Skip to main content Accessibility help
×
Hostname: page-component-848d4c4894-p2v8j Total loading time: 0 Render date: 2024-05-01T05:53:47.824Z Has data issue: false hasContentIssue false

References

Published online by Cambridge University Press:  29 February 2020

K. A. I. Nekaris
Affiliation:
Oxford Brookes University
Anne M. Burrows
Affiliation:
Duquesne University, Pittsburgh
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: 2020

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

Abbott, D. H., Barnett, D. K., Colman, R. J., Yamamoto, M. E., & Schultz-Darken, N. J. (2003). Aspects of common marmoset basic biology and life history important for biomedical research. Comparative Medicine, 53(4), 339350.Google Scholar
Abdullah, M., Mamat, M. P., Yaacob, M. R., Radam, A., & Fui, L. H. (2015). Estimate the conservation value of biodiversity in national heritage site: A case of Forest Research Institute Malaysia. Procedia Environmental Sciences, 30, 180185.CrossRefGoogle Scholar
Abugiche, S. A. (2008). Impact of hunting and bushmeat trade on biodiversity loss in Cameroon: A case study of the Banyang-Mbo Wildlife Sanctuary (unpublished doctoral dissertation). Brandenburg University of Technology (BTU), Cottbus, Germany.Google Scholar
Adams, L., Wilkinson, F., & MacDonald, S. E. (2017). Limits of spatial vision in Sumatran orangutans (Pongo abelii). Animal Behavior and Cognition, 4(3), 204222.CrossRefGoogle Scholar
Adipietro, K. A., Mainland, J. D., & Matsunami, H. (2012). Functional evolution of mammalian odorant receptors. PLoS Genetics, 8(7), e1002821.CrossRefGoogle ScholarPubMed
Aerts, P. (1998). Vertical jumping in Galago senegalensis: The quest for an obligate mechanical power amplifier. Philosophical Transactions of the Royal Society B: Biological Sciences, 353(1375), 16071620.CrossRefGoogle Scholar
Ahl, A. S. (1986). The role of vibrissae in behavior: A status review. Veterinary Research Communications, 10(1), 245268.CrossRefGoogle ScholarPubMed
Ahnelt, P. K., & Kolb, H. (2000). The mammalian photoreceptor mosaic-adaptive design. Progress in Retinal and Eye Research, 19(6), 711777.CrossRefGoogle ScholarPubMed
Akani, G. C., Amadi, N., Eniang, E. A., Luiselli, L., & Petrozzi, F. (2015). Are mammal communities occurring at a regional scale reliably represented in ‘hub’ bushmeat markets? A case study with Bayelsa State (Niger Delta, Nigeria). Folia Zoologica, 64(1), 7987.Google Scholar
Albouy, J., & Décaudin, J. M. (2018). Age differences in responsiveness to shocking prosocial campaigns. Journal of Consumer Marketing, 35(3), 328339.CrossRefGoogle Scholar
Aldrich, B. C., Molleson, L., & Nekaris, K. A. I. (2008). Vocalizations as a conservation tool: An auditory survey of the Andean titi monkey Callicebus oenanthe Thomas, 1924 (Mammalia: Primates: Pitheciidae) at Tarangue, Northern Peru. Contributions to Zoology, 77(1), 16.CrossRefGoogle Scholar
Alexa Traffic Rank. (2018). The top 500 sites on the web. Retrieved 21 October 2018, from www.alexa.com/topsites.Google Scholar
Alexander, R. M. (1981). The gaits of tetrapods: Adaptations for stability and economy. In Symposia of the Zoological Society of London, 48, 269287.Google Scholar
Allman, J. (1977). Evolution of the visual system in the early primates. Progress in Psychobiology and Physiological Psychology, 7(1), 53.Google Scholar
Allman, J. (1982). Reconstructing the evolution of the brain in primates through the use of comparative neurophysiological and neuroanatomical data. In Armstrong, E. & Falk, D. (Eds.), Primate brain evolution (pp. 1328). Boston, MA: Springer.CrossRefGoogle Scholar
Alterman, L. (1995). Toxins and toothcombs: Potential allospecific chemical defenses in Nycticebus and Perodicticus. In Alterman, L., Doyle, G. A., & Izard, M. K. (Eds.), Creatures of the dark (pp. 413424). Boston, MA: Springer.CrossRefGoogle Scholar
Alterman, L., & Freed, B. Z. (1997). Description and survey of three Nycticebus species in Bolikhamxay Province, Laos: Primate Society of Great Britain Scientific Meeting – New Perspectives on Nocturnal Primates, December 3, 1997. Primate Eye, 63, 16.Google Scholar
Altmann, J. (1974). Observational study of behavior: Sampling methods. Behaviour, 49(3–4), 227266.CrossRefGoogle ScholarPubMed
Alves, R. R., Souto, W. M., & Barboza, R. R. (2010). Primates in traditional folk medicine: A world overview. Mammal Review, 40(2), 155180.Google Scholar
Amato, K. R. (2016). Incorporating the gut microbiota into models of human and non‐human primate ecology and evolution. American Journal of Physical Anthropology, 159, 196215.CrossRefGoogle ScholarPubMed
Ambrose, L. (1999). Species diversity in West and Central African galagos (Primates, Galagonidae): The use of acoustic analysis (unpublished doctoral dissertation). Oxford Brookes University, Oxford, UK.Google Scholar
Ambrose, L. (2006). A survey of prosimians in the national parks and forest reserves of Uganda. In Newton-Fisher, N. E., Notman, H., Paterson, J. D., & Reynolds, V. (Eds.), Primates of Western Uganda (pp. 329343). New York, NY: Springer.CrossRefGoogle Scholar
Ambrose, L. (2013). Arctocebus aureus: Golden angwantibo. In Butynski, T. M., Kingdon, J., & Kalina, J. (Eds.), Mammals of Africa (Volume II: Primates) (pp. 402403). London: Bloomsbury Publishing.Google Scholar
Anadu, P. A., Elamah, P. O., & Oates, J. F. (1988). The bushmeat trade in southwestern Nigeria: A case study. Human Ecology, 16(2), 199208.Google Scholar
Anand, S., & Radhakrishna, S. (2017). Investigating trends in human–wildlife conflict: Is conflict escalation real or imagined?. Journal of Asia-Pacific Biodiversity, 10(2), 154161.CrossRefGoogle Scholar
Andrews, C. A., Rambeloarivony, H., Génin, F., & Masters, J. C. (2016). Why cheirogaleids are bad models for primate ancestors: A phylogenetic reconstruction. In Lehman, S. M., Radespiel, U., & Zimmermann, E. (Eds.), The dwarf and mouse lemurs of Madagascar: Biology, behavior and conservation biogeography of the Cheirogaleidae (pp. 94112). Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Anjum, F., Turni, H., Mulder, P. G., van der Burg, J., & Brecht, M. (2006). Tactile guidance of prey capture in Etruscan shrews. Proceedings of the National Academy of Sciences, 103(44), 1654416549.CrossRefGoogle ScholarPubMed
Ankel-Simons, F. (1983). A survey of living primates and their anatomy. New York, NY: Macmillan.Google Scholar
Anonymous. (2017). Number of internet users in Indonesia. Retrieved 5 August 2017, from www.statista.com/statistics/254456/number-of-internet-users-in-indonesia.Google Scholar
Anstis, S., Cavanagh, P., Maurer, D., & Lewis, T. (1987). Optokinetic technique for measuring infants’ responses to color. Applied Optics, 26(8), 15101516.CrossRefGoogle ScholarPubMed
Appleby, E. (2005). Mrs Blue Gum, some puppets and a remnant forest: Towards sustainability education through drama pedagogy. Australian Journal of Environmental Education, 21, 110.CrossRefGoogle Scholar
Araújo, C. S., Corrêa, L. P. D., da Silva, A. P. C., Prates, R. O., & Meira, W. (2014). It is not just a picture: Revealing some user practices in Instagram (pp. 1923), 2014 9th Latin American Web Congress, Brazil, 22–24 October 2014. New York, NY: IEEE.Google Scholar
Archer, J., & Monton, S. (2011). Preferences for infant facial features in pet dogs and cats. Ethology, 117(3), 217226.CrossRefGoogle Scholar
Arkley, K., Tiktak, G. P., Breakell, V., Prescott, T. J., & Grant, R. A. (2017). Whisker touch guides canopy exploration in a nocturnal, arboreal rodent, the Hazel dormouse (Muscardinus avellanarius). Journal of Comparative Physiology A, 203(2), 133142.Google Scholar
Arrese, C., Dunlop, S. A., Harman, A. M., Braekevelt, C. R., Ross, W. M., Shand, J., & Beazley, L. D. (1999). Retinal structure and visual acuity in a polyprotodont marsupial, the fat-tailed dunnart (Sminthopsis crassicaudata). Brain, Behavior and Evolution, 53(3), 111126.CrossRefGoogle Scholar
Arrese, C., Archer, M., Runham, P., Dunlop, S. A., & Beazley, L. D. (2000). Visual system in a diurnal marsupial, the numbat (Myrmecobius fasciatus): Retinal organization, visual acuity and visual fields. Brain, Behavior and Evolution, 55(4), 163175.CrossRefGoogle Scholar
Asher, R. J. (1998). Morphological diversity of anatomical strepsirrhinism and the evolution of the lemuriform toothcomb. American Journal of Physical Anthropology, 105(3), 355367.3.0.CO;2-Q>CrossRefGoogle ScholarPubMed
Ashwell, D., & Walston, N. (2008). An overview of the use and trade of plants and animals in traditional medicine systems in Cambodia. Ha Noi: TRAFFIC Southeast Asia, Greater Mekong Programme.Google Scholar
Atsalis, S. (2008). A natural history of the brown mouse lemur. Upper Saddle River, NJ: Pearson Prentice Hall.Google Scholar
AZA (Association of Zoos and Aquariums). (2014). Prosimians: Primates. Retrieved 20 April 2018, from www.stlzoo.org/animals/scienceresearch/reproductivemanagementcenter/contraceptionrecommendatio/contraceptionmethods/prosimians.Google Scholar
Azuma, D. L., Baldwin, J. A., & Noon, B. R. (1990). Estimating the occupancy of spotted owl habitat areas by sampling and adjusting for bias. Berkeley, CA: Pacific Southwest Research Station, Forest Service, US Department of Agriculture.CrossRefGoogle Scholar
Bäckman, J., Andersson, A., Alerstam, T., Pedersen, L., Sjöberg, S., Thorup, K., & Tøttrup, A. P. (2017). Activity and migratory flights of individual free‐flying songbirds throughout the annual cycle: Method and first case study. Journal of Avian Biology, 48(2), 309319.CrossRefGoogle Scholar
Baer, C. K., Ullrey, D. E., Schlegel, M. L., Agoramooethy, G., & Baer, D. J. (2010). Contemporary topics in wild mammal nutrition. In Kleiman, D. G., Thompson, K. V., & Baer, C. K. (Eds.), Wild Mammals in Captivity: Principals and Techniques for Zoo Management (pp. 85103). Chicago, IL: University of Chicago Press.Google Scholar
Baird, I. G. (1995). Lao PDR: An overview of traditional medicines derived from wild animals and plants. Selangor: TRAFFIC Southeast Asia.Google Scholar
Bajpai, S. U. N. I. L., Kapur, V. V., Thewissen, J. G. M., Das, D. P., Tiwari, B. N., Sharma, R. I. T. U., & Saravanan, N. (2005). Early Eocene primates from Vastan lignite mine, Gujarat, western India. Journal of the Palaeontological Society of India, 50(2), 4354.Google Scholar
Bajpai, S., Kay, R. F., Williams, B. A., Das, D. P., Kapur, V. V., & Tiwari, B. N. (2008). The oldest Asian record of Anthropoidea. Proceedings of the National Academy of Sciences, 105(32), 1109311098.CrossRefGoogle ScholarPubMed
Baker, S. E., Cain, R., Van Kesteren, F., Zommers, Z. A., D’cruze, N., & Macdonald, D. W. (2013). Rough trade: Animal welfare in the global wildlife trade. BioScience, 63(12), 928938.Google Scholar
Balestri, M. (2018). Ecology and conservation of the southern woolly lemur (Avahi meridionalis) in the Tsitongambarika Protected Area, south-eastern Madagascar (unpublished doctoral dissertation). Oxford Brookes University, Oxford, UK.Google Scholar
Barber-Meyer, S. M. (2010). Dealing with the clandestine nature of wildlife‐trade market surveys. Conservation Biology, 24(4), 918923.Google Scholar
Barrett, E. (1981). The present distribution and status of the slow loris in peninsular Malaysia. Malaysian Applied Biology, 10(2), 205211.Google Scholar
Barrett, E. (1984). The ecology of some nocturnal, arboreal mammals in the rainforest of Peninsular Malaysia (unpublished doctoral dissertation). University of Cambridge, Cambridge, UK.Google Scholar
Bartlett Society. (2018). First and early breeding records for mammals in the UK and Eire. Retrieved 7 August 2018, from www.zoohistory.co.uk/projects/first_breedings/primates.Google Scholar
Barton, R. A. (2006a). Olfactory evolution and behavioral ecology in primates. American Journal of Primatology: Official Journal of the American Society of Primatologists, 68(6), 545558.Google Scholar
Barton, R. A. (2006b). Primate brain evolution: Integrating comparative, neurophysiological, and ethological data. Evolutionary Anthropology: Issues, News, and Reviews, 15(6), 224236.CrossRefGoogle Scholar
Barton, R. A., Purvis, A., & Harvey, P. H. (1995). Evolutionary radiation of visual and olfactory brain systems in primates, bats and insectivores. Philosophical Transactions of the Royal Society B: Biological Sciences, 348(1326), 381392.Google Scholar
Bastir, M., Rosas, A., Stringer, C., Cuétara, J. M., Kruszynski, R., Weber, , …, & Ravosa, M. J. (2010). Effects of brain and facial size on basicranial form in human and primate evolution. Journal of Human Evolution, 58(5), 424431.CrossRefGoogle ScholarPubMed
Bauer, G. B., Gaspard, J. C. III, Colbert, D. E., Leach, J. B., Stamper, S. A., Mann, D., & Reep, R. (2012). Tactile discrimination of textures by Florida manatees (Trichechus manatus latirostris). Marine Mammal Science, 28(4), E456E471.CrossRefGoogle Scholar
Beard, K. C. (1990). Gliding behaviour and palaeoecology of the alleged primate family Paromomyidae (Mammalia, Dermoptera). Nature, 345(6273), 340.Google Scholar
Beard, K. C. (1991). Vertical postures and climbing in the morphotype of Primatomorpha: Implications for locomotor evolution in primate history. In Origines de la Bipédie chez les Hominidés (pp. 7987). Paris: Editions du CNRS (Centre national de la recherche scientifique)(Cahiers de Paléoanthropologie).Google Scholar
Beard, K. C., Dagosto, M., Gebo, D. L., & Godinot, M. (1988). Interrelationships among primate higher taxa. Nature, 331(6158), 712714.CrossRefGoogle ScholarPubMed
Bearder, S. K. (1987). Lorises, bushbabies and tarsiers: Diverse societies in solitary foragers. In Smuts, B. B., Cheney, D. L., Seyfarth, R. M., Wrangham, R. W., & Struhsaker, T. T. (Eds.), Primate Societies (pp. 1124). Chicago, IL: University of Chicago Press.Google Scholar
Bearder, S. K. (2007). A comparison of calling patterns in two nocturnal primates, Otolemur crassicaudatus and Galago moholi as a guide to predation risk. In Gursky, S. L. & Nekaris, K. A.I. (Eds.), Primate anti-predator strategies (pp. 206221). Boston, MA: Springer.Google Scholar
Bearder, S. K., & Doyle, G. A. (1974). Field and laboratory studies of social organisation in bushbabies (Galago sengalensis). Journal of Human Evolution, 3, 3750.Google Scholar
Bearder, S. K., & Honess, P. E. (1992). A survey of nocturnal primates and other mammals in the Korup National Park, Cameroon. New York, NY: Wildlife Conservation International.Google Scholar
Bearder, S. K., & Martin, R. D. (1980a). Acacia gum and its use by bushbabies, Galago senegalensis (Primates: Lorisidae). International Journal of Primatology, 1(2), 103128.Google Scholar
Bearder, S. T., & Martin, R. D. (1980b). The social organization of a nocturnal primate revealed by radio tracking. In Amlaner, C. J. & Macdonald, D. W. (Eds.), A handbook on biotelemetry and radio tracking (pp. 633648). Oxford: Pergamon Press.CrossRefGoogle Scholar
Bearder, S. K., Nekaris, K. A. I., & Buzzell, C. A. (2002). Dangers in the night: Are some nocturnal primates afraid of the dark? In Miller, L. (Ed.), Eat or be eaten: Predator sensitive foraging in primates (pp. 2143). Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Bearder, S. K., Ambrose, L., Harcourt, C., Honess, P., Perkin, A., Pimley, E., …, & Svoboda, N. (2003). Species-typical patterns of infant contact, sleeping site use and social cohesion among nocturnal primates in Africa. Folia Primatologica,74(5–6), 337354.CrossRefGoogle ScholarPubMed
Bearder, S. K., Nekaris, K. A. I., & Curtis, D. J. (2006). A re-evaluation of the role of vision in the activity and communication of nocturnal primates. Folia Primatologica, 77(1–2), 5071.Google Scholar
Bearder, S., Butynski, T. M., & Hoffmann, M. (2008a). Galago moholi. The IUCN Red List of Threatened Species. Retrieved 12 September 2019 from www.iucnredlist.org/species/8788/12932349.Google Scholar
Bearder, S., Oates, J. F., & Groves, C. P. (2008b). Arctocebus aureus. The IUCN Red List of Threatened Species. Retrieved 11 June 2018, from www.iucnredlist.org/species/2053/9211012.Google Scholar
Beerton-Joly, B. A., Piavuax, A., & Goffart, M. (1974). Quelques enzymes digestive chez un Prosimien, Perodicticus potto. Comptes rendus des sceances de la Societe de Biologie, 168, 140143.Google Scholar
