dore, K. M., Riley, E. P. & Fuentes, A. (eds) (2017). Ethnoprimatology: A Practical Guide to Research at the Human–Nonhuman Primate Interface. Cambridge: Cambridge University Press.Google Scholar
Dunn, A., Bergl, R., Byler, D., et al. (2014). Revised Regional Action Plan for the Conservation of the Cross River Gorilla (Gorilla gorilla diehli): 2014–2019. New York: IUCN/SSC Primate Specialist Group and Wildlife Conservation Society.Google Scholar
Dutton, P. E. (2012). Chimpanzee (Pan troglodytes ellioti) ecology in a Nigerian montane forest. PhD Dissertation, University of Canterbury.Google Scholar
Ellwanger, A. L., Riley, E. P., Niu, K. & Tan, C. L. (2017). Using a mixed-methods approach to elucidate the conservation implications of the human–primate interface in Fanjingshan National Nature Reserve, China. In Dore, K. M., Riley, E. P. & Fuentes, A. (eds) Ethnoprimatology: A Practical Guide to Research at the Human–Nonhuman Primate Interface. Cambridge: Cambridge University Press, pp. 257–70.Google Scholar
Estrada, A., Garber, P. A., Rylands, A. B., et al. (2017). Impending extinction crisis of the world’s primates: why primates matter. Science Advances, 3, 116.CrossRefGoogle ScholarPubMed
Etiendem, D. N., Funwi-Gabga, N., Tagg, N., Hens, L. & Indah, E. K. (2013). The Cross River gorillas (Gorilla gorilla diehli) at Mawambi Hills, south-west Cameroon: habitat suitability and vulnerability to anthropogenic disturbance. Folia Primatologica, 84, 1831.Google Scholar
Fedigan, L. M. (2010). Ethical issues faced by field primatologists: asking the relevant questions. American Journal of Primatology, 72, 754–71.Google Scholar
Frid, A. & Dill, L. (2002). Human-caused disturbance stimuli as a form of predation risk. Conservation Ecology, 6, 1126.Google Scholar
Fruth, B. & Hohmann, G. (1996). Nest building behavior in the great apes: the great leap forward? In McGrew, W. C., Merchant, L. F. & Nishida, T. (eds) Great Ape Societies. Cambridge: Cambridge University Press, pp. 225–40.Google Scholar
Fuentes, A. (2007). Monkey and human interconnections: the wild, the captive, and the in-between. In Cassidy, R. & Mullin, M. (eds) Where the Wild Things Are Now: Domestication Reconsidered. Oxford: Berg, pp. 123–45.Google Scholar
Fuentes, A. (2012). Ethnoprimatology and the anthropology of the human–primate interface. Annual Review of Anthropology, 41, 101–17.Google Scholar
Fuentes, A. & Wolfe, L. D. (2002). Primates Face to Face: The Conservation Implications of Human–Nonhuman Primate Interconnections. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Furuichi, T. & Hashimoto, C. (2004). Botanical and topographical factors influencing nesting-site selection by chimpanzees in Kalinzu Forest, Uganda. International Journal of Primatology, 25, 755–65.Google Scholar
Gagneux, P., Gonder, M. K., Goldberg, T. A. & Morin, P. A. (2001). Gene flow in wild chimpanzees: what genetic data tell us about chimpanzee movements over space and time. Philosophical Transactions of the Royal Society B, 356, 889–97.Google Scholar
Ganas, J. & Robbins, M. M. (2005). Ranging behaviour of the mountain gorillas (Gorilla beringei beringei) in Bwindi Impenetrable National Park, Uganda: a test of the ecological constraints model. Behavioral Ecology and Sociobiology, 58, 277–88.Google Scholar
Gezon, L. L. (2006). Global Visions, Local Landscapes: A Political Ecology of Conservation, Conflict, and Control in Northern Madagascar. Lanham, MD: Altamira Press.Google Scholar