Beetz, J. (2005). Role of private owners in the conservation of exotic species (Honors thesis). Colby College, Waterville, Maine, USA.Google Scholar
Behar-Cohen, F., Martinsons, C., Viénot, F., Zissis, G., Barlier-Salsi, A., Cesarini, J. P., …, & Attia, D. (2011). Light-emitting diodes (LED) for domestic lighting: Any risks for the eye?. Progress in Retinal and Eye Research, 30(4), 239257.Google Scholar
Bell, J. H., & Bromnick, R. D. (2003). The social reality of the imaginary audience: A grounded theory approach. Adolescence, 38(150), 205219.Google Scholar
Bene, J. C. K., Gamys, J., & Dufour, S. (2013). Marketing channel of hunting products in northern Nimba County, Liberia. Livestock Research for Rural Development, 25(1), 18.Google Scholar
Bennett, E. L., Caldecott, J., Kavanagh, M., & Sebastian, A. (1987). Current status of primates in Sarawak. Primate Conservation, 8, 184187.Google Scholar
Bennett, E. T. (1831). Characters of two species of Mammalia (one constituting a new genus) from Sierra Leone: Part I. Proceedings of the Committee of Science and Correspondence of the Zoological Society of London, 3, 109–12.Google Scholar
Berger, M., Vieira, M. A. R., & Guimaraes, J. A. (2012). Acute kidney injury induced by snake and arthropod venoms. In Polenakovic, M. (Ed.), Renal Failure: The Facts (pp.157186). London: IntechOpen.Google Scholar
Bernard, H. R. (2006). Research methods in anthropology (4th ed.). Oxford: AltaMira Press.Google Scholar
Bernard, H. R., & Gravlee, C. C. (2014). Handbook of methods in cultural anthropology (2nd ed.). Baltimore, MD: Rowman & Littlefield.Google Scholar
Bernede, L. (2008). Social organisation, ecology and conservation of Loris tardigradus tardigradus (Lorisiformes; Primates) (unpublished doctoral dissertation). Oxford Brookes University, Oxford, UK.Google Scholar
Bernede, L. (2009). A study of the social and behavioural ecology of the red slender loris (Loris tardigradus tardigradus) in Masmullah Proposed Forest Reserve, Sri Lanka (unpublished doctoral dissertation). Oxford Brookes University, Oxford, UK.Google Scholar
Bernede, L., Bearder, S. K., & Gunawardene, A. (2012). Habitat use by the red slender loris (Loris tardigradus tardigradus) in Masmullah Proposed Forest Reserve in Sri Lanka. In Masters, J., Gamba, M., & Génin, F. (Eds.), Leaping Ahead (pp. 7987). New York, NY: Springer.Google Scholar
Bersacola, E., Svensson, M. S., & Bearder, S. K. (2015). Niche partitioning and environmental factors affecting abundance of strepsirrhines in Angola. American Journal of Primatology, 77, 11791192.Google Scholar
Berthaume, M. A., & Schroer, K. (2017). Extant ape dental topography and its implications for reconstructing the emergence of early Homo. Journal of Human Evolution, 112, 1529.CrossRefGoogle ScholarPubMed
Bettinger, T. L., Kuhar, C. W., Lehnhardt, K., Cox, D., & Cress, D. (2010). Discovering the unexpected: Lessons learned from evaluating conservation education programs in Africa. American Journal of Primatology, 72(5), 445449.Google Scholar
Beyle, J., Bguyen, V. Q., Hendrie, D., & Nadler, T. (2014). Primates in the illegal wildlife trade in Vietnam. In Nadler, T. & Brockman, D. K. (Eds.), Primates of Vietnam (pp. 4350). Ninh Bình, Vietnam: Endangered Primate Rescue Center.Google Scholar
Bhandari, B. B., & Abe, O. (2000). Environmental education in the Asia-Pacific region: Some problems and prospects. International Review for Environmental Strategies, 1(1), 5777.Google Scholar
Bhatnagar, K. P., & Kallen, F. C. (1974). Cribriform plate of ethmoid, olfactory bulb and olfactory acuity in forty species of bats. Journal of Morphology, 142(1), 7189.Google Scholar
Bhatnagar, K. P., & Meisami, E. (1998). Vomeronasal organ in bats and primates: Extremes of structural variability and its phylogenetic implications. Microscopy Research and Technique, 43(6), 465475.Google Scholar
Bibby, C. J., Burgess, N. D., & Hill, D. A. (1992). Territory mapping techniques. In Bibby, C. J., Burgess, N. D., & Hill, D. A. (Eds.), Bird census techniques (pp. 8595). London: Academic Press.Google Scholar
Bicca-Marques, J. C., & Calegaro-Marques, C. (1995). Locomotion of black howlers in a habitat with discontinuous canopy. Folia Primatologica, 64, 5561.Google Scholar
Bidder, O. R., Campbell, H. A., Gómez-Laich, A., Urgé, P., Walker, J., Cai, Y., & Wilson, R. P. (2014). Love thy neighbour: Automatic animal behavioural classification of acceleration data using the k-nearest neighbour algorithm. PLoS One, 9(2), e88609.CrossRefGoogle ScholarPubMed
Bieber, C., Lebl, K., Stalder, G., Geiser, F., & Ruf, T. (2014). Body mass dependent use of hibernation: Why not prolong the active season, if they can?. Functional Ecology, 28(1), 167177.Google Scholar
Biegert, J. (1963). The evaluation of characteristics of the skull, hands and feet for primate taxonomy. In Washburn, S. L. (Ed.), Classification and human evolution (Vol. 37, pp. 116145). Chicago, IL: Aldine de Gruyter.Google Scholar
Bininda-Emonds, O. R. P., Cardillo, M., Jones, K. E., MacPhee, R. D.E, Beck, R. M. D., Grenyer, R., …, Purvis, A. (2007). The delayed rise of present-day mammals. Nature, 446, 507512.CrossRefGoogle ScholarPubMed
Bird, D. J., Amirkhanian, A., Pang, B., & Van Valkenburgh, B. (2014). Quantifying the cribriform plate: Influences of allometry, function, and phylogeny in Carnivora. The Anatomical Record, 297, 20802092.CrossRefGoogle ScholarPubMed
Bird, D. J., Murphy, W. J., Fox-Rosales, L., Hamid, I., Eagle, R. A., & Van Valkenburgh, B. (2018). Olfaction written in bone: Cribriform plate size parallels olfactory receptor gene repertoires in Mammalia. Proceedings of the Royal Society B: Biological Sciences, 285(1874), 20180100.CrossRefGoogle ScholarPubMed
BirdLife International (2018). Gracula robusta. The IUCN Red List of Threatened Species 2018: e.T103878817A129935549.Google Scholar
Bishop, A. (1962). Control of the hand in lower primates. Annals of the New York Academy of Sciences, 102(2), 316337.CrossRefGoogle ScholarPubMed
Blair, M. E., Sterling, E. J., & Hurley, M. M. (2011). Taxonomy and conservation of Vietnam’s primates: A review. American Journal of Primatology, 73(11), 10931106.CrossRefGoogle ScholarPubMed
Blair, M. E., Le, M. D., Sethi, G., Thach, H. M., Nguyen, V. T. H., Amato, G., …, Sterling, E. J. (2017). The importance of an interdisciplinary research approach to inform wildlife trade management in Southeast Asia. BioScience, 67(11), 9951003.Google Scholar
Blanchard, M. L., Furnell, S., Sellers, W. I., & Crompton, R. H. (2015). Locomotor flexibility in Lepilemur explained by habitat and biomechanics. American Journal of Physical Anthropology, 156(1), 5866.CrossRefGoogle ScholarPubMed
Blanco, M. B., Dausmann, K. H., Faherty, S. L., & Yoder, A. D. (2018). Tropical heterothermy is ‘cool’: The expression of daily torpor and hibernation in primates. Evolutionary Anthropology: Issues, News, and Reviews, 27(4), 147161.Google Scholar
Bloch, J. I., & Boyer, D. M. (2002). Grasping primate origins. Science, 298(5598), 16061610.CrossRefGoogle ScholarPubMed
Bloch, J. I., Fisher, D. C., Gingerich, P. D., Gunnell, G. F., Simons, E. L., & Uhen, M. D. (1997). Cladistic analysis and anthropoid origins. Science, 278(5346), 21342136.Google Scholar
Boddaert, P. (1785). Elenchus Animalium, Volumen 1. Sistens Quadrupedia huc usque nota, lorumque varietates. Rotterdam: C.R. Hake.Google Scholar
Boettcher, M., Leonard, K., Dickinson, E., Herrel, A., & Hartstone-Rose, A. (2019). Extraordinary grip strength and specialized myology in the hyper-derived hand of Perodicticus potto? Journal of Anatomy. DOI: http://doi.org/10.1111/joa.13051.Google Scholar
Boinski, S. H. (1986). The ecology of squirrel monkeys in Costa Rica (unpublished doctoral dissertation). University of Texas at Austin, Austin, TX, USA.Google Scholar
Boinski, S. H. (1987). Birth synchrony in squirrel monkeys (Saimiri oerstedi). Behavioral Ecology and Sociobiology, 21(6), 393400.Google Scholar
Boinski, S. (1988). Sex differences in the foraging behavior of squirrel monkeys in a seasonal habitat. Behavioral Ecology and Sociobiology, 23(3), 177186.CrossRefGoogle Scholar
Bojarski, C., & Bernard, R. T. F. (1988). Comparison of the morphology of the megachiropteran and microchiropteran eye. African Zoology, 23(3), 155160.Google Scholar
Bonds, A. B., Casagrande, V. A., Norton, T. T., & DeBruyn, E. J. (1987). Visual resolution and sensitivity in a nocturnal primate (Galago) measured with visual evoked potentials. Vision Research, 27(6), 845857.Google Scholar
Bonhote, J. L. (1907). On a collection of mammals made by Dr. Vassal in Annam. Proceedings of the Zoological Society of London, 77(1), 311.CrossRefGoogle Scholar
Borgatti, S. P., & Halgin, D. S. (2011). Mapping culture: Freelists, pilesorting, triads and consensus analysis. The Ethnographer’s Toolkit, 3, 160.Google Scholar
Borgatti, S. P., Everett, M. G., & Freeman, L. C. (Eds.). (2002). Ucinet 6 for Windows: Software for Social Network Analysis. Harvard, MA: Analytic Technologies.Google Scholar
Born Free Foundation. (2014). Pet shop primates: An investigation into the sale of non-human primates by licensed pet shops in England. Retrieved 13 September 2019, from www.bornfree.org.uk/storage/media/content/files/Publications/Pet-Shop-Primates.pdf.Google Scholar
Bosman, W. (Ed.). (1705). A new and accurate description of the coast of Guinea, divided into the Gold, the Slave, and the Ivory coasts. London: Sir Alfred Jones.Google Scholar
Bottcher-Law, L. (2001). Habitat design. In Fitch-Snyder, H. & Schulze, H. (Eds.), Management of lorises in captivity, a husbandry manual for Asian lorisines (Nycticebus & Loris ssp.) (pp. 7273). San Diego, CA: Zoological Society of San Diego.Google Scholar
Bourlière, F., & Petter-Rousseaux, A. (1953). L’Homeothermie imparfaite de certains prosimiens. Comptes rendus des séances de la Société de biologie et de ses filiales, 147, 15941595.Google Scholar
Bouvier, A. (1879). Guide du naturaliste: Revue bibliographique des sciences naturelles bulletin mensuel. Paris: Librairie des sciences naturelles.Google Scholar
Boyd, D. M., & Ellison, N. B. (2007). Social network sites: Definition, history, and scholarship. Journal of Computer‐Mediated Communication, 13(1), 210230.Google Scholar
Boyer, D. M. (2008). Relief index of second mandibular molars is a correlate of diet among prosimian primates and other euarchontan mammals. Journal of Human Evolution, 55(6), 11181137.CrossRefGoogle ScholarPubMed
Boyer, D. M., & Bloch, J. I. (2008). Evaluating the Mitten-Gliding hypothesis for Paromomyidae and Micromomyidae (Mammalia, ‘Plesiadapiformes’) using comparative functional morphology of new paleogene skeletons. In Sargis, E. J. & Dagosto, M. (Eds.), Mammalian evolutionary morphology (pp. 233-284) New York: Springer.CrossRefGoogle Scholar
Boyer, D. M., Evans, A. R., & Jernvall, J. (2010a). Evidence of dietary differentiation among late Paleocene–early Eocene plesiadapids (Mammalia, Primates). American Journal of Physical Anthropology, 142(2), 194210.Google Scholar
Boyer, D. M., Seiffert, E. R., & Simons, E. L. (2010b). Astragalar morphology of Afradapis, a large adapiform primate from the earliest late Eocene of Egypt. American Journal of Physical Anthropology, 143(3), 383402.Google Scholar
Boyer, D. M., Costeur, L., & Lipman, Y. (2012). Earliest record of Platychoerops (Primates, Plesiadapidae), a new species from Mouras quarry, Mont de Berru, France. American Journal of Physical Anthropology, 149(3), 329346.Google Scholar
Boyer, D. M., Kaufman, S., Gunnell, G. F., Rosenberger, A. L., & Delson, E. (2014). Managing 3D digital data sets of morphology: MorphoSource is a new project-based data archiving and distribution tool. American Journal of Physical Anthropology, 153, 8484.Google Scholar
Boyer, D. M., Gunnell, G. F., Kaufman, S., & McGeary, T. M. (2016). Morphosource: Archiving and sharing 3-d digital specimen data. The Paleontological Society Papers, 22, 157181.Google Scholar
Bradley, J. C., Waliczek, T. M., & Zajicek, J. M. (1999). Relationship between environmental knowledge and environmental attitude of high school students. The Journal of Environmental Education, 30(3), 1721.Google Scholar
Brandon-Jones, D., Eudey, A. A., Geissmann, T., Groves, C. P., Melnick, D. J., Morales, J. C., …, & Stewart, C. B. (2004). Asian primate classification. International Journal of Primatology, 25(1), 97164.CrossRefGoogle Scholar
Braun-Courville, D. K., & Rojas, M. (2009). Exposure to sexually explicit web sites and adolescent sexual attitudes and behaviors. Journal of Adolescent Health, 45(2), 156162.CrossRefGoogle ScholarPubMed
Breuer, T., & Mavinga, F. B. (2010). Education for the conservation of great apes and other wildlife in Northern Congo: The importance of nature clubs. American Journal of Primatology, 72(5), 454461.CrossRefGoogle ScholarPubMed
Brockelman, W. Y., & Srikosamatara, S. (1993). Estimation of density of gibbon groups by use of loud songs. American Journal of Primatology, 29(2), 93108.CrossRefGoogle ScholarPubMed
Bronikowski, A. M., & Altmann, J. (1996). Foraging in a variable environment: Weather patterns and the behavioral ecology of baboons. Behavioral Ecology and Sociobiology, 39(1), 1125.Google Scholar
Brown, A. E. (1909). The tuberculin test in monkeys: With notes on the temperature of mammals. Proceedings of the Zoological Society of London, 79(1), 8190.Google Scholar
Brown, D. D., Kays, R., Wikelski, M., Wilson, R., & Klimley, A. P. (2013). Observing the unwatchable through acceleration logging of animal behavior. Animal Biotelemetry, 1(1), 20.CrossRefGoogle Scholar
Brown, G. R. (2001). Using poroximity measures to describe mother–infant relationships. Folia Primatologica, 72(2), 80.Google Scholar
Buck, L., & Axel, R. (1991). A novel multigene family may encode odorant receptors: A molecular basis for odor recognition. Cell, 65(1), 175187.CrossRefGoogle ScholarPubMed
Buckanoff, H. D., Frederick, C., & Murphy, H. W. (2006). Hand‐rearing a potto Perodicticus potto at Franklin Park Zoo, Boston. International Zoo Yearbook, 40(1), 302312.Google Scholar
Buckland, S. T., Anderson, D. R., Burnham, K. P., & Laake, J. L. (1993). Distance sampling: Estimating abundance of biological populations. London: Chapman & Hall.Google Scholar
Buckland, S. T., Anderson, D. R., Burnham, K. P., Laake, J. L., Borchers, D. L., & Thomas, L. (Eds.).(2001). Introduction to distance sampling: Estimating abundance of biological populations (new ed.). Oxford: Oxford University Press.CrossRefGoogle Scholar
Buckland, S. T., Anderson, D. R., Burnham, K. P., Laake, J. L., Borchers, D. L., & Thomas, L. (Eds.). (2004). Advanced distance sampling. Oxford: Oxford University Press.CrossRefGoogle Scholar
Buffon, G. L. L. C. (1765). Histoire Naturelle Générale et Particulière, vol. 13. Paris: L’Imprimerie Royal.Google Scholar
Burke Deleon, V., Smith, T. D., De Leon, V. B., & Smith, T. D. (2014). Mapping the nasal airways: Using histology to enhance CT-based three-dimensional reconstruction in Nycticebus. The Anatomical Record, 297, 21132120.Google Scholar
Burnham, K. P., & Anderson, D. R. (2002). Model selection and multimodel inference: A practical information-theoretic approach (2nd ed.). New York, NY: Springer.Google Scholar
Burnham, K. P., Anderson, D. R., & Laake, J. L. (1980). Estimation of density from line transect sampling of biological populations. Wildlife Monographs, 72, 3202.Google Scholar
Burrows, A. M., & Nash, L. T. (2010). Searching for dental signals of exudativory in galagos. In Burrows, A. M. & Nash, L. T. (Eds.), The evolution of exudativory in primates (pp. 211233). New York, NY: Springer.CrossRefGoogle Scholar
Burrows, A. M., Hartstone‐Rose, A., & Nash, L. T. (2015). Exudativory in the Asian loris, Nycticebus: Evolutionary divergence in the toothcomb and M 3. American Journal of Physical Anthropology, 158(4), 663672.Google Scholar
Burrows, A. M., Nash, L. T., Hartstone-Rose, A., Silcox, M. T., López-Torres, S., & Selig, K. R. (2019a). Dental signatures for exudativory in living Primates, with comparisons to other gouging mammals. The Anatomical Record. DOI: http://doi.org/10.1002/ar.24048.CrossRefGoogle Scholar
Burrows, A. M., Hartstone-Rose, A., Nash, L. T., Silcox, M. T., Selig, K. R., & López-Torres, S. (2019b). The uncertainty of the potto and exudate-feeding in Lorisidae. American Journal of Physical Anthropology, 168 (S67), 3536.Google Scholar
Bush, E. R., Baker, S. E., & Macdonald, D. W. (2014). Global trade in exotic pets 2006–2012. Conservation Biology, 28(3), 663676.Google Scholar
Bušina, T., Pasaribu, N. & Kouba, M. (2018). Ongoing illicit trade of Sumatran Laughingthrush Garrulax bicolor: one-year market monitoring in Medan, North Sumatra. Kukila 21, 2734.Google Scholar
Butynski, T. M., & de Jong, Y. A. (2004). Natural history of the Somali lesser galago (Galago gallarum). Journal of East African Natural History, 93(1), 2339.CrossRefGoogle Scholar
Butynski, T. M., & de Jong, Y. A. (2007). Distribution of the potto Perodicticus potto (Primates: Lorisidae) in eastern Africa, with a description of a new subspecies from Mount Kenya. Journal of East African Natural History, 96(2), 113148.Google Scholar
Butynski, T. M., & de Jong, Y. A. (2017). The Mount Kenya potto is a subspecies of the eastern potto Perodicticus ibeanus. Primate Conservation, 31, 4952.Google Scholar
Butynski, T. M., Ehardt, C. L., & Struhsaker, T. T. (1998). Notes on two dwarf galagos (Galagoides udzungwensis and Galagoides orinus) in the Udzungwa Mountains, Tanzania. Primate Conservation, 18, 6975.Google Scholar
Butynski, T. M., de Jong, Y. A., Perkin, A. W., Bearder, S. K., & Honess, P. E. (2006). Taxonomy, distribution, and conservation status of three species of dwarf galagos (Galagoides) in Eastern Africa. Primate Conservation, 83(18), 6380.CrossRefGoogle Scholar
Byrnes, G., & Jayne, B. C. (2012). The effects of three-dimensional gap orientation on bridging performance and behavior of brown tree snakes (Boiga irregularis). Journal of Experimental Biology, 215(15), 26112620.CrossRefGoogle ScholarPubMed
Cabana, F. (2016). Using feeding ecology to influence captive Slow Loris (Nycticebus spp.) nutrition and husbandry (unpublished doctoral dissertation). Oxford Brookes University, Oxford, UK.Google Scholar
Cabana, F., & Nekaris, K. A. I. (2015). Diets high in fruits and low in gum exudates promote the occurrence and development of dental disease in pygmy slow loris (Nycticebus pygmaeus). Zoo Biology, 34(6), 547553.Google Scholar
Cabana, F., & Plowman, A. (2014). Pygmy slow loris Nycticebus pygmaeus: Natural diet replication in captivity. Endangered Species Research, 23(3), 197204.Google Scholar
Cabana, F., Dierenfeld, E., Wirdateti, W., Donati, G., & Nekaris, K. A. I. (2017). The seasonal feeding ecology of the Javan slow loris (Nycticebus javanicus). American Journal of Physical Anthropology, 162(4), 768781.Google Scholar
Cabana, F., Dierenfeld, E. S., Donati, G., & Nekaris, K. A. I. (2018a). Exploiting a readily available but hard to digest resource: A review of exudativorous mammals identified thus far and how they cope in captivity. Integrative Zoology, 13(1), 94111.Google Scholar
Cabana, F., Jasmi, R., & Maguire, R. (2018b). Great ape nutrition: Low‐sugar and high‐fibre diets can lead to increased natural behaviours, decreased regurgitation and reingestion, and reversal of prediabetes. International Zoo Yearbook, 52(1), 4861.Google Scholar
Cabrera, A. N. G. E. L. (1908). Sobre los loris, y en especial sobre la forma filipina. Boletın De La Real Sociedad Espanola De Historia Natural. Madrid, 8, 135139.Google Scholar
Caine, N. G. (1990). Unrecognized anti-predator behaviour can bias observational data. Animal Behaviour, 39, 195197.Google Scholar
Caine, N. G., & Mundy, N. I. (2000). Demonstration of a foraging advantage for trichromatic marmosets (Callithrix geoffroyi) dependent on food colour. Proceedings of the Royal Society B: Biological Sciences, 267(1442), 439444.CrossRefGoogle ScholarPubMed
Calisi, R. M., & Bentley, G. E. (2009). Lab and field experiments: Are they the same animal?. Hormones and Behavior, 56(1), 110.Google Scholar
Campera, M. (2018). Ecological flexibility and conservation of Fleurette’s sportive lemur, Lepilemur fleuretae, in the lowland rainforest of Ampasy, Tsitongambarika Protected Area (unpublished doctoral dissertation). Oxford Brookes University, Oxford, UK.Google Scholar
Camera, M., Balestri, M., Chimienti, M., Nijman, V., Nekaris, K.A.I., & Donati, G. (2019). Temporal niche separation between the two ecologically similar nocturnal primates Avahi meridionalis and Lepilemur fleuretae. Behavioral Ecology and Sociobiology, 73, 55. https://doi.org/10.1007/s00265-019-2664-1.Google Scholar
Cant, J. G. (1992). Positional behavior and body size of arboreal primates: A theoretical framework for field studies and an illustration of its application. American Journal of Physical Anthropology, 88(3), 273283.Google Scholar
Giang, Cao Thi Huong. (2015). Conservation of the slow loris: Using molecular approaches to study genetic diversity of populations in Vietnam (unpublished undergraduate honors thesis). Vietnam National University, Hanoi University of Science, Hanoi, Vietnam.Google Scholar
Cao, Y., Roy, S., Sachdev, R. N., & Heck, D. H. (2012). Dynamic correlation between whisking and breathing rhythms in mice. Journal of Neuroscience, 32(5), 16531659.Google Scholar
Carpaneto, G., & Germi, F. (1989). The mammals in the zoological culture of the Mbuti Pygmies in north-eastern Zaire/I mammiferi nella cultura zoologica dei Pigmei Mbuti nello Zaire nord-orientale. Hystrix, the Italian Journal of Mammalogy, 1(1), 183.Google Scholar
Cartmill, M. (1970). The orbits of arboreal mammals: A reassessment of the arboreal theory of primate evolution (unpublished doctoral dissertation). University of Chicago, Chicago, IL, USA.Google Scholar
Cartmill, M. (1972). Arboreal adaptations and the origin of the order primates. In Tuttle, R. H. (Ed.), The functional and evolutionary biology of primates (pp. 97122). New York, NY: Aldine de Gruyter.Google Scholar
Cartmill, M. (1974). Rethinking primate origins. Science, 184(4135), 436443.Google Scholar
Cartmill, M. (1975). Strepsirhine basicranial structures and the affinities of the Cheirogaleidae. In Luckett, W. & Szalay, F. (Eds.), Phylogeny of the primates: A multidisciplinary approach (pp. 313354). New York, NY: Plenum Press.Google Scholar
Cartmill, M. (1980). Morphology, function, and evolution of the anthropoid postorbital septum. In: Ciochon, R. L. & Chiarelli, A. B. (Eds.), Evolutionary biology of the New World monkeys and continental drift (pp. 243274). New York, NY: Plenum Press.Google Scholar
Cartmill, M. (1992). New views on primate origins. Evolutionary Anthropology: Issues, News, and Reviews, 1(3), 105111.CrossRefGoogle Scholar
Cartmill, M. (2012). Primate origins, human origins, and the end of higher taxa. Evolutionary Anthropology: Issues, News, and Reviews, 21(6), 208220.CrossRefGoogle ScholarPubMed
Cartmill, M., & Milton, K. (1977). The lorisiform wrist joint and the evolution of ‘brachiating’ adaptations in the Hominoidea. American Journal of Physical Anthropology, 47(2), 249272.Google Scholar
Case, L. P., Daristotle, L., Hayek, M. G., & Raasch, M. F. (Eds.). (2013). Canine and feline mutrition: A resource for companion animal professionals (3rd ed.). Maryland Heights, MO: Mosby Inc.Google Scholar
Casewell, N. R., Wüster, W., Vonk, F. J., Harrison, R. A., & Fry, B. G. (2013). Complex cocktails: The evolutionary novelty of venoms. Trends in Ecology & Evolution, 28(4), 219229.CrossRefGoogle ScholarPubMed
Catania, K. C. (1999). A nose that looks like a hand and acts like an eye: The unusual mechanosensory system of the star-nosed mole. Journal of Comparative Physiology A, 185(4), 367372.Google Scholar
Catania, K. C., & Henry, E. C. (2006). Touching on somatosensory specializations in mammals. Current Opinion in Neurobiology, 16(4), 467473.CrossRefGoogle ScholarPubMed
Caton, J. M., Lawes, M., & Cunningham, C. (2000). Digestive strategy of the south-east African lesser bushbaby, Galago moholi. Comparative Biochemistry and Physiology Part A: Molecular & Integrative Physiology, 127(1), 3948.CrossRefGoogle ScholarPubMed
Caughley, G. (Ed.). (1977). Analysis of Vertebrate Populations. Chichester: Wiley.Google Scholar
Chaimanee, Y., Yamee, C., Marandat, B., & Jaeger, J. J. (2007). First Middle Miocene rodents from the Mae Moh Basin (Thailand): Biochronological and paleoenvironmental implications. Bulletin of Carnegie Museum of Natural History, 204(39), 157164.CrossRefGoogle Scholar
Champion, H. G. & Seth, S. K. (Eds.). (1968). A revised survey of forests types of India. New Delhi: Manager of Publication.Google Scholar
Chan, L. K. (2007). Glenohumeral mobility in primates. Folia Primatologica, 78(1), 118.Google Scholar
Chan, L. K. (2008). The range of passive arm circumduction in primates: Do hominoids really have more mobile shoulders? American Journal of Physical Anthropology, 136(3), 265277.CrossRefGoogle ScholarPubMed
Chandler, D., & Munday, R. (2016). A dictionary of social media. Oxford: Oxford University Press.CrossRefGoogle Scholar
Chaplin, G., Jablonski, N. G., Sussman, R. W., & Kelley, E. A. (2014). The role of piloerection in primate thermoregulation. Folia Primatologica, 85(1), 117.Google Scholar
Chappell, J., Phillips, A. C., van Noordwijk, M. A., Setia, T. M., & Thorpe, S. K. (2015). The ontogeny of gap crossing behaviour in Bornean orangutans (Pongo pygmaeus wurmbii). PLoS One, 10(7), e0130291.Google Scholar
Charles-Dominique, P. (1977a). Ecology and behaviour of nocturnal primates: Prosimians of equatorial West Africa. New York, NY: Columbia University Press.Google Scholar
Charles‐Dominique, P. (1977b). Urine marking and territoriality in Galago alleni (Waterhouse, 1837—Lorisoidea, Primates): A field study by radio‐telemetry. Zeitschrift Für Tierpsychologie, 43(2), 113138.Google Scholar
Charles-Dominique, P. (1978). Solitary and gregarious prosimians: Evolution of social structures in primates. In Chivers, D. J. & Joysey, K. A. (Eds.), Recent advances in primatology (vol. 3, pp.139149). New York, NY: Academic Press.Google Scholar
Charles-Dominique, P., & Bearder, S. K. (1979). Field studies of lorisid behavior: Methodological aspects. In Doyle, G. A. & Martin, R. D. (Eds.), The study of prosimian behavior (pp. 567629). London: Academic Press.CrossRefGoogle Scholar
Charles-Dominique, P., & Martin, R. D. (1970). Evolution of lorises and lemurs. Nature, 227(5255), 257.CrossRefGoogle ScholarPubMed
Charles-Dominique, P., & Petter, J. J. (1980). Ecology and social life of Phaner furcifer. In Charles-Dominique, P., Cooper, H. M., Hladik, A., Hladik, C. M., Pages, E., Pariente, G. F., …, & Petter, J. J. (Eds.), Nocturnal Malagasy primates: Ecology, physiology and behaviour (pp. 7596). New York, NY: Academic Press.Google Scholar
Chatterjee, H. J., Ho, S. Y., Barnes, I., & Groves, C. (2009). Estimating the phylogeny and divergence times of primates using a supermatrix approach. BMC Evolutionary Biology, 9(1), 259.CrossRefGoogle ScholarPubMed
Chatterjee, S., & Scotese, C. R. (1999). The breakup of Gondwana and the evolution and biogeography of the Indian plate. Proceedings of the Indian National Science Academy Part A, 65(3), 397426.Google Scholar
Chen, J. H., Pan, D., Groves, C., Wang, Y. X., Narushima, E., Fitch-Snyder, H., …, & Fu, Y. X. (2006). Molecular phylogeny of Nycticebus inferred from mitochondrial genes. International Journal of Primatology, 27(4), 11871200.Google Scholar
Chen, M. J. (2014). Investigation on the distribution and home range of Nycticebus pygmaeus in Pingbian Daweishan: With a discussion on the conservation attitude of local communities (unpublished masters dissertation). Yunnan Normal University, Yunnan, China.Google Scholar
Chen, Z., Zhang, Y., Shi, L., Liu, R., & Wang, Y. (1993). Studies on the chromosomes of genus Nycticebus. Primates, 34(1), 4753.Google Scholar
Cheng, J. C. H., & Monroe, M. C. (2012). Connection to nature: Children’s affective attitude toward nature. Environment and Behavior, 44(1), 3149.Google Scholar
Cheyne, S. M. (2006). Wildlife reintroduction: Considerations of habitat quality at the release site. BMC Ecology, 6(1), 5.Google Scholar
Chiarelli, A. B. (1972). Taxonomic atlas of living primates. London: Academic Press.Google Scholar
Chimienti, M., Cornulier, T., Owen, E., Bolton, M., Davies, I. M., Travis, J. M., & Scott, B. E. (2016). The use of an unsupervised learning approach for characterizing latent behaviors in accelerometer data. Ecology and Evolution, 6(3), 727741.Google Scholar
Chivers, D. J., & Hladik, C. M. (1980). Morphology of the gastrointestinal tract in primates: Comparisons with other mammals in relation to diet. Journal of Morphology, 166(3), 337386.Google Scholar
Chng, S. C. L, Eaton, J. A., Krishnasamy, K, Shepherd, C. R., & Nijman, V. (2015). In the market for extinction: An inventory of Jakarta’s bird markets. Kuala Lumpur: TRAFFIC.Google Scholar
Choi, L., Liu, Z., Matthews, C. E., & Buchowski, M. S. (2011). Validation of accelerometer wear and nonwear time classification algorithm. Medicine and Science in Sports and Exercise, 43(2), 357.Google Scholar
Choudhury, J. K., Biswas, S. R., Islam, S. M., Rahman, O., & Uddin, S. N. (2004). Biodiversity of Shatchari Reserved Forest, Habiganj. Dhaka: IUCN Bangladesh Country Office.Google Scholar
CITES. (2013). A guide to using the CITES Trade Database, Version 8. Retrieved 20 July 2018, from https://trade.cites.org/cites_trade_guidelines/en-CITES_Trade_Database_Guide.pdf.Google Scholar
CITES. (2019). Convention on International Trade in Endangered Species of Wild Fauna and Flora. Retrieved 31 January 2019, from www.cites.org.Google Scholar
CITES Trade Database. (2018). CITES Trade Database. Retrieved 18 September 2016, from http://trade.cites.org/#.Google Scholar
Clark, A. B. (1985). Sociality in a nocturnal ‘solitary’ prosimian: Galago crassicaudatus. International Journal of Primatology, 6, 581600.Google Scholar
Clemens, W. A. (2004). Purgatorius (Plesiadapiformes, Primates?, Mammalia), a Paleocene immigrant into northeastern Montana: Stratigraphic occurrences and incisor proportions. Bulletin of Carnegie Museum of Natural History, 2004(36), 314.CrossRefGoogle Scholar
Clemmons, G., Granatosky, M. C., Schmitt, D., & Hanna, J. B. (2018). Kinematic strategies are scale-dependent during vertical climbing in primates. American Journal of Physical Anthropology, 165, 5051.Google Scholar
Clutton-Brock, T. H. (1985). Size, sexual dimorphism, and polygyny in primates. In Jungers, W. L. (Ed.), Size and scaling in primate biology (pp. 5160). Boston, MA: Springer.CrossRefGoogle Scholar
Cohen, E. (2009). The wild and the humanized: Animals in Thai tourism. Anatolia, 20(1), 100118.Google Scholar
Coimbra, J. P., Kaswera-Kyamakya, C., Gilissen, E., Manger, P. R., & Collin, S. P. (2016). The topographic organization of retinal ganglion cell density and spatial resolving power in an unusual arboreal and slow-moving strepsirhine primate, the potto (Perodicticus potto). Brain, Behavior and Evolution, 87(1), 418.Google Scholar
Coimbra-Filho, A. F., & Mittermeier, R. A. (1977). Tree-gouging, exudate-eating, and the ‘short-tusked’ condition in Callithrix and Cebuella. In Kleiman, D. (Ed.), The biology and conservation of the Callitrichidae (pp. 105115). Washington, DC: Smithsonian Institution Press.Google Scholar
Coimbra-Filho, A. F., Rocha, N. C., & Pissinatti, A. (1980). Morfofisiologia de ceco e sua correlação com o tipo odentológico em Callitrichidae (Platyrrhini, Primates). Revista Brasileira de Biologia, 40, 177185.Google Scholar
Coleman, M. N., & Ross, C. F. (2004). Primate auditory diversity and its influence on hearing performance. The Anatomical Record Part A: Discoveries in Molecular, Cellular, and Evolutionary Biology, 281(1), 11231137.CrossRefGoogle ScholarPubMed
Colicev, A., Malshe, A., Pauwels, K., & O’Connor, P. (2018). Improving consumer mindset metrics and shareholder value through social media: The different roles of owned and earned media. Journal of Marketing, 82(1), 3756.Google Scholar