Imong, I., Robbins, M. M., Mundry, R., Bergl, R. & Kühl, H. S. (2014). Distinguishing ecological constraints from human activity in species range fragmentation: the case of Cross River gorillas. Animal Conservation, 17, 323–31.Google Scholar
Jacobson, S. K. (2010). Effective primate conservation education: gaps and opportunities. American Journal of Primatology, 72, 414–19.Google Scholar
Jameson, C. (2012). Gorilla Guardian update: expansion of the community-based monitoring network. Gorilla Journal, 45, 1315.Google Scholar
Koops, K., McGrew, W. C., de Vries, H. & Matsuzawa, T. (2012). Nest-building by chimpanzees (Pan troglodytes verus) at Seringbara, Nimba Mountains: antipredation, thermoregulation, and antivector hypotheses. International Journal of Primatology, 33, 356–80.Google Scholar
Kühl, H., Maisels, F., Ancrenaz, M. & Williamson, E. A. (2008). Best Practice Guidelines for Surveys and Monitoring of Great Ape Populations. Gland: IUCN SSC Primate Specialist Group.Google Scholar
Malone, N., Palmer, A. & Wade, A. H. (2017). Incorporating the ethnographic perspective: the value, process, and responsibility of working with human participants. In Dore, K. M., Riley, E. P. & Fuentes, A. (eds) Ethnoprimatology: A Practical Guide to Research at the Human–Nonhuman Primate Interface. Cambridge: Cambridge University Press, pp. 176–89.Google Scholar
Mboh, H. & Warren, Y. (2007). Large mammal survey of the proposed Takamanda National Park. Unpublished report. Wildlife Conservation Society.Google Scholar
McFarland, K. L. (2007). Ecology of Cross River gorillas (Gorilla gorilla diehli) on Afi Mountain, Cross River State, Nigeria. PhD Dissertation, City University of New York.Google Scholar
McLennan, M. R., Spagnoletti, N. & Hockings, K. J. (2017). The implications of primate behavioral flexibility for sustainable human–primate coexistence in anthropogenic habitats. International Journal of Primatology, 38, 105–21.Google Scholar
Morgan, B., Adeleke, A., Bassey, T., et al. (2011). Regional Action Plan for the Conservation of the Nigeria–Cameroon Chimpanzee (Pan troglodytes ellioti). San Diego, CA: IUCN/SSC Primate Specialist Group and Sociological Society of San Diego.Google Scholar
Oates, J. F., Doumbe, O., Dunn, A., et al. (2016). Pan troglodytes ssp. ellioti. The IUCN Red List of Threatened Species 2016. Available at: www.iucnredlist.org/details/40014/0 (accessed 28 December 2016).Google Scholar
Oishi, T. (2013). Human–gorilla and gorilla–human: dynamics of human–animal boundaries and interethnic relationships in the central African rainforest. Revue de Primatology, 5, 123.Google Scholar
QGIS. (2014). QGIS 2.12.3: a free and open source geographic information system. Available at: www.qgis.org/en/site.Google Scholar
Radhakrishna, S. (2017). Culture, conflict, and conservation: living with nonhuman primates in Northeastern India. In Dore, K. M., Riley, E. P. & Fuentes, A. (eds) Ethnoprimatology: A Practical Guide to Research at the Human–Nonhuman Primate Interface. Cambridge: Cambridge University Press, pp. 271–83.Google Scholar
Remis, M. J. (1997). Ranging and grouping patterns of a western lowland gorilla group at Bai Hokou, Central African Republic. American Journal of Primatology, 43, 111–33.Google Scholar
Riley, E. P. & Ellwanger, A. L. (2013). Methods in ethnoprimatology: exploring the human–nonhuman primate interface. In Sterling, E., Bynum, N. & Blair, M. (eds) Primate Ecology and Conservation: A Handbook of Techniques. Oxford: Oxford University Press, pp. 128–50.Google Scholar
Rothman, J. M., Pell, A. N., Dierenfeld, E. S. & McCann, C. M. (2006). Plant choice in the construction of night nests by gorillas in the Bwindi Impenetrable National Park, Uganda. American Journal of Primatology, 68, 361–8.Google Scholar