Congdon, K. A., & Ravosa, M. J. (2016). Get a grip: Substrate orientation and digital grasping pressures in Strepsirrhines. Folia Primatologica, 87(4), 224243.Google Scholar
Conroy, G. C., Senut, B., Gommery, D., Pickford, M., & Mein, P. (1992). Brief communication: New primate remains from the Miocene of Namibia, southern Africa. American Journal of Physical Anthropology, 99, 487492.Google Scholar
Constantino, P. J., & Wright, B. W. (2009). The importance of fallback foods in primate ecology and evolution. American Journal of Physical Anthropology, 140(4), 599602.Google Scholar
Cooke, S. J., Thorstad, E. B., & Hinch, S. G. (2004). Activity and energetics of free‐swimming fish: Insights from electromyogram telemetry. Fish and Fisheries, 5(1), 2152.CrossRefGoogle Scholar
Corlett, R. T. (2007). The impact of hunting on the mammalian fauna of tropical Asian forests. Biotropica, 39(3), 292303.Google Scholar
Cornwallis, C. K., West, S. A., Davis, K. E., & Griffin, A. S. (2010). Promiscuity and the evolutionary transition to complex societies. Nature, 466(7309), 969.Google Scholar
Costantini, D., Sebastiano, M., Goossens, B., & Stark, D. J. (2017). Jumping in the night: An investigation of the leaping activity of the western tarsier (Cephalopachus bancanus borneanus) using accelerometers. Folia Primatologica, 88(1), 4656.CrossRefGoogle ScholarPubMed
Coultas, D. S. (2002). Bioacoustic analysis of the loud call of two species of slender loris (Loris tardigradus and Loris lydekkerianus nordicus) from Sri Lanka (unpublished Master’s dissertation). Oxford Brookes University, Oxford, UK.Google Scholar
Covert, H. H. (2002). The earliest fossil primates and the evolution of prosimians: Introduction. In Hartwig, W. C. (Ed.), The primate fossil record (pp. 1320). Cambridge: Cambridge University Press.Google Scholar
Covey, R., & McGraw, W. S. (2014). Monkeys in a West African bushmeat market: Implications for cercopithecid conservation in eastern Liberia. Tropical Conservation Science, 7(1), 115125.Google Scholar
Cowgill, U. M. (1969). Some observations on the prosimian Perodicticus potto. Folia Primatologica, 11(1–2), 144150.CrossRefGoogle ScholarPubMed
Cowgill, U. M. (1974). Co-operative behaviour in Perodicticus. In Martin, R. D., Doyle, G. A., & Walker, A. C. (Eds.), Prosimian biology (pp. 261272). London: Duckworth.Google Scholar
Cowgill, U. M., & States, S. J. (1982). Lactose intolerance in captive nocturnal prosimians (Perodicticus potto): A twenty-one year record. Primates, 23(4), 598604.Google Scholar
Cowgill, U. M., & Zeman, L. B. (1980). Life span in captive nocturnal prosimians (Perodicticus potto) with reproductive and mortality records. Primates, 21(3), 437439.CrossRefGoogle Scholar
Cowgill, U. M., States, S. J., & States, K. J. (1989). A twenty‐five‐year chronicle of a group of captive nocturnal prosimians (Perodicticus potto). Mammal Review, 19(2), 8389.Google Scholar
Crandall, L., (1964). The management of wild animals in captivity. Chicago, IL: University of Chicago Press.Google Scholar
Craven, B. A., Paterson, E. G., & Settles, G. S. (2009). The fluid dynamics of canine olfaction: Unique nasal airflow patterns as an explanation of macrosmia. Journal of the Royal Society Interface, 7(47), 933943.CrossRefGoogle ScholarPubMed
Creamer, J., & Philips, T. (Eds.). (2005). My mate’s a primate: Evaluating our relationship and behaviour towards our fellow primates. London: Animal Defenders International.Google Scholar
Crofoot, M. C., Lambert, T. D., Kays, R., & Wikelski, M. C. (2010). Does watching a monkey change its behaviour? Quantifying observer effects in habituated wild primates using automated radiotelemetryAnimal Behaviour80(3), 475480.Google Scholar
Crompton, R. H. (1980). A leap in the dark: Locomotor behaviour and ecology in Galago senegalensis and Galago crassicaudatus (unpublished doctoral dissertation). Harvard University, Cambridge, MA, USA.Google Scholar
Crompton, R. H. (1983). Age differences in locomotion in two subtropical Galaginae. Primates. 24: 241259.Google Scholar
Crompton, R. H. (1984). Foraging, habitat structure, and locomotion in two species of Galago. In Rodman, P. S. & Cant, J. G. H. (Eds.), Adaptations for foraging in non-human primates (pp. 73111). Columbia, NY: Columbia University Press.Google Scholar
Crompton, R. H. (1995). ‘Visual predation,’ habitat structure, and the ancestral primate niche. In Alterman, L., Doyle, G. A., & Izard, M. K. (Eds.), Creatures of the dark: The nocturnal prosimians (pp. 1130). Boston, MA: Springer.Google Scholar
Cronin, D. T., Riaco, C., Linder, J. M., Bergl, R. A., Gonder, M. K., O’Connor, M. P., & Hearn, G. W. (2016). Impact of gun-hunting on monkey species and implications for primate conservation on Bioko Island, Equatorial Guinea. Biological Conservation, 197, 180189.Google Scholar
Crook, J. H., & Gartlan, J. S. (1966). Evolution of primate societies. Nature, 210, 12001203.Google Scholar
Curtin, S. (2009). Wildlife tourism: The intangible, psychological benefits of human–wildlife encounters. Current Issues in Tourism, 12(5–6), 451474.CrossRefGoogle Scholar
Curtis, D. J. (1995). Functional anatomy of the trunk musculature in the slow loris (Nycticebus coucang). American Journal of Physical Anthropology, 97(4), 367379.CrossRefGoogle ScholarPubMed
Curtis, D. J., Howden, M., Curtis, F., McColm, I., Scrine, J., Blomfield, T., …, & Ryan, T. (2013). Drama and environment: Joining forces to engage children and young people in environmental education. Australian Journal of Environmental Education, 29(2), 182201.Google Scholar
Dacier, A., de Luna, A. G., Fernández‐Duque, E., & Di Fiore, A. (2011). Estimating population density of Amazonian titi monkeys (Callicebus discolor) via playback point counts. Biotropica, 43(2), 135140.CrossRefGoogle Scholar
Dagosto, M. (1988). Implications of postcranial evidence for the origin of euprimates. Journal of Human Evolution, 17(1–2), 3556.Google Scholar
Dagosto, M., & Yamashita, N. (1998). Effect of habitat structure on positional behavior and support use in three species of lemur. Primates, 39(4), 459472.CrossRefGoogle Scholar
Dalecki, D., Child, S. Z., Raeman, C. H., & Carstensen, E. L. (1995). Tactile perception of ultrasound. Journal of the Acoustical Society of America, 97(5), 31653170.Google Scholar
Van Tien, Dao (1960). Sur une nouvelle espece de Nycticebus au Vuetnam. Zoologischer Anzeiger, 164, 240243.Google Scholar
D’Aout, K. & Vereecke, E. E. (Eds.). (2011). Primate locomotion: Linking field and laboratory research. New York, NY: Springer.Google Scholar
Dartnall, H. J. A., Arden, G. B., Ikeda, H., Luck, C. P., Rosenberg, M. E., Pedler, C. M. H., & Tansley, K. (1965). Anatomical, electrophysiological and pigmentary aspects of vision in the bush baby: An interpretative study. Vision Research, 5(8–9), 399424.Google Scholar
Das, N., Biswas, J., Das, J., Ray, P. C., Sangma, A., & Bhattacharjee, P. C. (2009). Status of Bengal slow loris Nycticebus bengalensis (Primates: Lorisidae) in gibbon wildlife sanctuary, Assam, India. Journal of Threatened Taxa, 1, 558561.Google Scholar
Das, S., Dutta, S., Mangalam, M., Verma, R. K., Rath, S., Kumara, H. N., & Singh, M. (2011). Prioritizing remnant forests for the conservation of Mysore slender lorises (Loris lyddekerianus lyddekerianus) in Karnataka, India through estimation of population density. International Journal of Primatology, 32(5), 1153.CrossRefGoogle Scholar
Das, N., Nekaris, K. A. I., & Bhattacharjee, P. C. (2014). Medicinal plant exudativory by the Bengal slow loris Nycticebus bengalensis. Endangered Species Research, 23(2), 149157.Google Scholar
Das, N., Biswas, J., Bhattacharya, K., & Nekaris, K. A. I. (2016). Observations on the Bengal slow loris Nycticebus bengalensis in Pakke Tiger Reserve, Arunachal Pradesh, India. Asian Primates Journal, 6(1), 2732.Google Scholar
Daschbach, N. J., Schein, M. W., & Haines, D. E. (1983). Cage-size effects on locomotor, grooming and agonistic behaviours of the slow loris, Nycticebus coucang (Primates, Lorisidae). Applied Animal Ethology, 9(3–4), 317330.CrossRefGoogle Scholar
Daugherty, C. H., Cree, A., Hay, J. M., & Thompson, M. B. (1990). Neglected taxonomy and continuing extinctions of tuatara (Sphenodon). Nature, 347(6289), 177.Google Scholar
Dausmann, K. H. (2005). Measuring body temperature in the field: Evaluation of external vs. implanted transmitters in a small mammal. Journal of Thermal Biology, 30(3), 195202.CrossRefGoogle Scholar
Dausmann, K. H. (2008). Hypometabolism in primates: Torpor and hibernation. In Lovegrove, B. G. & McKechnie, A. E. (Eds.), Hypometabolism in animals: Torpor, hibernation and cryobiology (pp. 327336). Pietermaritzburg: Interpak Books.Google Scholar
Dausmann, K. H. (2014). Flexible patterns in energy savings: Heterothermy in primates. Journal of Zoology, 292(2), 101111.CrossRefGoogle Scholar
Davies, B. (2005). Black market: Inside the endangered species trade in Asia. San Rafael, CA: Earth Aware Editions.Google Scholar
de Boer, L. E. (1973). Cytotaxonomy of the Lorisoidea (primates: Prosimii): I. Chromosome studies and karyological relationships in the Galagidae. Genetica, 44(2), 155193.Google Scholar
de Waal, F. B., & Gavrilets, S. (2013). Monogamy with a purpose. Proceedings of the National Academy of Sciences, 110(38), 1516715168.Google Scholar
De Winton, W. E. (1902). VIII: Notices of two new species of Potto from the French Congo territory. Journal of Natural History, 9(49), 4749.Google Scholar
Deblauwe, I., & Janssens, G. P. (2008). New insights in insect prey choice by chimpanzees and gorillas in southeast Cameroon: The role of nutritional value. American Journal of Physical Anthropology, 135(1), 4255.CrossRefGoogle ScholarPubMed
Debyser, I. W. J. (1995). Prosimian juvenile mortality in zoos and primate centers. International Journal of Primatology, 16(6), 889907.Google Scholar
Dehnhardt, G., & Dücker, G. (1996). Tactual discrimination of size and shape by a California sea lion (Zalophus californianus). Animal Learning & Behavior, 24(4), 366374.Google Scholar
Dehnhardt, G., & Kaminski, A. (1995). Sensitivity of the mystacial vibrissae of harbour seals (Phoca vitulina) for size differences of actively touched objects. Journal of Experimental Biology, 198(11), 23172323.Google Scholar
delBarco‐Trillo, J., Burkert, B. A., Goodwin, T. E., & Drea, C. M. (2011). Night and day: The comparative study of strepsirrhine primates reveals socioecological and phylogenetic patterns in olfactory signals. Journal of Evolutionary Biology, 24(1), 8298.Google Scholar
DeMenocal, P. (1995). Plio-Pleistocene African climate. Science, 270 (5233), 5359.CrossRefGoogle ScholarPubMed
Demes, B., Jungers, W., & Nieschalk, U. (1990). Size-and speed-related aspects of quadrupedal walking in slender and slow lorises. In Jouffroy, F., Stack, M., & Niemitz, C. (Eds.), Gravity, posture and locomotion in primates (Editrice II) (pp. 175197). Florence: Sedicesimo.Google Scholar
Demes, B., Larson, S. G., Stern, J. T. Jr, Jungers, W. L., Biknevicius, A. R., & Schmitt, D. (1994). The kinetics of primate quadrupedalism: ‘Hindlimb drive’ reconsidered. Journal of Human Evolution, 26(5–6), 353374.CrossRefGoogle Scholar
Demes, B., Fleagle, J. G., & Lemelin, P. (1998). Myological correlates of prosimian leaping. Journal of Human Evolution, 34(4), 385399.CrossRefGoogle ScholarPubMed
Demes, B., Fleagle, J. G., & Jungers, W. L. (1999). Takeoff and landing forces of leaping strepsirhine primates. Journal of Human Evolution, 37(2), 279292.CrossRefGoogle ScholarPubMed
Dene, H., Goodman, M., Prychodko, W., & Moore, W. (1976). Immunodiffusion systematics of the primates. Folia Primatologica, 25(1), 3561.Google Scholar
Denzinger, A., & Schnitzler, H. U. (2013). Bat guilds, a concept to classify the highly diverse foraging and echolocation behaviors of microchiropteran bats. Frontiers in Physiology, 4, 164.CrossRefGoogle ScholarPubMed
Deschênes, M., Moore, J., & Kleinfeld, D. (2012). Sniffing and whisking in rodents. Current Opinion in Neurobiology, 22(2), 243250.Google Scholar
Di Minin, E., Fink, C., Tenkanen, H., & Hiippala, T. (2018). Machine learning for tracking illegal wildlife trade on social media. Nature Ecology & Evolution, 2(3), 406.Google Scholar
Dillman, D. A., Smyth, J. D., & Christian, L. M. (2016). Internet, phone, mail and mixed-mode surveys: The tailored design method. Reis, 154, 161176.Google Scholar
Diogo, R., & Wood, B. (2011). Soft‐tissue anatomy of the primates: Phylogenetic analyses based on the muscles of the head, neck, pectoral region and upper limb, with notes on the evolution of these muscles. Journal of Anatomy, 219(3), 273359.Google Scholar
Dixson, A. F. (1995). Sexual selection and the evolution of copulatory behavior in nocturnal prosimians. In Altermann, L., Doyle, G. A., & Izard, M. K. (Eds.), Creatures of the dark: The nocturnal prosimians (pp. 93118). New York, NY: Plenum Press.Google Scholar
Djagoun, C. A., Akpona, H. A., Mensah, G. A., Nuttman, C., & Sinsin, B. (2013). Wild mammals trade for zootherapeutic and mythic purposes in Benin (West Africa): Capitalizing species involved, provision sources, and implications for conservation. In Alves, R. R. N. & Rosa, I. L. (Eds.), Animals in traditional folk medicine (pp. 367381). Berlin: Springer.CrossRefGoogle Scholar
Dkhissi‐Benyahya, O., Szel, A., Degrip, W. J., & Cooper, H. M. (2001). Short and mid‐wavelength cone distribution in a nocturnal Strepsirrhine primate (Microcebus murinus). Journal of Comparative Neurology, 438(4), 490504.Google Scholar
Dolins, F. L., Jolly, A., Rasamimanana, H., Ratsimbazafy, J., Feistner, A. T., & Ravoavy, F. (2010). Conservation education in Madagascar: Three case studies in the biologically diverse island‐continent. American Journal of Primatology, 72(5), 391406.Google Scholar
Dörfl, J. (1982). The musculature of the mystacial vibrissae of the white mouse. Journal of Anatomy, 135(Pt 1), 147.Google Scholar
Doyle, G. A. (1979). Development of behavior in prosimians with special reference to the lesser bushbaby, Galago senegalensis moholi. In Doyle, G. A. & Martin, R. D. (Eds.), The study of prosimian behavior (pp. 157206). New York, NY: Academic Press.Google Scholar
Duarte, M., Hanna, J., Sanches, E., Liu, Q., & Fragaszy, D. (2012). Kinematics of bipedal locomotion while carrying a load in the arms in bearded capuchin monkeys (Sapajus libidinosus). Journal of Human Evolution, 63(6), 851858.CrossRefGoogle ScholarPubMed
Duckworth, J. W. (1994). Field sightings of the pygmy loris, Nycticebus pygmaeus, in Laos. Folia Primatologica, 63(2), 99101.Google Scholar
Duckworth, J. W. (1997). Mammals in Similajau National Park, Sarawak, in 1995. Sarawak Museum Journal, 51, 171192.Google Scholar
Duckworth, J. W. (1998). The difficulty of estimating population densities of nocturnal forest mammals from transect counts of animals. Journal of Zoology, 246(4), 443486.Google Scholar
Duckworth, J. W., Timmins, R. J., Thewlis, R. C. M., Evans, T. D., & Anderson, G. Q. A. (1994). Field observations of mammals in Laos, 1992–1993. Natural History Bulletin of the Siam Society, 42, 177205.Google Scholar
Duckworth, J. W., Batters, G., Belant, J. L., Bennett, E. L., Brunner, J., Burton, J., …, & Wirth, R. (2012). Why South-east Asia should be the world’s priority for averting imminent species extinctions, and a call to join a developing cross-institutional programme to tackle this urgent issue. SAPIENS: Surveys and Perspectives Integrating Environment and Society, 5(2). Retrieved 13 September 2019, from https://journals.openedition.org/sapiens/1327.Google Scholar
Ducrocq, S., Manthi, F. K., & Lihoreau, F. (2011). First record of a parapithecid primate from the Oligocene of Kenya. Journal of Human Evolution, 61(3), 327331.Google Scholar
Dumont, E. R. (1997). Cranial shape in fruit, nectar, and exudate feeders: Implications for interpreting the fossil record. American Journal of Physical Anthropology, 102(2), 187202.3.0.CO;2-W>CrossRefGoogle ScholarPubMed
Dunbar, R. I. (2009). The social brain hypothesis and its implications for social evolution. Annals of Human Biology, 36(5), 562572.Google Scholar
Dunn, R. H., Rose, K. D., Rana, R. S., Kumar, K., Sahni, A., & Smith, T. (2016). New euprimate postcrania from the early Eocene of Gujarat, India, and the strepsirrhine–haplorhine divergence. Journal of Human Evolution, 99, 2551.CrossRefGoogle ScholarPubMed
Dykyj, D. (1980). Locomotion of the slow loris in a designed substrate context. American Journal of Physical Anthropology, 52(4), 577586.Google Scholar
Eaglen, R. H. (1980). Toothcomb homology and toothcomb function in extant strepsirrhines. International Journal of Primatology, 1, 276–87.Google Scholar
Eaglen, R. H. (1986). Morphometrics of the anterior dentition in strepsirhine primates. American Journal of Physical Anthropology, 71(2), 185201.CrossRefGoogle ScholarPubMed
Eaton, J. A., Shepherd, C. R., Rheindt, F. E., Harris, J. B. C., van Balen, B., Wilcove, D. S., Collar, N. J. (2015). Trade-driven extinctions and near-extinctions of avian taxa in Sundaic Indonesia. Forktail, 31, 1-12.Google Scholar
Ebara, S., Kumamoto, K., Matsuura, T., Mazurkiewicz, J. E., & Rice, F. L. (2002). Similarities and differences in the innervation of mystacial vibrissal follicle–sinus complexes in the rat and cat: A confocal microscopic study. Journal of Comparative Neurology, 449(2), 103119.Google Scholar
eBizMBA Guide. (2018). Top 15 most popular social networking sites. Retrieved 13 September 2019, from www.ebizmba.com/articles/social-networking-websites.Google Scholar
Ehrlich, A., & Macbride, L. (1989). Mother–infant interactions in captive slow lorises (Nycticebus coucang). American Journal of Primatology, 19(4), 217228.Google Scholar
Ehrlich, A., & Musicant, A. (1977). Social and individual behaviors in captive slow lorises. Behaviour, 60, 195220.Google Scholar
Eisentraut, M. (1956a). Der Winterschlaf mit seinen oekologischen und physiologischen Begleiterscheinungen. Jena: G. Fischer.Google Scholar
Eisentraut, M. (1956b). Temperaturschwankungen bei niederen Säugetieren. Z Säugetierk, 21, 4952.Google Scholar
Eisentraut, M. (1960). Heat regulation in primitive mammals and in tropical species. Bulletin of the Museum of Comparative Zoology at Harvard College, 124, 3143.Google Scholar
Eisentraut, M. (1961). Beobachtungen über den Wärmehaushalt bei Halbaffen. Biologisches Zentralblatt, 80, 319325.Google Scholar
Eiting, T. P., Smith, T. D., Perot, J. B., & Dumont, E. R. (2014). The role of the olfactory recess in olfactory airflow. Journal of Experimental Biology, 217(10), 17991803.Google Scholar
Eklöf, J., Šuba, J., Petersons, G., & Rydell, J. (2014). Visual acuity and eye size in five European bat species in relation to foraging and migration strategies. Envrionmental and Experimental Biology, 12, 16.Google Scholar
Elangovan, V., & Marimuthu, G. (2001). Effect of moonlight on the foraging behaviour of a megachiropteran bat Cynopterus sphinx. Journal of Zoology, 253(3), 347350.Google Scholar
Ellingson, L. D., Schwabacher, I. J., Kim, Y., Welk, G. J., & Cook, D. B. (2016). Validity of an integrative method for processing physical activity data. Medicine and Science in Sports and Exercise, 48(8), 16291638.CrossRefGoogle ScholarPubMed
Elliot, O., & Elliot, M. (1967). Field notes on the slow loris in Malaya. Journal of Mammalogy, 48(3), 497498.CrossRefGoogle Scholar
Elvidge, A. M. (2013). Captive breeding programmes for nocturnal prosimians (unpublished doctoral dissertation). Oxford Brookes University, Oxford, UK.Google Scholar
Endler, J. A. (1993). The color of light in forests and its implications. Ecological Monographs, 63(1), 127.CrossRefGoogle Scholar
Engler, M., & Parry-Jones, R. (2007). Opportunity or threat: The role of the European Union in global wildlife trade. Brussels: TRAFFIC Europe.Google Scholar
Enlow, D. H. (1996). Handbook of facial growth (2nd ed.). Philadelphia, PA: Saunders.Google Scholar
Eppley, T. M., Ganzhorn, J. U., & Donati, G. (2015). Cathemerality in a small, folivorous primate: Proximate control of diel activity in Hapalemur meridionalis. Behavioral Ecology and Sociobiology, 69(6), 9911002.CrossRefGoogle Scholar
Epps, J. (1974). Social interaction of Perodicticus potto kept in captivity in Kampala, Uganda. In Martin, R. D., Doyle, G. A., & Walker, A. C. (Eds.), Prosimian biology (pp. 233244). London: Duckworth.Google Scholar
Erkert, H. G. (1989). Characteristics of the circadian activity rhythm in common marmosets (Callithrix j. jacchus). American Journal of Primatology, 17(4), 271286.Google Scholar
Erkert, H. G., & Kappeler, P. M. (2004). Arrived in the light: Diel and seasonal activity patterns in wild Verreaux’s sifakas (Propithecus v. verreauxi; Primates: Indriidae). Behavioral Ecology and Sociobiology, 57(2), 174186.CrossRefGoogle Scholar
Erkert, H. G., Nagel, B., & Stephani, I. (1986). Light and social effects on the free-running circadian activity rhythm in common marmosets (Callithrix jacchus; Primates): Social masking, pseudo-splitting, and relative coordination. Behavioral Ecology and Sociobiology, 18(6), 443452.Google Scholar
ESRC (Economic and Social Research Council). (2015). ESRC framework for research ethics (updated January 2015). Retrieved January 2019, from https://esrc.ukri.org.Google Scholar
Estrada, A., Raboy, B. E., & Oliveira, L. C. (2012). Agroecosystems and primate conservation in the tropics: A review. American Journal of Primatology, 74(8), 696711.CrossRefGoogle ScholarPubMed
Estrada, A., Garber, P. A., Rylands, A. B., Roos, C., Fernández-Duque, E., Di Fiore, A., …, & Rovero, F. (2017). Impending extinction crisis of the world’s primates: Why primates matter. Science Advances, 3(1), e1600946.Google Scholar
Evans, T. (2003). A study of resin-tapping and livelihoods in southern Mondulkiri, Cambodia: With implications for conservation and forest management. Phnom Penh: Wildlife Conservation Society Cambodia Program.Google Scholar
Evans, T. D., Duckworth, J. W., & Timmins, R. J. (2000). Field observations of larger mammals in Laos, 1994–1995. Mammalia, 64(1), 55100.Google Scholar
Fa, J. E., & Garcia, Y. (2001). Commercial bushmeat hunting in the Monte Mitra forests, Equatorial Guinea: Extent and impact. Animal Biodiversity and Conservation, 24(1), 3152.Google Scholar
Fa, J. E., Seymour, S., Dupain, J. E. F., Amin, R., Albrechtsen, L., & Macdonald, D. (2006). Getting to grips with the magnitude of exploitation: Bushmeat in the Cross–Sanaga Rivers region, Nigeria and Cameroon. Biological Conservation, 129(4), 497510.CrossRefGoogle Scholar
Fabre, P. H., Rodrigues, A., & Douzery, E. J. (2009). Patterns of macroevolution among Primates inferred from a supermatrix of mitochondrial and nuclear DNA. Molecular Phylogenetics and Evolution, 53(3), 808.CrossRefGoogle ScholarPubMed
Fairbanks, L. A., & Hinde, K. (2013). Behavioral responses of mothers and infants to variation in maternal condition: Adaptation, compensation, and resilience. In Clancy, K. B. H., Hinde, K., & Rutherford, J. N. (Eds.), Building babies: Primate development in proximate and ultimate perspective (pp. 281302). New York, NY: Springer.CrossRefGoogle Scholar
Fam, S. D., Lee, B. P. Y.-H., & Shekelle, M. (2014). The conservation status of slow lorises Nycticebus spp. in Singapore. Endangered Species Research, 25(1), 6977.Google Scholar
Fan, P., Scott, M. B., Fei, H., & Ma, C. (2013). Locomotion behavior of cao vit gibbon (Nomascus nasutus) living in karst forest in Bangliang Nature Reserve, Guangxi, China. Integrative Zoology, 8(4), 356364.CrossRefGoogle ScholarPubMed
FAO. (2016). State of the world’s forests 2016: Forests and agriculture – land-use challenges and opportunities. Rome: UN Food and Agriculture Organization.Google Scholar
FAWC. (1979). Farm Animal Welfare Council press statement. Retrieved 3 February 2019, from https://webarchive.nationalarchives.gov.uk/20121010012427/www.fawc.org.uk/freedoms.htm.Google Scholar
Fawcett, L. (2002). Children’s wild animal stories: Questioning inter-species bonds. Canadian Journal of Environmental Education, 7(2), 125139.Google Scholar
Feddema, K. (2015). Perception of primates in online content: Investigating the attitudes and understanding of the adolescent student (unpublished Master’s dissertation). Oxford Brookes University, Oxford, UK.Google Scholar
Fedigan, L. M. (1990). Vertebrate predation in Cebus capucinus: Meat eating in a neotropical monkey. Folia Primatologica, 54(3–4), 196205.CrossRefGoogle Scholar
Feldhamer, G. A., Drickamer, L. C., Vessey, S. H., Merritt, J. F., & Krajewski, C. (2007). Mammalogy: Adaptation, diversity, ecology. Baltimore, MD: Johns Hopkins University Press.Google Scholar
Felsenstein, J. (1981). Evolutionary trees from gene frequencies and quantitative characters: Finding maximum likelihood estimates. Evolution, 35(6), 12291242.Google Scholar
Feng, Q., Wang, Y., & Li, C. (1993). Reproduction of slow loris (Nycticebus intermedius) in captivity. Zoological Research, 14(1), 2126.Google Scholar
Feng, Q., Wang, Y., & Li, C. (1994). Progress of studies on intermediate slow loris in captivity. Zhongguo Lingzhanglei Yanjiu Tongxun/Chinese Primate Research and Conservation News, 3(1), 1112.Google Scholar
Feng, Q., Wang, Y., & Li, C. (1996). Postnatal growth and development in the intermediate slow loris (Nycticebus intermedius) in captivity. Zoological Research, 17, 443450.Google Scholar
Fernández-Duque, E., & Erkert, H. G. (2006). Cathemerality and lunar periodicity of activity rhythms in owl monkeys of the Argentinian Chaco. Folia Primatologica, 77(1–2), 123138.Google Scholar
Fernández-Duque, E., De La Iglesia, H., & Erkert, H. G. (2010). Moonstruck primates: Owl monkeys (Aotus) need moonlight for nocturnal activity in their natural environment. PLoS One, 5(9), e12572.Google Scholar
Ferrari, S. F. (1993). Ecological differentiation in the Callitrichidae. In Rylands, A. B. (Ed.), Marmosets and tamarins: Systematics, behaviour, and ecology (pp. 314328). Oxford: Oxford University Press.Google Scholar
Fieberg, J., & Kochanny, C. O. (2005). Quantifying home‐range overlap: The importance of the utilization distribution. The Journal of Wildlife Management, 69(4), 13461359.Google Scholar
Field, A. (2009). Discovering statistics using SPSS (3rd ed.). London: Sage.Google Scholar
Fietz, J. (1999). Monogamy as a rule rather than exception in nocturnal lemurs: The case of the fat‐tailed dwarf lemur, Cheirogaleus medius. Ethology, 105(3), 255272.Google Scholar
Fietz, J., & Dausmann, K. H. (2006). Big is beautiful: Fat storage and hibernation as a strategy to cope with marked seasonality in the fat-tailed dwarf lemurs (Cheirogaleus medius). In Gould, L. & Sauther, M.L (Eds.), Lemurs: Ecology and adaptation (pp. 97110). Boston, MA: Springer.Google Scholar
Finke, M. D. (2003). Gut loading to enhance the nutrient content of insects as food for reptiles: A mathematical approach. Zoo Biology, 22(2), 147162.CrossRefGoogle Scholar
Finke, M. D. (2013). Complete nutrient content of four species of feeder insects. Zoo Biology, 32(1), 2736.Google Scholar
Finley, R. B. (1959). Observation of nocturnal animals by red light. Journal of Mammalogy, 40(4), 591594.CrossRefGoogle Scholar
Finstermeier, K., Zinner, D., Brameier, M., Meyer, M., Kreuz, E., Hofreiter, M., & Roos, C. (2013). A mitogenomic phylogeny of living primates. PLoS One, 8(7), e69504.Google Scholar
Fisher, H. S., Swaisgood, R. R., & Fitch-Snyder, H. (2003a). Odor familiarity and female preferences for males in a threatened primate, the pygmy loris Nycticebus pygmaeus: Applications for genetic management of small populations. Naturwissenschaften, 90(11), 509512.Google Scholar
Fisher, H. S., Swaisgood, R., & Fitch-Snyder, H. (2003b). Countermarking by male pygmy lorises (Nycticebus pygmaeus): Do females use odor cues to select mates with high competitive ability?. Behavioral Ecology and Sociobiology, 53(2), 123130.CrossRefGoogle Scholar
Fitch-Snyder, H. (2000). Reproductive patterns in a breeding colony of pygmy lorises (Nycticebus pygmaeus) measured by behavioral and physiological correlates of gonadal activity (Master’s thesis). San Diego State University, San Diego, CA, USA.Google Scholar
Fitch-Snyder, H. (2014). Status and history of captive loris primates in North America, 25th Congress of the International Primatological Society, Vietnam, 11–16 August 2014.Google Scholar
Fitch-Snyder, H. (2015). Regional studbook for pygmy lorises (Nycticebus pygmaeus). San Diego, CA: San Diego Zoo Global.Google Scholar
Fitch-Snyder, H., & Ehrlich, A. (2003). Mother–infant interactions in slow lorises (Nycticebus bengalensis) and pygmy lorises (Nycticebus pygmaeus). Folia Primatologica, 74(5–6), 259271.Google Scholar
Fitch‐Snyder, H., & Jurke, M. (2003). Reproductive patterns in pygmy lorises (Nycticebus pygmaeus): Behavioral and physiological correlates of gonadal activity. Zoo Biology, 22(1), 1532.Google Scholar
Fitch-Snyder, H., & Perez, L. (1989). Correlation and evaluation of various methods used to monitor estrus in two subspecies of Lorisidae (abstract). American Journal of Primatology, 18, 140.Google Scholar
Fitch-Snyder, H., & Schulze, H. (2001). Management of lorises in captivity: A husbandry manual for Asian loridae (Nycticebus & Loris ssp.). San Diego, CA: Center for Reproduction of Endangered Species, Zoological Society of San Diego.Google Scholar
Fitch-Snyder, H., & Thanh, V. N. (2002). A preliminary survey of lorises (Nycticebus spp.) in northern Vietnam. Asian Primates, 8(2), 16.Google Scholar
Fitch-Snyder, H., Schulze, H., & Streicher, U. (2008). Enclosure design for captive slow and pygmy lorises. In Shekelle, M., Groves, C., Maryanto, I., Schulze, H., & Fitch-Snyder, H. (Eds.), Primates of the oriental night (pp. 123135). Jakarta: Research Center for Biology, Indonesian Institute of Sciences.Google Scholar
Fleagle, J. G. (1999). Primate adaptation and evolution (2nd ed.). New York, NY: Academic Press.Google Scholar
Fleagle, J. G. (2013). Primate adaptation and evolution (3rd ed.). New York, NY: Academic Press.Google Scholar
Fleagle, J. G., & Simons, E. L. (1978). Micropithecus clarki, a small ape from the Miocene of Uganda. American Journal of Physical Anthropology, 49(4), 427440.Google Scholar
Flesher, K. M. (2015). The distribution, habitat use, and conservation status of three Atlantic Forest monkeys (Sapajus xanthosternos, Callicebus melanochir, Callithrix sp.) in an agroforestry/forest mosaic in Southern Bahia, Brazil. International Journal of Primatology, 36(6), 11721197.Google Scholar
Flynn, L. J., & Morgan, M. E. (2005). New lower primates from the Miocene Siwaliks of Pakistan. In Lieberman, D. E., Smith, R. J., & Kelley, J. (Eds.), Interpreting the past: Essays on human, primate, and mammal evolution in honour of David Pilbeam (pp. 81101). Boston, MA: Brill Academic Publishers.CrossRefGoogle Scholar
Fobes, J. L., & King, J. E. (1977). Prosimian sensory capacities. Primates, 18(3), 713730.Google Scholar
Fobes, J. L., & King, J. E. (1982). Vision: The dominant primate modality. In Fobes, J. L. & King, J. E. (Eds.), Primate behavior (pp. 219243). New York, NY: Academic Press.Google Scholar
Fontaneto, D., Tommaseo-Ponzetta, M., Galli, C., Risé, P., Glew, R. H., & Paoletti, M. G. (2011). Differences in fatty acid composition between aquatic and terrestrial insects used as food in human nutrition. Ecology of Food and Nutrition, 50(4), 351367.Google Scholar
Fooden, J. (1991). Eastern limit of distribution of the slow loris, Nycticebus coucang. International Journal of Primatology, 12(3), 287290.Google Scholar
Forster, A. (1933). La ‘pince palmaire’ et la ‘pince plantaire’ de Perodicticus potto. Archives D’anatomie, D’histologie Et D’embryologie, 17, 181246.Google Scholar
Forsythe, E. C., & Ford, S. M. (2011). Craniofacial adaptations to tree‐gouging among marmosets. The Anatomical Record, 294(12), 21312139.CrossRefGoogle ScholarPubMed
Fortes‐Marco, L., Lanuza, E., & Martinez‐Garcia, F. (2013). Of pheromones and kairomones: What receptors mediate innate emotional responses?. The Anatomical Record, 296(9), 13461363.CrossRefGoogle ScholarPubMed
Fowler, J., Cohen, L., & Jarvis, P. (1998). Practical statistics for field biology (2nd ed.). Chichester: Wiley.Google Scholar