Sawyer, S. C. (2012). The ecology and conservation of the Critically Endangered Cross River gorilla in Cameroon. PhD Dissertation, University of California.Google Scholar
Setchell, J. M. (2013). Editorial: the top 10 questions in primatology. International Journal of Primatology, 34, 647–61.Google Scholar
Setchell, J. M., Fairet, E., Shutt, K., Waters, S. & Bell, S. (2016). Biosocial conservation: integrating biological and ethnographic methods to study human–primate interactions. International Journal of Primatology, 38, 126.Google Scholar
Sponsel, L. E. (1997). The human niche in Amazonia: explorations in ethnoprimatology. In Kinzey, W. G. (ed.) New World Primates. New York: Aldine de Gruyter, pp. 143–65.Google Scholar
Strier, K. B. (2011). Conservation. In Campbell, C. J., Fuentes, A., MacKinnon, K. C., Bearder, S. K. & Stumpf, R. M. (eds) Primates in Perspective. New York: Oxford University Press, pp. 664–75.Google Scholar
Struhsaker, T. T. (1997). Ecology of an African Rain Forest: Logging in Kibale and the Conflict Between Conservation and Exploitation. Gainesville, FL: University of Florida Press.Google Scholar
Sunderland, T. C. H., Comiskey, J. A., Besong, S., et al. (2003). Vegetation assessment of Takamanda Forest Reserve, Cameroon. In Comiskey, J. A., Sunderland, T. C. H. & Sunderland-Groves, J. L. (eds) Takamanda: The Biodiversity of an African Rainforest. Washington, DC: Smithsonian Institution, pp. 1954.Google Scholar
Sunderland-Groves, J. L., Ekinde, A. & Mboh, H. (2009). Cross River gorilla (Gorilla gorilla diehli) nesting behaviour at Kagwene Mountain, Cameroon: implications for assessing group size, age structure and density at other Cross River gorilla localities. International Journal of Primatology, 30, 253–66.Google Scholar
SYSTAT (2009). SYSTAT Version 13. SYSTAT Software. Available at: www.systatsoftware.com.Google Scholar
Thalmann, O., Wegmann, D., Spitzner, M., et al. (2011). Historical sampling reveals dramatic demographic changes in western gorilla populations. BMC Evolutionary Biology, 11, 8595.Google Scholar
Tutin, C. E. G. (1996). Ranging and social structure of lowland gorillas in the Lope Reserve, Gabon. In McGrew, W. C., Marchant, L. F. & Nishida, T. (eds) Great Ape Societies. Cambridge: Cambridge University Press, pp. 5870.Google Scholar
Tutin, C. E. G., Parnell, R. J., White, L. J. T. & Fernandez, M. (1995). Nest building by lowland gorillas in the Lopé Reserve, Gabon: environmental influences and implications for censusing. International Journal of Primatology, 16, 5376.Google Scholar
Tutin, C. E. G., White, L. J. T. & Mackanga-Missandzouo, A. (1997). The use by rain forest mammals of natural forest fragments in an equatorial African savanna. Conservation Biology, 11, 1190–203.Google Scholar
Warren, Y. & Ekinde, A. (2007). Large mammal recce survey of the Mone Forest Reserve. Unpublished report. Wildlife Conservation Society.Google Scholar
White, F. (1983). The Vegetation of Africa, a Descriptive Memoir to Accompany the UNESCO/AETFAT/UNSO Vegetation Map of Africa (3 plates, Northwestern African, Northeastern Africa, and Southern Africa, 1:5,000,000). Paris: UNESCO.Google Scholar
White, L. & Edwards, A. (2000). Methods for assessing the status of animal populations. In White, L. & Edwards, A. (eds) Conservation Research in the African Rain Forests: A Technical Handbook. New York: Wildlife Conservation Society, pp. 191201.Google Scholar

References