Fox, R. C., & Scott, C. S. (2011). A new, early Puercan (earliest Paleocene) species of Purgatorius (Plesiadapiformes, Primates) from Saskatchewan, Canada. Journal of Paleontology, 85(3), 537548.Google Scholar
Fragaszy, D. M. (1986). Time budgets and foraging behavior in wedge-capped capuchins (Cebus olivaceus): Age and sex differences. In Taub, D. M. & King, F. A. (Eds.), Current perspective in primate social dynamics (pp. 159174). New York, NY: Van Nostrand.Google Scholar
Fragaszy, D. M. (1990). Sex and age differences in the organization of behavior in wedge-capped capuchins, Cebus olivaceus. Behavioral Ecology, 1(1), 8194.Google Scholar
Franzen, J. L., Gingerich, P. D., Habersetzer, J., Hurum, J. H., Von Koenigswald, W., & Smith, B. H. (2009). Complete primate skeleton from the middle Eocene of Messel in Germany: Morphology and paleobiology. PLoS One, 4(5), e5723.CrossRefGoogle ScholarPubMed
Frederick, C. (1998). Observations of paternal care in Perodicticus potto at the Cincinnati Zoo and Botanical Garden. Folia Primatologica, 69(5), 312.Google ScholarPubMed
Fredrick, C., & Fernandes, D. (1994). Increased activity in a nocturnal primate through lighting manipulation: The case of the potto Perodicticus potto. International Zoo Yearbook, 33(1), 219228.Google Scholar
Frederick, C., & Fernandes, D. (1996). Behavioral changes in pottos (Perodicticus potto): Effects of naturalizing an exhibit. International Journal of Primatology, 17(3), 389399.Google Scholar
Fujiwara, S. I., Endo, H., & Hutchinson, J. R. (2011). Topsy‐turvy locomotion: Biomechanical specializations of the elbow in suspended quadrupeds reflect inverted gravitational constraints. Journal of Anatomy, 219(2), 176191.CrossRefGoogle ScholarPubMed
Fuller, D., Machemer, R., & Knighton, R. W. (1978). Retinal damage produced by intraocular fiber optic light. American Journal of Ophthalmology, 85(4), 519537.Google Scholar
Fuller, G., Kuhar, C. W., Dennis, P. M., & Lukas, K. E. (2013). A survey of husbandry practices for lorisid primates in North American zoos and related facilities. Zoo Biology, 32(1), 88100.Google Scholar
Fuller, G., Lukas, K. E., Kuhar, C., & Dennis, P. M. (2014). A retrospective review of mortality in lorises and pottos in North American zoos, 1980–2010. Endangered Species Research, 23(3), 205217.CrossRefGoogle Scholar
Fuller, G., Nijman, V., Wirdateti, W., & Nekaris, K. A. I. (2016a). Do chemical cues in the venom of slow lorises repel avian predators?. Emu, 116(4), 435439.Google Scholar
Fuller, G., Raghanti, M. A., Dennis, P. M., Kuhar, C. W., Willis, M. A., Schook, M. W., & Lukas, K. E. (2016b). A comparison of nocturnal primate behavior in exhibits illuminated with red and blue light. Applied Animal Behaviour Science, 184, 126134.Google Scholar
Fuller, G., Eggen, W. F., Wirdateti, W., & Nekaris, K. A. I. (2018). Welfare impacts of the illegal wildlife trade in a cohort of confiscated greater slow lorises, Nycticebus coucang. Journal of Applied Animal Welfare Science, 21(3), 224238.Google Scholar
Fung, H. T., & Wong, O. F. (2016). Clinical quiz: A potentially toxic primate bite. Hong Kong Journal of Emergency Medicine, 23(5), 301303.Google Scholar
Gain, P. (2002). Bangladesh environment: Facing the 21st century. Dhaka: Society of Environment and Human Development.Google Scholar
Gamage, S., Liyanage, W. K. D. D., Weerakoon, D., & Gunwardena, A. (2009). Habitat quality and availability of the Sri Lanka red slender loris Loris tardigradus tardigradus (Mammalia: Primates: Lorisidae) in the Kottawa Arboretum. Journal of Threatened Taxa, 1(2), 6571.Google Scholar
Gamage, S., Groves, C. P., Manikar, F. M. M., Turner, C. S., Padmalal, K. U. K. G., & Kotagama, S. W. (2017). The taxonomy, distribution, and conservation status of the slender loris (Primates, Lorisidae: Loris) in Sri Lanka. Primate Conservation, 31, 83106.Google Scholar
Gambaryan, P. P. (1974). How mammals run: Anatomical adaptations. New York, NY: Wiley.Google Scholar
Garber, P. A. (1992). Vertical clinging, small body size, and the evolution of feeding adaptations in the Callitrichinae. American Journal of Physical Anthropology, 88(4), 469482.Google Scholar
Gardiner, M., Nekaris, K. A. I., & Samuel, P. (2017). The cytotoxic effect of slow loris (Nycticebus) venom on human cancer. Folia Primatologica, 88(2), 171.Google Scholar
Gardiner, M., Weldon, A., Poindexter, S. A., Gibson, N., & Nekaris, K. A. I. (2018). Survey of practitioners handling slow lorises (Primates: Nycticebus): An assessment of the harmful effects of slow loris bites. Journal of Venom Research, 9, 1.Google ScholarPubMed
Garrett, E. C. (2015). Was there a sensory trade-off in primate evolution? The vomeronasal groove as a means of understanding the vomeronasal system in the fossil record (unpublished doctoral dissertation). City University of New York, New York, NY, USA.Google Scholar
Garrett, E. C., & Steiper, M. E. (2014). Strong links between genomic and anatomical diversity in both mammalian olfactory chemosensory systems. Proceedings of the Royal Society B: Biological Sciences, 281(1783), 20132828.Google Scholar
Garrett, E. C., Dennis, J. C., Bhatnagar, K. P., Durham, E. L., Burrows, A. M., Bonar, C. J., …, & Smith, T. D. (2013). The vomeronasal complex of nocturnal strepsirhines and implications for the ancestral condition in primates. The Anatomical Record, 296(12), 18811894.Google Scholar
Gaston, K. J., Visser, M. E., & Hölker, F. (2015). The biological impacts of artificial light at night: The research challenge. Philosophical Transactions of the Royal Society B: Biological Sciences, 370(1667). DOI: https://doi.org/10.1098/rstb.2014.0133.Google Scholar
Gautier-Hion, A. (1980). Seasonal variations of diet related to species and sex in a community of Cercopithecus monkeys. Journal of Animal Ecology, 49(1), 237269.Google Scholar
Gavrilov, L. R., Gersuni, G. V., Ilyinsky, O. B., Sirotyuk, M. G., Tsirulnikov, E. M., & Shchekanov, E. E. (1976). The effect of focused ultrasound on the skin and deep nerve structures of man and animal. Progress in Brain Research, 43, 279292.Google Scholar
Gazzano, A., Zilocchi, M., Massoni, E., & Mariti, C. (2013). Dogs’ features strongly affect people’s feelings and behavior toward them. Journal of Veterinary Behavior: Clinical Applications and Research, 8(4), 213220.Google Scholar
Gebo, D. L. (1986). Miocene lorisids: The foot evidence. Folia Primatologica, 47(4), 217225.CrossRefGoogle Scholar
Gebo, D. L. (1987). Locomotor diversity in prosimian primates. American Journal of Primatology, 13(3), 271281.Google Scholar
Gebo, D. L. (1989). Postcranial adaptation and evolution in Lorisidae. Primates, 30(3), 347367.CrossRefGoogle Scholar
Gebo, D. L. (2002). Adapiformes: Phylogeny and adaptation. In Hartwig, W. C. (Ed.), The primate fossil record (pp. 2144). Cambridge: Cambridge University Press.Google Scholar
Gebo, D. L. (2014). Primate comparative anatomy. Baltimore, MD: Johns Hopkins University Press.Google Scholar
Gebo, D. L., MacLatchy, L., & Kityo, R. (1997a). A new lorisid humerus from the Early Miocene of Uganda. Primates, 38(4), 423427.Google Scholar
Gebo, D. L., MacLatchy, L., Kityo, R., Deino, A., Kingston, J., & Pilbeam, D. (1997b). A hominoid genus from the early Miocene of Uganda. Science, 276(5311), 401404.Google Scholar
Geerdts, M. S., Van de Walle, G. A., & LoBue, V. (2016). Learning about real animals from anthropomorphic media. Imagination, Cognition and Personality, 36(1), 526.Google Scholar
Geiser, F. (2013). Hibernation. Current Biology, 23(5), R188R193.Google Scholar
Geiser, F., & Ruf, T. (1995). Hibernation versus daily torpor in mammals and birds: Physiological variables and classification of torpor patterns. Physiological Zoology, 68(6), 935966.CrossRefGoogle Scholar
Génin, F. (2003). Female dominance in competition for gum trees in the grey mouse lemur. Revue d’Écologie, 58, 397410.Google Scholar
Génin, F. G., Masters, J. C., & Ganzhorn, J. U. (2010). Gummivory in cheirogaleids: Primitive retention or adaptation to hypervariable environments?. In Burrows, A. M. & Nash, L. T. (Eds.), The evolution of exudativory in primates (pp. 123140). New York, NY: Springer.Google Scholar
Geoffroy Saint-Hilaire, E. (1796). Mammifères:Mémoire sur les rapports naturels des makis Lemur, L. et description d’une espèce nouvelle de Mammifère. Magasin Encyclopédique, 2(1), 2050.Google Scholar
Geoffroy Saint-Hilaire, E. (1811). Sur les espèces du genre Loris, mammifères de l’ordre Quadrumanes. Annales Du Muséum D’Histoire Naturelle, Paris, 17, 164165.Google Scholar
Geoffroy Saint-Hilaire, E. (1812). Suite au tableau des Quadrumanes. Annales Du Muséum D’Histoire Naturelle, Paris, 19, 156170.Google Scholar
Gestich, C. C., Caselli, C. B., Nagy‐Reis, M. B., Setz, E. Z., & da Cunha, R. G. (2017). Estimating primate population densities: The systematic use of playbacks along transects in population surveys. American Journal of Primatology, 79(2), e22586.Google Scholar
Giam, X. (2017). Global biodiversity loss from tropical deforestation. Proceedings of the National Academy of Sciences, 114(23), 57755777.Google Scholar
Gibbons, W. J., & Andrews, K. M. (2004). PIT tagging: Simple technology at its best. Bioscience, 54(5), 447454.CrossRefGoogle Scholar
Gilad, Y., Wiebe, V., Przeworski, M., Lancet, D., & Pääbo, S. (2004). Loss of olfactory receptor genes coincides with the acquisition of full trichromatic vision in primates. PLoS Biology, 2(1), e5.Google Scholar
Gilbert, C. C., & Maiolino, S. A. (2015). Comment to ‘Primates in the Eocene’ by Gingerich (2012). Palaeobiodiversity and Palaeoenvironments, 95(2), 237241.Google Scholar
Gilbert, F. W., & Warren, W. E. (1995). Psychographic constructs and demographic segments. Psychology & Marketing, 12(3), 223237.CrossRefGoogle Scholar
Gill, F. B., & Wolf, L. L. (1975). Economics of feeding territoriality in the golden‐winged sunbird. Ecology, 56(2), 333345.CrossRefGoogle Scholar
Gingerich, P. D. (1973). Anatomy of the temporal bone in the Oligocene anthropoid Apidium and the origin of Anthropoidea. Folia Primatologica, 19(5), 329337.Google Scholar
Gingerich, P. D. (1975). Dentition of Adapis parisiensis and the origin of lemuriform primates. In Tattersall, I. & Sussman, R. W. (Eds.), Lemur biology (pp. 6580). New York, NY: Plenum.Google Scholar
Gingerich, P. D. (1977). Homologies of the anterior teeth in Indriidae and a functional basis for dental reduction in primates. American Journal of Physical Anthropology, 47(3), 387393.CrossRefGoogle Scholar
Gingerich, P. D. (1984). Primate evolution: Evidence from the fossil record, comparative morphology, and molecular biology. American Journal of Physical Anthropology, 27(S5), 5772.Google Scholar
Gingerich, P. D. (2012). Primates in the Eocene. In Lehman, T. & Schaal, S. F.K. (Eds.), The world at the time of Messel: Puzzles in palaeobiology, palaeoenvironment, and the history of early primates (pp. 6768). Frankfurt am Main: Senckenberg Gesellschaft für Naturforschung.Google Scholar
Gingerich, P. D., & Sahni, A. (1979). Indraloris and Sivaladapis: Miocene adapid primates from the Siwaliks of India and Pakistan. Nature, 279(5712), 415.CrossRefGoogle ScholarPubMed
Gingerich, P. D., & Schoeninger, M. (1977). The fossil record and primate phylogeny. Journal of Human Evolution, 6(5), 483505.Google Scholar
Gingerich, P. D., Franzen, J. L., Habersetzer, J., Hurum, J. H., & Smith, B. H. (2010). Darwinius masillae is a Haplorhine: Reply to Williams et al. (2010). Journal of Human Evolution, 59(5), 574579.Google Scholar
Glassman, D. M., & Wells, J. P. (1984). Positional and activity behavior in a captive slow loris: A quantitative assessment. American Journal of Primatology, 7(2), 121132.Google Scholar
Global Federation of Animal Sanctuaries (2013). Standards for prosimian sanctuaries. Retrieved 13 September 2019, from www.sanctuaryfederation.org/wp-content/uploads/2017/09/ProsimianStandards_Dec2015.pdf.Google Scholar
Godfrey, L. R., Winchester, J. M., King, J. M., Boyer, D. M., & Jernvall, J. (2012). Dental topography indicates ecological contraction of lemur communities. American Journal of Physical Anthropology, 148, 215227.Google Scholar
Godinot, M. (1998). A summary of adapiform systematics and phylogeny. Folia Primatologica, 69(Suppl. 1), 218249.CrossRefGoogle Scholar
Godinot, M. (2006). Lemuriform origins as viewed from the fossil record. Folia Primatologica, 77(6), 446464.CrossRefGoogle ScholarPubMed
Godinot, M. (2010). Paleogene prosimians. In Werdelin, L. & Sanders, W. J. (Eds.), Cenozoic mammals of Africa (pp. 319333). Berkeley, CA: University of California Press.Google Scholar
Godinot, M. (2015). Fossil record of the Primates from the Paleogene to the Oligocene. In Henke, W. & Tattersall, I. (Eds.), Handbook of paleoanthropology (pp. 11371259). Berlin: Springer.Google Scholar
Goffart, M., Missotten, L., Faidherbe, J., & Watillon, M. (1976). A duplex retina and the electroretinogram in the nocturnal Perodicticus potto. Archives Internationales De Physiologie Et De Biochimie, 84(3), 493516.Google Scholar
Goldingay, R., & Possingham, H. (1995). Area requirements for viable populations of the Australian gliding marsupial Petaurus australis. Biological Conservation, 73(2), 161167.Google Scholar
Gomez, A., Petrzelkova, K., Yeoman, C. J., Vlckova, K., Mrázek, J., Koppova, I., …, & Torralba, M. (2015). Gut microbiome composition and metabolomic profiles of wild western lowland gorillas (Gorilla gorilla gorilla) reflect host ecology. Molecular Ecology, 24(10), 25512565.Google Scholar
Gómez-Laich, A. G., Wilson, R. P., Quintana, F., & Shepard, E. L. (2008). Identification of imperial cormorant Phalacrocorax atriceps behaviour using accelerometers. Endangered Species Research, 10, 2937.Google Scholar
Gonçalves, B. S., Cavalcanti, P. R., Tavares, G. R., Campos, T. F., & Araujo, J. F. (2014). Nonparametric methods in actigraphy: An update. Sleep Science, 7(3), 158164.Google Scholar
Gong, J. D., Cheng, G., Xue, D. Y., Guo, Y. J., & Yang, J. B. (2012). Status of Jingpo people’s traditional medicinal knowledge of animals in Longchuan, Yunnan [J]. Journal of Yunnan Agricultural University (Natural Science), 3.Google Scholar
Goodman, M. (1967). Deciphering primate phylogeny from macromolecular specificities. American Journal of Physical Anthropology, 26(2), 255275.Google Scholar
Goodman, M., Porter, C. A., Czelusniak, J., Page, S. L., Schneider, H., Shoshani, J., …, & Groves, C. P. (1998). Toward a phylogenetic classification of primates based on DNA evidence complemented by fossil evidence. Molecular Phylogenetics and Evolution, 9(3), 585598.Google Scholar
Goonan, P. M. (1993). Behaviour and reproduction of the slender loris (Loris tardigradus) in captivity. Folia Primatologica, 60(3), 146157.Google Scholar
Goto, R., & Kumakura, H. (2013). The estimated mechanical advantage of the prosimian ankle joint musculature, and implications for locomotor adaptation. Journal of Anatomy, 222, 538546.Google Scholar
Gould, S. J. (1975). Allometry in primates, with emphasis on scaling and the evolution of the brain. Contributions to Primatology, 5, 244292.Google Scholar
Gould, S. J. (1977). Ontogeny and phylogeny. Cambridge, MA: Belknap Press.Google Scholar
Granatosky, M. C., Lemelin, P., Chester, S. G., Pampush, J. D., & Schmitt, D. (2014a). Functional and evolutionary aspects of axial stability in euarchontans and other mammals. Journal of Morphology, 275(3), 313327.Google Scholar
Granatosky, M. C., Miller, C. E., Boyer, D. M., & Schmitt, D. (2014b). Lumbar vertebral morphology of flying, gliding, and suspensory mammals: Implications for the locomotor behavior of the subfossil lemurs Palaeopropithecus and Babakotia. Journal of Human Evolution, 75, 4052.CrossRefGoogle ScholarPubMed
Granatosky, M. C., Schmitt, D., & Hanna, J. (2019). Comparison of spatiotemporal gait characteristics between vertical climbing and horizontal walking in primates. Journal of Experimental Biology, 222(2). DOI: http://doi.org/10.1242/jeb.185702.Google Scholar
Grand, T. (1967). The functional anatomy of the ankle and foot of the slow loris (Nycticebus coucang). American Journal of Physical Anthropology, 26(2), 207218.Google Scholar