Alvard, M. S., Robinson, J. G., Redford, K. H. & Kaplan, H. (1997). The sustainability of subsistence hunting in the Neotropics. Conservation Biology, 11, 977–82.Google Scholar
Balée, W. (2015). Cultural Forests of the Amazon: A Historical Ecology of People and Their Landscapes. Tuscaloosa, AL: University of Alabama Press.Google Scholar
Barnett, A. A., Borges, S. H., de Castilho, C. V., Neri, F. M. & Shapley, R. L. (2002). Primates of the Jaúnational park, Amazonas, Brazil. Neotropical Primates 10, 6570.Google Scholar
Bodmer, R. E. (1994). Managing wildlife with local communities: the case of the Reserva Comunal Tamshiyacu-Tahuayo. In Western, D., Wright, M. & Strum, S. (eds) Natural Connections. Washington, DC: Island Press, pp. 113–34.Google Scholar
Bodmer, R. & Puertas, P. E. (2000). Community-based comanagment of wildlife in the Peruvian Amazon. In Robinson, J. G. and Bennet, E. L. (eds) Hunting for Sustainability in Tropical Forests. New York: Columbia University Press, pp. 395409.Google Scholar
Bodmer, R. & Robinson, J. (2004). Evaluating the sustainability of hunting in the neotropics. In Silvius, K., Bodmer, R. & Fragoso, J. (eds) People in Nature: Wildlife Conservation in South and Central America. New York: Columbia University Press, pp. 299323.Google Scholar
Bodmer, R. E., Eisenberg, J. F. & Redford, K. H. (1997). Hunting and the likelihood of extinction of Amazonian mammals. Conservation Biology, 11, 460–6.Google Scholar
Carneiro, R. L. (1970). Hunting and hunting magic among the Amahuaca of the Peruvian Montana. Ethnology, 9, 331–41.Google Scholar
Cartelle, C. & Hartwig, W. C. (1996). A new extinct primate among the Pleistocene megafauna of Bahia, Brazil. Proceedings of the National Academy of Sciences, 93, 6405–9.Google Scholar
Chagnon, N. A. (1968). Yanomamo: The Fierce People. New York: Holt, Rinehart and Winston.Google Scholar
Constantino, P. A. L., Fortini, L. B., Kaxinawa, F. R. S., et al. (2008). Indigenous collaborative research for wildlife management in Amazonia: the case of the Kaxinawa, Acre, Brazil. Biological Conservation, 141, 27182729.Google Scholar
Cormier, L. A. (2003). Kinship with Monkeys: The Guaja Foragers of Eastern Amazonia. New York: Columbia University Press.Google Scholar
Cormier, L. A. (2006). A preliminary review of neotropical primates in the subsistence and symbolism of indigenous lowland South American people. Ecological and Environmental Anthropology, 2, 1432.Google Scholar
Crocker, J. C. (1985). Vital Souls, Bororo Cosmology, Natural Symbolism, and Shamanism. Tucson, AZ: University of Tucson Press.Google Scholar
de Souza-Mazurek, D., Roselis, R., Pedrinho, T., et al. (2000). Subsistence hunting among the Waimiri Atroari Indians in Central Amzonia, Brazil. Biodiversity and Conservation, 9, 579–96.Google Scholar
de Thoisy, B., Renoux, F. & Julliot, C. (2005). Hunting in northern French Guiana and its impacts on primate communities. Oryx, 39, 149–57.Google Scholar
de Thoisy, B., Richard-Hansen, C. & Peres, C. A. (2009). Impacts of subsistence game hunting on Amazonian primates. In Garber, P. A., Estrada, A., Bicca-Marques, J. C., Heymann, E. W. & Strier, K. B. (eds) South American Primates: Comparative Perspectives in the Study of Behavior. New York: Springer, pp. 389412.Google Scholar
Edmundson, R. G. (1922). Journal of the Travels of Father Samuel Fritz in the River of the Amazonas between 1686 and 1723. London: Hakluyt Society.Google Scholar
Endo, W., Peres, C. A., Salas, E., et al. (2010). Game vertebrate densities in hunted and nonhunted forest sites in Manu National Park, Peru. Biotropica, 42, 251–61.Google Scholar