Grand, T. (1984). Motion economy within the canopy: Four strategies for mobility. In Rodman, P. S. & Cant, J. G.H. (Eds.), Adaptations for foraging in non-human primates: Contributions to an organismal biology of prosimians, monkeys, and apes (pp. 5472). New York, NY: Columbia University Press.Google Scholar
Grant, R. A., Haidarliu, S., Kennerley, N. J., & Prescott, T. J. (2013). The evolution of active vibrissal sensing in mammals: Evidence from vibrissal musculature and function in the marsupial opossum Monodelphis domestica. Journal of Experimental Biology, 216(18), 34833494.Google Scholar
Grant, R. A., Delaunay, M. G., & Haidarliu, S. (2017). Mystacial whisker layout and musculature in the guinea pig (Cavia porcellus): A social, diurnal mammal. The Anatomical Record, 300(3), 527536.Google Scholar
Grant, R. A., Breakell, V., & Prescott, T. J. (2018). Whisker touch sensing guides locomotion in small, quadrupedal mammals. Proceedings of the Royal Society B: Biological Sciences, 285(1880), 20180592.Google Scholar
Gray, A. E., Wirdateti, , & Nekaris, K. A. I. (2015). Trialling exudate-based enrichment efforts to improve the welfare of rescued slow lorises Nycticebus spp. Endangered Species Research, 27(1), 2129.Google Scholar
Gray, J. E. (1821). On the natural arrangement of vertebrose animals. London Medical Repository, 15(1), 296310.Google Scholar
Gray, J. E. (1863). Revision of the species of lemuroid animals, with the description of some new species. Proceedings of the Zoological Society of London, 1863, 129152.Google Scholar
GrayJ. E. (1870). Catalogue of monkeys, lemurs, and fruit-eating bats in the collection of the British Museum. London: British Museum.Google Scholar
Greatorex, Z. F., Olson, S. H., Singhalath, S., Silithammavong, S., Khammavong, K., Fine, A. E., …, & Gilbert, M. (2016). Wildlife trade and human health in Lao PDR: An assessment of the zoonotic disease risk in markets. PLoS One, 11(3), e0150666.Google Scholar
Greene, L. K., Grogan, K. E., Smyth, K. N., Adams, C. A., Klager, S. A., & Drea, C. M. (2016). Mix it and fix it: Functions of composite olfactory signals in ring-tailed lemurs. Royal Society Open Science, 3(4), 160076.Google Scholar
Greengrass, E. (2011). Exploring the dynamics of bushmeat hunting and trade in Sapo National Park. Cambridge: Fauna and Flora International.Google Scholar
Gregory, W. M. (1920). On the structure and relations of Notharctus, an American Eocene primate. Memoirs of the American Museum of Natural History, 3, 49243.Google Scholar
Gregory, W. K. (1922). The origin and evolution of the human dentition. Baltimore, MD: Williams and Wilkins.Google Scholar
Gris, K. V., Coutu, J. P., & Gris, D. (2017). Supervised and unsupervised learning technology in the study of rodent behavior. Frontiers in Behavioral Neuroscience, 11, 141.Google Scholar
Groves, C. P. (1971). Systematics of the genus Nycticebus. In Biegert, J. & Leutenegger, W. (Eds.), Taxonomy, anatomy, reproduction: Proceedings of the third International Congress of Primatology (vol. , pp. 4453). Basel: Karger.Google Scholar
Groves, C. P. (1998). Pseudopotto martini: A new potto. African Primates, 3, 4243.Google Scholar
Groves, C. P. (2001). Primate taxonomy. Washington, DC: Smithsonian Institution Press.Google Scholar
Groves, C. P. (2004). Taxonomy and biogeography of primates in Vietnam and neighbouring regions. In Nadler, T., Streicher, U., & Ha, T. L. (Eds.), Conservation of Primates in Vietnam (pp. 1522). Hanoi: Frankfurt Zoological Society.Google Scholar
Groves, C. P. (2005). Order Primates. In Wilson, D. E. & Reeder, D. M. (Eds.), Mammal species of the world: A taxonomic and geographic reference (3rd ed.) (pp. 111184). Baltimore, MD: Johns Hopkins University Press.Google Scholar
Groves, C. P. (2008). Extended family: Long lost cousins. Arlington, VA: Conservation International.Google Scholar
Groves, C. P., & Maryanto, I. (2008). Craniometry of slow lorises (genus Nycticebus) of insular Southeast Asia. In Shekelle, M., Maryanto, I., Groves, C., Schulze, H., & Fitch-Snyder, H. (Eds.), Primates of the Oriental Night (pp. 115122). Treubia: LIPI Press.Google Scholar
Grubb, P., Butynski, T. M., Oates, J. F., Bearder, S. K., Disotell, T. R., Groves, C. P., & Struhsaker, T. T. (2003). Assessment of the diversity of African primates. International Journal of Primatology, 24(6), 13011357.Google Scholar
Grundmann, E., & Didier, S. (2000). Adaptation of orphaned chimpanzees (Pan troglodytes troglodytes) and orangutans (Pongo pygmaeus) to reintroduction to the forest: Activity budgets, feeding and foraging behaviour. Folia Primatologica, 71(4), 195.Google Scholar
Guillera-Arroita, G., Lahoz-Monfort, J. J., Milner-Gulland, E. J., Young, R. P., & Nicholson, E. (2010). Using occupancy as a state variable for monitoring the Critically Endangered Alaotran gentle lemur Hapalemur alaotrensis. Endangered Species Research, 11(2), 157166.Google Scholar
Gunnell, G. F., Boyer, D. M., Friscia, A. R., Heritage, S., Manthi, F. K., Miller, E. R., …, & Seiffert, E. R. (2018). Fossil lemurs from Egypt and Kenya suggest an African origin for Madagascar’s aye-aye. Nature Communications, 9(1), 3193.Google Scholar
Gunther, M. M. (1989). Funktionsmorphologische Untersuchungen zum Sprungverhalten an mehreren halbaffenarten (Galago moholi, Galago (Otolemur) garnetti, Lemur catta) (unpublished doctoral dissertation). Freie Universitat, Berlin, Germany.Google Scholar
Gupta, K. K. (2005). Relationships of male condition, sociality, ranging and habitat use in the slender loris (Loris tardigradus) in Kalakad-Mundanthurai Tiger Reserve, India. American Journal of Physical Anthropology, 40, 127.Google Scholar
Gupta, K. K. (2007). Socioecology and conservation of the slender loris (Loris tardigradus) in southern India (unpublished doctoral disseration). Arizona State University, Tempe, AZ, USA.Google Scholar
Gupta, K. K., & Katti, M. (2007). Size matters: Male mating tactics of slender lorises in a human modified ecosystem, Presentation at the Society for Integrative and Comparative Biology annual meeting, Phoenix, 4–8 January 2007. Henderson, VA: The Society for Integrative and Comparative Biology.Google Scholar
Gursky, S. (2003). Lunar philia in a nocturnal primate. International Journal of Primatology, 24(2), 351367.Google Scholar
Gursky, S. (2015). Ultrasonic vocalizations by the spectral tarsier, Tarsius spectrum. Folia Primatologica, 86(3), 153163.CrossRefGoogle ScholarPubMed
Gursky, S., & Moser, C. (2018). Echolocation in a nocturnal primate? American Journal of Physical Anthropology, 165, 109.Google Scholar
Gursky, S., & Nekaris, K. A. I. (2003). An introduction to mating, birthing and rearing systems of nocturnal prosimians. Folia Primatologica, 74(5–6), 272284.Google Scholar
Gursky-Doyen, S. (2010). Intraspecific variation in the mating system of spectral tarsiers. International Journal of Primatology, 31(6), 11611173.Google Scholar
Gyambibi, A., & Lemelin, P. (2013). Comparative and quantitative myology of the forearm and hand of prosimian primates. The Anatomical Record, 296(8), 11961206.Google Scholar
Hagey, L. R., Fry, B. G., & Fitch-Snyder, H. (2007). Talking defensively, a dual use for the brachial gland exudate of slow and pygmy lorises. In Gursky, S. L. & Nekaris, K. A. I. (Eds.), Primate anti-predator strategies (pp. 253272). Boston, MA: Springer.Google Scholar
Haidarliu, S., Simony, E., Golomb, D., & Ahissar, E. (2010). Muscle architecture in the mystacial pad of the rat. The Anatomical Record, 293(7), 11921206.Google Scholar
Haig, S. M., Beever, E. A., Chambers, S. M., Draheim, H. M., Dugger, B. D., Dunham, S., …, & Lopes, I. F. (2006). Taxonomic considerations in listing subspecies under the US Endangered Species Act. Conservation Biology, 20(6), 15841594.CrossRefGoogle Scholar
Hall, R. J., & Langtimm, C. A. (2001). The US National Amphibian Research and Monitoring Initiative and the role of protected areas. The George Wright Forum, 18(2), 1425.Google Scholar
Halle, S., & Stenseth, N. C. (2000). Chronoecology: New light through old windows – a conclusion. In Halle, S. & Stenseth, N. C. (Eds.), Activity patterns in small mammals (pp. 275284). Berlin: Springer.Google Scholar
Halsey, L. G., Shepard, E. L., & Wilson, R. P. (2011). Assessing the development and application of the accelerometry technique for estimating energy expenditure. Comparative Biochemistry and Physiology Part A: Molecular & Integrative Physiology, 158(3), 305314.Google Scholar
Hanna, J. B. (2006). Kinematics of vertical climbing in lorises and Cheirogaleus medius. Journal of Human Evolution, 50(4), 469478.Google Scholar
Hanna, J. B., & Schmitt, D. (2011a). Interpreting the role of climbing in primate locomotor evolution: Are the biomechanics of climbing influenced by habitual substrate use and anatomy?. International Journal of Primatology, 32(2), 430444.CrossRefGoogle Scholar
Hanna, J. B., & Schmitt, D. (2011b). Locomotor energetics in primates: Gait mechanics and their relationship to the energetics of vertical and horizontal locomotion. American Journal of Physical Anthropology, 145(1), 4354.Google Scholar
Hanna, J. B., & Venkataraman, V. (in press). Experimental research on foot use and function during climbing by primates. In Zeininger, A., Hatala, K., Wunderlich, R. E., & Schmitt, D. (Eds.), Evolution of the primate foot. New York: Springer.Google Scholar
Hanna, J. B., Schmitt, D., & Griffin, T. M. (2008). The energetic cost of climbing in primates. Science, 320(5878), 898898.Google Scholar
Hanna, J. B., Chan, L. K., & Schmitt, D. (2011). Loris locomotor behavior in relation to skeletal morphology: Disjunction between assumed mobility and utilized range of motion. American Journal of Physical Anthropology, 144, 155156.Google Scholar
Hanna, J. B., Schmitt, D., Wright, K., Eshchar, Y., Visalberghi, E., & Fragaszy, D. (2015). Kinetics of bipedal locomotion during load carrying in capuchin monkeys. Journal of Human Evolution, 85, 149156.CrossRefGoogle ScholarPubMed
Hanna, J. B., Granatosky, M. C., Rana, P., & Schmitt, D. (2017). The evolution of vertical climbing in primates: Evidence from reaction forces. Journal of Experimental Biology, 220(17), 30393052.Google Scholar
Happold, D., & Happold, M. (1992). Termites as food for the thick-tailed bushbaby (Otolemur crassicaudatus) in Malawi. Folia Primatologica, 58, 118120.Google Scholar
Harcourt, A. H., Harvey, P. H., Larson, S. G., & Short, R. V. (1981). Testis weight, body weight and breeding system in primates. Nature, 293(5827), 55.Google Scholar
Harcourt, C. S. (1986). Seasonal variation in the diet of South African galagos. International Journal of Primatology, 7(5), 491506.Google Scholar
Harcourt, C. S. (1981). An examination of the function of urine washing in Galago senegalensis. Zeitschrift Für Tierpsychologie, 55, 119128.Google Scholar
Harcourt, C. S., & Nash, L. T. (1986a). Species differences in substrate use and diet between sympatric galagos in two Kenyan coastal forests. Primates, 27, 4152.CrossRefGoogle Scholar
Harcourt, C. S., & Nash, L. T. (1986b). Social organization of galagos in Kenyan coastal forests: I. Galago zanzibaricus. American Journal of Primatology, 10(4), 339355.Google Scholar
Hardy, S. A., & Carlo, G. (2011). Moral identity: What is it, how does it develop, and is it linked to moral action?. Child Development Perspectives, 5(3), 212218.Google Scholar
Harmon, L. J., Weir, J. T., Brock, C. D., Glor, R. E., & Challenger, W. (2008). GEIGER: Investigating evolutionary radiations. Bioinformatics, 24(1), 129131.CrossRefGoogle ScholarPubMed
Harris, R. L., Davies, N. W., & Nicol, S. C. (2012). Chemical composition of odorous secretions in the Tasmanian short-beaked echidna (Tachyglossus aculeatus setosus). Chemical Senses, 37(9), 819836.Google Scholar
Harris, S., Cresswell, W. J., Forde, P. G., Trewhella, W. J., Woollard, T., & Wray, S. (1990). Home‐range analysis using radio‐tracking data: A review of problems and techniques particularly as applied to the study of mammals. Mammal Review, 20(2–3), 97123.Google Scholar
Harrison, T. (1989). A new species of Micropithecus from the middle Miocene of Kenya. Journal of Human Evolution, 18(6), 537557.Google Scholar
Harrison, T. (2010). Later Tertiary Lorisiformes. In Werdelin, L. & Sanders, W. J. (Eds.), Cenozoic mammals of Africa (pp. 333349) Berkeley, CA: University of California Press.CrossRefGoogle Scholar
Harrison, T., & Andrews, P. (2009). The anatomy and systematic position of the early Miocene proconsulid from Meswa Bridge, Kenya. Journal of Human Evolution, 56(5), 479496.Google Scholar
Harvey, E. S., Butler, J. J., McLean, D. L., & Shand, J. (2012). Contrasting habitat use of diurnal and nocturnal fish assemblages in temperate Western Australia. Journal of Experimental Marine Biology and Ecology, 426, 7886.CrossRefGoogle Scholar
Harvey, P. H., & Clutton‐Brock, T. H. (1985). Life history variation in primates. Evolution, 39(3), 559581.Google Scholar
Harzhauser, M., Kroh, A., Mandic, O., Piller, W. E., Göhlich, U., Reuter, M., & Berning, B. (2007). Biogeographic responses to geodynamics: A key study all around the Oligo–Miocene Tethyan Seaway. Zoologischer Anzeiger: A Journal of Comparative Zoology, 246(4), 241256.CrossRefGoogle Scholar
Hashimoto, C. (1995). Population census of the chimpanzees in the Kalinzu Forest, Uganda: Comparison between methods with nest counts. Primates, 36(4), 477488.Google Scholar
Hastie, J., & McCrea-Steele, T. (2014). Wanted – dead or alive: Exposing online wildlife trade. London: International Fund for Animal Welfare.Google Scholar
Hawes, J., & Ogden, J. (2001). Birth management and hand rearing. In Fitch-Snyder, H. & Schulze, H. (Eds.), Management of lorises in captivity: A husbandry manual for Asian lorisines (Nycticebus & Loris ssp.) (pp. 4550). San Diego, CA: Zoological Society of San Diego.Google Scholar
Hayden, S., & Teeling, E. C. (2014). The molecular biology of vertebrate olfaction. The Anatomical Record, 297(11), 22162226.Google Scholar
Haydon, M. J. (1994). The ecology and management of rain forest ungulates in Sabah, Malaysia: Implications of forest disturbance. Aberdeen: Institute of Tropical Biology, University of Aberdeen.Google Scholar
Hayne, D. W. (1949). Calculation of size of home range. Journal of Mammalogy, 30(1), 118.Google Scholar
Hedewig, R. (1980). Vergleichende anatomische Untersuchungen an der Jacobsonschen Organen von Nycticebus coucang Boddaert, 1785 (Prosimiae, Lorisidae) und Galago crassicaudatus E. Geoffroy, 1812 (Prosimiae, Lorisidae). Gegenbaurs Morphologische Jahrbuch Leipzig, 126, 676722.Google Scholar
Heesy, C. P. (2008). Ecomorphology of orbit orientation and the adaptive significance of binocular vision in primates and other mammals. Brain, Behavior and Evolution, 71(1), 5467.Google Scholar
Hemingway, C. A. (1999). Time budgets and foraging in a Malagasy primate: Do sex differences reflect reproductive condition and female dominance?. Behavioral Ecology and Sociobiology, 45(3–4), 311322.Google Scholar
Hending, D., Andrianiaina, A., Rakotomalala, Z., & Cotton, S. (2017). Range extension and behavioural observations of the recently described Sheth’s dwarf lemur (Cheirogaleus shethi). Folia Primatologica, 88(5), 401408.Google Scholar
Hending, D., Andrianiaina, A., Rakotomalala, Z., & Cotton, S. (2018). The use of vanilla plantations by lemurs: Encouraging findings for both lemur conservation and sustainable agroforestry in the Sava Region, northeast Madagascar. International Journal of Primatology, 39(1), 141153.Google Scholar
Heritage, S. (2014). Modeling olfactory bulb evolution through primate phylogeny. PLoS One, 9(11), e113904.CrossRefGoogle ScholarPubMed
Herrera, J. P., & Dávalos, L. M. (2016). Phylogeny and divergence times of lemurs inferred with recent and ancient fossils in the tree. Systematic Biology, 65(5), 772791.Google Scholar
Hershkovitz, P. (1977). Living New World monkeys (Platyrrhini). Chicago, IL: University of Chicago Press.Google Scholar
Hertenstein, B., Zimmermann, E., & Rahmann, H. (1987). The development of visual acuity in tree shrews, bushbabies, and slow loris (Abstract). International Journal of Primatology, 8(5), 533.Google Scholar