Erikson, P. (2001). Myth and material culture: Matis blowguns, palm trees, and ancestor spirits. In Rival, L. and Whitehead, N. (eds) Beyond the Visible and the Material: The Amerindianization of Society in the Work of Peter Rivière. Oxford: Oxford University Press, pp. 101121Google Scholar
Fa, J. E., Peres, C. A. & Meeuwig, J. (2002). Bushmeat exploitation in tropical forests: an international comparison. Conservation Biology, 16, 232–37.Google Scholar
Franzen, M. (2006). Evaluating the sustainability of hunting: a comparison of harvest profiles across three Huaorani communities. Environmental Conservation, 33, 3645.Google Scholar
Freese, C. H., Heltne, P. G., Castro, N. & Whitesides, G. (1982). Patterns and determinants of monkey densities in Peru and Bolivia, with notes on distribution. International Journal of Primatology 3, 5390.CrossRefGoogle Scholar
Fricke, T. (2005). Taking culture seriously: making the social survey ethnographic. In Weisner, T. S. (ed.) Discovering Successful Pathways in Children’s Development: Mixed Methods in the Study of Childhood and Family Life. Chicago, IL: University of Chicago Press, pp. 185218.Google Scholar
Gross, D. R. (1975). Protein capture and cultural development in the Amazon Basin. American Anthropologist, 77, 526–49.Google Scholar
Hamas, R. (1991). Wildlife conservation in tribal societies. In Oldfield, M. & Alcorn, J. (eds) Biodiversity: Culture, Conservation, and Ecodevelopment. Denver, CO: Westview, pp. 172–99.Google Scholar
Hamas, R. (2007). The ecologically noble savage debate. Annual Review of Anthropology, 36, 177190.Google Scholar
Heckenberger, M. J. (1998). Manioc agriculture and sedentism in Amazonia: the Upper Xingu example. Antiquity, 72, 633–48.Google Scholar
Henfrey, T. B. (2002). Ethnoecology, resource use, conservation, and development in a Wapishana community in South Rupununi, Guyana. PhD Thesis, University of Kent at Canterbury.Google Scholar
Hill, K., McMillan, G. & Farina, R. (2003). Hunting-related changes in game encounter rates from 1994 to 2001 in the Mbaracayu Reserve, Paraguay. Conservation Biology, 17, 1312–23.Google Scholar
Holmberg, A. R. (1969). Nomads of the Long Bow: The Siriono of Eastern Bolivia. Prospect Heights, NY: Waveland Press.Google Scholar
Iwamura, T., Lambin, E. F., Silvius, K. M., Luzar, J. B. & Fragoso, J. M. V. (2014). Agent-based modeling of hunting and subsistence agriculture on indigenous lands: understanding interactions between social and ecological systems. Environmental Modelling and Software, 58, 109–27.Google Scholar
Jerozolimski, A. & Peres, C. A. (2003). Bringing home the biggest bacon: a cross-site analysis of the structure of hunter-kill profiles in Neotropical forests. Biological Conservation, 111, 415425.Google Scholar
Kensinger, K. M., Rabineau, P., Tanner, H., Ferguson, S. G. & Dawson, A. (1975). The Cashinahua of Eastern Peru. In Dwyer, J. P. (ed.) Studies in Anthropology and Material Culture, Volume 1. Providence, RI: The Haffenferrer Museum of Anthropology, Brown University.Google Scholar
Koster, J. (2008). The impact of hunting with dogs on wildlife harvests in the Bosawas Reserve, Nicaragua. Environmental Conservation, 35, 211–20.Google Scholar
Levi, T., Shepard, G. H. Jr, Ohl-Schacherer, J., Peres, C. A. & Yu, D. W. (2009). Modeling the long-term sustainability of indigenous hunting in Manu National Park, Peru: landscape-scale management implications for Amazonia. Journal of Applied Ecology, 46, 804–14.Google Scholar
Levi, T., Shepard, G. H., Ohl-Schacherer, J., et al. (2011). Spatial tools for modeling the sustainability of subsistence hunting in tropical forests. Ecological Applications, 21, 18021818.Google Scholar
Levy, R. I. & Hollan, D. W. (2014). Person-centered interviewing and observation. In Bernard, R. H. & Gravlee, C. C. (eds) Handbook of Methods in Cultural Anthropology. London: Rowman & Littlefield, pp. 313342.Google Scholar
Lizarralde, M. (2002). Ethnoecology of monkeys among the Bari of Venezuela: perception, use, and conservation. In Fuentes, A. & Wolfe, L. D. (eds) Primates Face to Face: The Conservation Implications of Human and Nonhuman Primate Interconnections. Cambridge: Cambridge University Press, pp. 85100.Google Scholar
Meggars, B. J. (1971). Amazonia: Man and Culture in a Counterfeit Paradise. Chicago, IL: Aldine, Atherton, Inc.Google Scholar
Mena, V. P., Stallings, J. R., Regalado, J. B. & Cueva, R. L. (2000). The sustainability of current hunting practices by the Huaorani. In Robinson, J. G. & Bennett, E. L. (eds) Hunting for Sustainability in Tropical Forests. New York: Columbia University Press, pp. 5778.Google Scholar
Mentore, G. P. (2005). Of Passionate Curves and Desirable Cadences: Themes on Waiwai Social Being. London: University of Nebraska Press.Google Scholar
Milner-Gulland, E. J. & Akcakaya, H. R. (2001). Sustainability indices for exploited populations. Trends in Ecology and Evolution, 18, 351–7.Google Scholar
Mittermeier, R. A. (1991). Hunting and its effects on wild primate populations in Suriname. In Robinson, J. G. & Redford, K. H. (eds) Neotropical Wildlife Use and Conservation. Chicago, IL: University of Chicago Press, pp. 93106.Google Scholar
Nasi, R., Brown, D., Wilkie, D., et al. (2008). Conservation and Use of Wildlife-Based Resources: The Bushmeat Crisis. Bogor: Secretariat of the Convention on Biological Diversity and Center for International Forestry Research (CIFOR).Google Scholar
Nasi, R., Taber, A. & Van Vliet, N. (2011). Empty forests, empty stomachs? Bushmeat and livelihoods in the Congo and Amazon basins. International Forestry Review, 13, 355–68.Google Scholar
Nepstad, D., Schwartzman, S., Bamberger, B., et al. (2006). Inhibition of Amazon deforestation and fire by parks and indigenous lands. Conservation Biology, 20, 6573.Google Scholar
Novaro, A. J., Redford, K. H. & Bodmer, R. E. (2000). Effect of hunting in source–sink systems in the Neotropics. Conservation Biology, 14, 713–21.Google Scholar
Nuñez-Iturri, G. & Howe, H. F. (2007). Bushmeat and the fate of trees with seeds dispersed by large primates in a lowland rainforest in western Amazonia. Biotropica, 39, 348–54.Google Scholar
Ohl, J., Wezel, A., Shepard, G. H., Yu, D. W. (2008). Swidden agriculture in a protected area: the Matsigenka native communities of Manu National Park, Peru. Environmental Development and Sustainability, 10, 827–43.Google Scholar
Ohl-Schacherer, J., Shepard, G. H. Jr, Kaplan, H., et al. (2007). The sustainability of subsistence hunting by Matsigenka native communities in Manu National Park, Peru. Conservation Biology, 21, 1174–85.Google Scholar
Oliver, J. R. (2001). The archaeology of forest foraging and agricultural production in Amazonia. In McEwan, C., Neves, E. G. & Barreto, C. (eds) The Unknown Amazon: Culture in Nature in Ancient Brazil. London: British Museum Press, pp. 5085.Google Scholar
Parathian, H. E. & Maldonado, A. M. (2010). Human–nonhuman primate interactions amongst Tikuna people: perceptions and local initiatives for resource management in Amacayacu in the Colombian Amazon. American Journal of Primatology, 72, 855–65.Google Scholar
Peres, C. A. (1990). Effects of hunting on western Amazonian primate communities. Biological Conservation, 54, 4759.Google Scholar
Peres, C. A. (1994). Indigenous reserves and nature conservation in Amazonian forests. Conservation Biology, 8, 586–88.Google Scholar
Peres, C. A. (2000). Effects of subsistence hunting on vertebrate community structure in Amazonian forests. Conservation Biology, 17, 240–53.Google Scholar
Peres, C. A. & Nascimento, H. S. (2006). Impact of game hunting by the Kayapó of southeastern Amazonia: implications for wildlife conservation in Amazonian indigenous reserves. Biodiversity and Conservation, 15, 2627–53.Google Scholar
Peres, C. A. & Palacios, E. (2007). Basin-wide effects of game harvest on vertebrate population densities in Amazonian forests: implications for animal-mediated seed dispersal. Biotropica, 39, 304–15.Google Scholar
Peres, C. A., Emilio, T., Schietti, J., Desmoulière, S. J. M. & Levi, T. (2016). Dispersal limitation induces long-term biomass collapse in overhunted Amazonian forests. Proceedings of the National Academy of Sciences, 113, 892–7.Google Scholar
RAISG (2012). Amazonía bajo presión. Available at: www.raisg.socioambiental.org.Google Scholar
Redford, K. (1991). The ecologically noble savage. Orion 9, 24–9.Google Scholar
Redford, K. H. & Robinson, J. G. (1991). Subsistence and commercial uses of wildlife in Latin America. In Robinson, J. G. & Redford, K. H. (eds) Neotropical Wildlife Use and Conservation. Chicago, IL: University of Chicago Press, pp. 623.Google Scholar
Richard-Hansen, C. & Hansen, E. (2004). Hunting and wildlife management in French Guiana: current aspects and future prospects. In Silvius, K., Bodmer, R. & Fragoso, J. (eds) People in Nature: Wildlife Conservation in South and Central America. New York: Columbia University Press, pp. 400410.Google Scholar
Robinson, J. G. & Bennett, E. L. (2004). Having your wildlife and eating it too: an analysis of hunting sustainability across tropical ecosystems. Animal Conservation, 7, 397408.Google Scholar
Robinson, J. G. & Redford, K. H. (1991). Sustainable harvest of Neotropical wildlife. In Robinson, J. G. & Redford, K. H. (eds) Neotropical Wildlife Use and Conservation. Chicago, IL: University of Chicago Press, pp. 415–29.Google Scholar
Robinson, J. G. & Redford, K. H. (1994). Measuring the sustainability of hunting in tropical forests. Oryx, 28, 249–56.Google Scholar
Ross, E. B. (1978). Food taboos, diet, and hunting strategy: the adaptation to animals in Amazon cultural ecology. Current Anthropology, 19, 136.Google Scholar
Salas, L. A. & Kim, J. B. (2002). Spatial factors and stochasticity in the evaluation of sustainable hunting of tapirs. Conservation Biology, 16, 8696.Google Scholar
Schomburgk, R. (1836). Report of an expedition into the interior of British Guiana in 1835–1836. Journal of the Royal Geographical Society, 4, 224–84.Google Scholar
Schwartzman, S. & Zimmerman, B. (2005). Conservation alliances with indigenous peoples of the Amazon. Conservation Biology, 19, 721–7.Google Scholar
Shaffer, C. A., Milstein, M. S., Yukuma, C., Marawanaru, E. & Suse, P. (2017a). Sustainability and comanagement of subsistence hunting in an indigenous reserve in Guyana. Conservation Biology, 31, 119–31.Google Scholar
Shaffer, C. A., Marawanaru, E. & Yukuma, C. (2017b). An ethnoprimatological approach to assessing the sustainability of primate subsistence hunting of indigenous Waiwai in the Konashen Community Owned Conservation Concession, Guyana. In Dore, K. M., Riley, E. P. & Fuentes, A. (eds) Ethnoprimatology: A Practical Guide to Research on the Human–Nonhuman Primate Interface. Cambridge: Cambridge University Press, pp. 219–31.Google Scholar
Shaffer, C. A., Yukuma, C., Marawanaru, E. & Suse, P. (2017c). Assessing the sustainability of Waiwai subsistence hunting in Guyana by comparison of static indices and spatially explicit, biodemographic models. Animal Conservation. DOI: 10.1111/acv.12366.Google Scholar
Shaffer, C. A., Milstein, M. S., Yukuma, C., et al. (2018). Integrating ethnography and hunting sustainability modeling for primate conservation in an indigenous reserve in Guyana. International Journal of Primatology. In review.Google Scholar
Shepard, G. H. (2002). Primates in Matsigenka: subsistence and world view. In Fuentes, A. & Wolfe, L. D. (eds) Primates Face to Face: The Conservation Implications of Human and Nonhuman Primate Interconnections. Cambridge: Cambridge University Press, pp. 101–36.Google Scholar
Shepard, G. H., Rummenhoeller, K., Ohl, J. & Yu, D. W. (2010). Trouble in paradise: indigenous populations, anthropological policies, and biodiversity conservation in Manu National Park, Peru. Journal of Sustainable Forestry, 29, 252301.Google Scholar
Shepard, G. H., Levi, T., Neves, E. G., Peres, C. A. & Yu, D. W. (2012). Hunting in ancient and modern Amazonia: rethinking sustainability. American Anthropologist, 114, 652–67.Google Scholar
Sirén, A. & Parvinen, K. (2015). A spatial bioeconomic model of the harvest of wild plants and animals. Ecological Economics, 116, 201–10.Google Scholar
Sirén, A., Hambäck, P. & Machoa, J. (2004). Including spatial heterogeneity and animal dispersal when evaluating hunting: a model analysis and an empirical assessment in an Amazonian community. Conservation Biology, 18, 1315–29.Google Scholar
Stafford, C. A., Preziosi, R. F. & Sellers, W. I. (2017). A pan-neotropical analysis of hunting preferences. Biodiversity and Conservation, 26, 1877–97.Google Scholar
Townsend, W. R. (2000). The sustainability of subsistence hunting by the Sirionó Indians of Bolivia. In Robinson, J. G. & Bennett, E. L. (eds) Hunting for Sustainability in Tropical Forests. New York: Columbia University Press, pp. 267–82.Google Scholar
Van Vliet, N., Fa, J. & Nasi, R. (2015). Managing hunting under uncertainty: from one-off ecological indicators to resilience approaches in assessing the sustainability of bushmeat hunting. Ecology and Society, 20. DOI: 10.5751/ES-07669-200307.Google Scholar
Verissimo, A., Barreto, P., Tarifa, R. & Uhl, C. (1995). Extraction of a high-value natural resource from Amazonia: the case of mahogany. Forest Ecology and Management, 72, 3960.Google Scholar
Wallace, A. R. (1835). A Narrative of Travels on the Amazon and Rio Negro. London: Reeve and Co.Google Scholar
Waterton, C. (1825). Wanderings in South America, the North-West of the United States, and the Antilles. London: Cassell and Company.Google Scholar
Weinbaum, K. Z., Brashares, J. S., Golden, C. D. & Getz, W. M. (2013). Searching for sustainability: are assessments of wildlife harvests behind the times? Ecology Letters, 16, 99111.Google Scholar
Yde, J. (1965). Material culture of the Waiwai. Copenhagen: National Museum of Denmark.Google Scholar
Zapata-Ríos, G., Urgilés, C. & Suárez, E. (2009). Mammal hunting by the Shuar of the Ecuadorian Amazon: is it sustainable? Oryx, 43, 375–85.Google Scholar
Zimmerman, B., Peres, C. A., Malcolm, J. & Turner, T. (2001). Conservation and development alliances with the Kayapo of south-eastern Amazonia, a tropical forest indigenous peoples. Environmental Conservation, 28, 1022.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
×