Herzog, H. A., & Mathews, S. (1997). Personality and attitudes toward the treatment of animals. Society & Animals, 5(2), 169175.Google Scholar
Hildebrand, M. (1976). Analysis of tetrapod gaits: General considerations and symmetrical gaits. In Herman, R., Grillner, S., Stein, P., & Stuary, D. (Eds.), Neural control of locomotion (pp. 203236). Boston, MA: Springer.Google Scholar
Hill, R. A., & Dunbar, R. I. M. (2002). Climatic determinants of diet and foraging behaviour in baboons. Evolutionary Ecology, 16(6), 579593.Google Scholar
Hill, R. P., Gaines, J., & Wilson, R. M. (2008). Consumer behavior, extended-self, and sacred consumption: An alternative perspective from our animal companions. Journal of Business Research, 61(5), 553562.Google Scholar
Hill, W. C. O. (1936). The affinities of the lorisoids. Ceylon Journal of Science B, 19, 287314.Google Scholar
Hill, W. C. O. (1947a). The lorisoid genus Arctocebus: Observations based on the type material. Proceedings of the Royal Society of Edinburgh. Section B: Biology, 62(3), 248265.Google Scholar
Hill, W. C. O. (1947b). Untitled discussion of the biology of lorisids. Proceedings of the Zoological Society of London, 117(1), 278.Google Scholar
Hill, W. C. O. (1953a). Early records of the slender loris and its allies. Proceedings of the Zoological Society of London, 123(1), 4347.CrossRefGoogle Scholar
Hill, W. C. O. (1953b). Primates: Comparative anatomy and taxonomy: I – Strepsirhini. Edinburgh: Edinburgh University Press.Google Scholar
Hinde, K., & Milligan, L. A. (2011). Primate milk: Proximate mechanisms and ultimate perspectives. Evolutionary Anthropology: Issues, News, and Reviews, 20(1), 923.CrossRefGoogle ScholarPubMed
Hinde, K., Skibiel, A. L., Foster, A. B., Del Rosso, L., Mendoza, S. P., & Capitanio, J. P. (2015). Cortisol in mother’s milk across lactation reflects maternal life history and predicts infant temperament. Behavioral Ecology, 26(1), 269281.CrossRefGoogle ScholarPubMed
Hinsley, A., Lee, T. E., Harrison, J. R., & Roberts, D. L. (2016). Estimating the extent and structure of trade in horticultural orchids via social media. Conservation Biology, 30(5), 10381047.Google Scholar
Hirasaki, E., Kumakura, H., & Matano, S. (2000). Biomechanical analysis of vertical climbing in the spider monkey and the Japanese macaque. American Journal of Physical Anthropology, 113(4), 455472.Google Scholar
Hirokawa, Y., & Kumakura, H. (2003). Functional analysis of the thigh muscles during locomotion in the Garnet galago (Galago garnetti). Anthropological Science, 111(2), 187201.Google Scholar
Hladik, C. M. (1979). Diet and ecology of prosimians. In Doyle, G. A. & Martin, R. D. (Eds.), The study of prosimian behavior (pp. 307357). New York, NY: Academic Press.Google Scholar
Hladik, C. M., & Simmen, B. (1996). Taste perception and feeding behavior in non-human primates and human populations. Evolutionary Anthropology: Issues, News, and Reviews, 5(2), 5871.Google Scholar
Hoffmann, J. N., Montag, A. G., & Dominy, N. J. (2004). Meissner corpuscles and somatosensory acuity: The prehensile appendages of primates and elephants. The Anatomical Record Part A, 281(1), 11381147.Google Scholar
Hoffstetter, R. (1977). Phylogénie des primates: Confrontation des resultats obtenus par les diverses voies d’approche du probleme. Bulletins Et Mémoires De La Société D’Anthropologie De Paris, 4(4), 327346.Google Scholar
Hofmann, T., Ellenberg, H., & Roth, H. H. (1999). Bushmeat: A natural resource of the moist forest regions of West Africa: With particular consideration of two duiker species in Côte d’Ivoire and Ghana. Eschborn: Deutsche Gesellschaft für Technische Zusammenarbeit.Google Scholar
Hofner, A. N. (2016). The value of voices: Moving beyond ecology in primate conservation (unpublished Master’s dissertation). Oxford Brookes University, Oxford, UK.Google Scholar
Hofner, A. N., Robinson, C. A. J., & Nekaris, K. A. I. (2018). Preserving Preuss’s red colobus (Piliocolobus preussi): An ethnographic analysis of hunting, conservation, and changing perceptions of primates in Ikenge-Bakoko, Cameroon. International Journal of Primatology, 39(5), 895917.Google Scholar
Hohenbrink, P., Mundy, N. I., Zimmermann, E., & Radespiel, U. (2013). First evidence for functional vomeronasal 2 receptor genes in primates. Biology Letters, 9(1), 20121006.Google Scholar
Holbrook, M. B. (2008). Pets and people: Companions in commerce? Journal of Business Research, 61(5), 546552.Google Scholar
Holiday Watchdog. (2016). Marmaris holidays. Retrieved 11 March 2016, from www.holidaywatchdog.com/Marmaris-Hoilday-Reviews-221.html.Google Scholar
Holland, G., & Tiggemann, M. (2017). ‘Strong beats skinny every time’: Disordered eating and compulsive exercise in women who post fitspiration on Instagram. International Journal of Eating Disorders, 50(1), 7679.CrossRefGoogle ScholarPubMed
Hopwood, A. T. (1933). Miocene primates from Kenya. Zoological Journal of the Linnean Society, 38(260), 437464.Google Scholar
Horvath, J. E., Weisrock, D. W., Embry, S. L., Fiorentino, I., Balhoff, J. P., Kappeler, P., …, & Yoder, A. D. (2008). Development and application of a phylogenomic toolkit: Resolving the evolutionary history of Madagascar’s lemurs. Genome Research, 18(3), 489499.Google Scholar
Hu, H., Liu, W., & Cao, M. (2008). Impact of land use and land cover changes on ecosystem services in Menglun, Xishuangbanna, Southwest China. Environmental Monitoring and Assessment, 146(1–3), 147156.Google Scholar
Huber, H. F., & Lewis, K. P. (2011). An assessment of gum‐based environmental enrichment for captive gummivorous primates. Zoo Biology, 30(1), 7178.CrossRefGoogle ScholarPubMed
Huber, R., Hill, S. L., Holladay, C., Biesiadecki, M., Tononi, G., & Cirelli, C. (2004). Sleep homeostasis in Drosophila melanogaster. Sleep, 27(4), 628639.Google Scholar
Hughes, A. (1977). The topography of vision in mammals of contrasting life style: Comparative optics and retinal organisation. In Crescitelli, F. (Ed.), The visual system in vertebrates (pp. 613756). Berlin: Springer.Google Scholar
Hughes, A. C. (2017). Understanding the drivers of Southeast Asian biodiversity loss. Ecosphere, 8(1), e01624.Google Scholar
Hughes, C. (2013). Exploring children’s perceptions of cheetahs through storytelling: Implications for cheetah conservation. Applied Environmental Education & Communication, 12(3), 173186.Google Scholar
Hughey, L. F., Hein, A. M., Strandburg-Peshkin, A., & Jensen, F. H. (2018). Challenges and solutions for studying collective animal behaviour in the wild. Philosophical Transactions of the Royal Society B: Biological Sciences, 373(1746), 20170005.Google Scholar
Hunt, K. D., Cant, J. G., Gebo, D. L., Rose, M. D., Walker, S. E., & Youlatos, D. (1996). Standardized descriptions of primate locomotor and postural modes. Primates, 37(4), 363387.CrossRefGoogle Scholar
Hunter, J. J., Morgan, J. I., Merigan, W. H., Sliney, D. H., Sparrow, J. R., & Williams, D. R. (2012). The susceptibility of the retina to photochemical damage from visible light. Progress in Retinal and Eye Research, 31(1), 2842.Google Scholar
Huntington, H. P. (2000). Using traditional ecological knowledge in science: Methods and applications. Ecological Applications, 10(5), 12701274.Google Scholar
Hurlbert, S. H. (1978). The measurement of niche overlap and some relatives. Ecology, 59(1), 6777.Google Scholar
Huxley, T. H. (1864). On the angwantibo (Arctocebus claabarensis, Gray) of Old Calabar. Proceedings of the Zoological Society of London, 32, 314585.Google Scholar
Hylander, W. L., & Ravosa, M. J. (1992). An analysis of the supraorbital region of primates: A morphometric and experimental approach. In Smith, P. and Tchernov, E. (Eds.), Structure, function and evolution of teeth (pp. 223255). London: Freund.Google Scholar
Ibarra-Soria, X., Levitin, M. O., & Logan, D. W. (2014). The genomic basis of vomeronasal-mediated behaviour. Mammalian Genome, 25(1–2), 7586.CrossRefGoogle ScholarPubMed
IFAW. (2005). Caught in the web: Wildlife trade on the internet. London: IFAW.Google Scholar
International Board of Zoological Nomenclature. (2002). Opinion 1995 (Case 3004). Lorisidae Gray, 1821, Galagidae Gray, 1825 and Indriidae Burnett, 1828 (Mammalia, Primates): Conserved as the correct original spellings. Bulletin of Zoological Nomenclature, 59(1), 6567.Google Scholar
Irwin, M. T., Jean‐Luc Raharison, F., Rakotoarimanana, H., Razanadrakoto, E., Ranaivoson, E., Rakotofanala, J., & Randrianarimanana, C. (2007). Diademed sifakas (Propithecus diadema) use olfaction to forage for the inflorescences of subterranean parasitic plants (Balanophoraceae: Langsdorffia sp., and Cytinaceae: Cytinus sp.). American Journal of Primatology, 69(4), 471476.CrossRefGoogle ScholarPubMed
Ishida, H., & Pickford, M. (1997). A new Late Miocene hominoid from Kenya: Samburupithecus kiptalami gen. et sp. nov. Comptes Rendus de l’Académie des Sciences-Series IIA: Earth and Planetary Science, 325(10), 823829.Google Scholar
Ishida, H., Kawabata, N., & Matano, S. (1983). Mode of descending and functional morphology of the biceps femoris muscle in the slow loris (Nycticebus coucang). Annales Des Sciences Naturelles, Zoologie, 5, 6774.Google Scholar
Ishida, H., Jouffroy, F., & Nakano, Y. (1990). Comparative dynamics of pronograde and upside down horizontal quadrupedalism in the slow loris (Nycticebus coucang). In Jouffroy, F., Stack, M., & Niemitz, C. (Eds.), Gravity, posture and locomotion in primates (pp. 209220) (2nd ed.). Florence: Sedicesimo.Google Scholar
Ishida, H., Hirasaki, E., & Matano, S. (1992). Locomotion of the slow loris between discontinuous substrates. Topics in Primatology, 3, 139152.Google Scholar
Isler, K. (2004). Footfall patterns, stride length and speed of vertical climbing in spider monkeys (Ateles fusciceps robustus) and woolly monkeys (Lagothrix lagotricha). Folia Primatologica, 75(3), 133149.Google Scholar
Isler, K. (2005). 3D‐kinematics of vertical climbing in hominoids. American Journal of Physical Anthropology, 126(1), 6681.Google Scholar
IUCN. (2018). IUCN Red List of Threatened Species: version 2018-1. Retrieved 27 July 2018, from www.iucnredlist.org.Google Scholar
IUCN-CEC. (2010). Communicating biodiversity. Gland: IUCN.Google Scholar
Izard, M. K., & Nash, L. T. (1988). Contrasting reproductive parameters in Galago senegalensis braccatus and G. s. moholi. International Journal of Primatology, 9(6), 519527.Google Scholar
Izard, M. K., & Rasmussen, D. T. (1985). Reproduction in the slender loris (Loris tardigradus malabaricus). American Journal of Primatology, 8, 153165.Google Scholar
Izard, M. K., & Weisenseel, K. (1989). Comparative reproduction of the lorisidae. American Journal of Primatology,18, 140.Google Scholar
Izard, M. K., Weisenseel, K. A., & Ange, R. L. (1988). Reproduction in the slow loris (Nycticebus coucang). American Journal of Primatology, 16(4), 331339.Google Scholar
Jack, K. M., Lenz, B. B., Healan, E., Rudman, S., Schoof, V. A., & Fedigan, L. (2008). The effects of observer presence on the behavior of Cebus capucinus in Costa Rica. American Journal of Primatology, 70(5), 490494.Google Scholar
Jacobs, G. H. (1977). Visual capacities of the owl monkey (Aotus trivirgatus): II. Spatial contrast sensitivity. Vision Research, 17(7), 821825.CrossRefGoogle ScholarPubMed
Jacobs, G. H. (1981). Comparative color vision. New York, NY: Academic Press.Google Scholar
Jacobs, G. H. (2008). Primate color vision: A comparative perspective. Visual Neuroscience, 25(5–6), 619633.Google Scholar
Jacobs, G. H. (2013). Losses of functional opsin genes, short-wavelength cone photopigments, and color vision: A significant trend in the evolution of mammalian vision. Visual Neuroscience, 30(1–2), 3953.Google Scholar
Jacobs, G. H., Neitz, M., & Neitz, J. (1996). Mutations in S-cone pigment genes and the absence of colour vision in two species of nocturnal primate. Proceedings of the Royal Society B: Biological Sciences, 263(1371), 705710.Google Scholar
Jacobs, G. H., Deegan, J. F. II, Tan, Y., & Li, W. H. (2002). Opsin gene and photopigment polymorphism in a prosimian primate. Vision Research, 42(1), 1118.CrossRefGoogle Scholar
Jacobs, L. L. (1981). Tooth comb in Nycticeboides simpsoni from the Miocene Siwaliks. Nature, 289, 585586.Google Scholar
Jacobson, S. K. (2010). Effective primate conservation education: Gaps and opportunities. American Journal of Primatology, 72(5), 414419.Google Scholar
Jacobson, S. K., McDuff, M. D., & Monroe, M. C. (2015). Conservation education and outreach techniques (2nd ed.). Oxford: Oxford University Press.Google Scholar
Jaeger, J. J., Beard, K. C., Chaimanee, Y., Salem, M., Benammi, M., Hlal, O., …, & Valentin, X. (2010). Late middle Eocene epoch of Libya yields earliest known radiation of African anthropoids. Nature, 467(7319), 10951098.Google Scholar
Janssen, J., & Chng, S. C. (2018). Biological parameters used in setting captive‐breeding quotas for Indonesia’s breeding facilities. Conservation Biology, 32(1), 1825.Google Scholar
Jayne, B. C., Lehmkuhl, A. M., & Riley, M. A. (2014). Hit or miss: Branch structure affects perch choice, behaviour, distance and accuracy of brown tree snakes bridging gaps. Animal Behaviour, 88, 233241.Google Scholar
Jeanniard‐du‐Dot, T., Guinet, C., Arnould, J. P., Speakman, J. R., & Trites, A. W. (2017). Accelerometers can measure total and activity‐specific energy expenditures in free‐ranging marine mammals only if linked to time–activity budgets. Functional Ecology, 31(2), 377386.Google Scholar
Jenkins, D. J., Axelsen, M., Kendall, C. W., Augustin, L. S., Vuksan, V., & Smith, U. (2000). Dietary fibre, lente carbohydrates and the insulin-resistant diseases. British Journal of Nutrition, 83(S1), S157-S163.Google Scholar
Jenkins, D. V. (1971). Animal life of Kinabalu National Park. Malaysian Nature Journal, 24, 177183.Google Scholar
Jenkins, F. A. Jr (1970). Anatomy and function of expanded ribs in certain edentates and primates. Journal of Mammalogy, 51(2), 288301.Google Scholar
Jewell, P. A., & Oates, J. F. (1969a). Ecological observations on the lorisoid primates of African lowland forest. Zoologica Africana, 4(2), 231248.Google Scholar
Jewell, P. A., & Oates, J. F. (1969b). Breeding activity in prosimians and small rodents in West Africa. Journal of Reproduction and Fertility, 6, 2338.Google Scholar
Ji, W., & Jiang, X. (2004). Primatology in China. International Journal of Primatology, 25(5), 10771092.Google Scholar
Jiang, Z. G., Jiang, J. P., Wang, Y. Z., Zhang, E., Zhang, Y. Y., Li, L. L., …, & Dong, L. (2016). Red list of China’s vertebrates. Biodiversity Science, 24(5), 500551.Google Scholar
Jiang, Z. G., Liu, S. Y., Wu, Y., Jiang, X. L., & Zhou, K. Y. (2017). China’s mammal diversity. Biodiversity Science, 25(8), 886895.Google Scholar
Joesch, M., & Meister, M. (2016). A neuronal circuit for colour vision based on rod–cone opponency. Nature, 532(7598), 236.Google Scholar
Johns, A. D. (1983). Ecological effects of selective logging in a West Malaysian rain-forest (unpublished doctoral dissertation). University of Cambridge, Cambridge, UK.Google Scholar
Johns, A. D. (1986). Effects of selective logging on the behavioral ecology of West Malaysian primates. Ecology, 67(3), 684694.Google Scholar
Johns, A. D., & Skorupa, J. P. (1987). Responses of rain-forest primates to habitat disturbance: A review. International Journal of Primatology, 8(2), 157.Google Scholar
Johnsen, S. (2012). The optics of life: A biologist’s guide to light in nature. Princeton, NJ: Princeton University Press.Google Scholar
Jorgensen, R. M., & Jayne, B. C. (2017). Three-dimensional trajectories affect the epaxial muscle activity of arboreal snakes crossing gaps. Journal of Experimental Biology, 220(19), 35453555.Google Scholar
Jouffroy, F. K. (1962). La Musculature des membres chez les Lémuriens de Madagascar. Paris: Mammalia.Google Scholar
Jouffroy, F. K., & Petter, A. (1990). Gravity-related kinematic changes in lorisine horizontal locomotion in relation to position of the body. In Jouffroy, F., Stack, M., & Niemitz, C. (Eds.), Gravity, posture and locomotion in primates (pp. 199208) (2nd ed.). Florence: Sedicesimo.Google Scholar
Jouffroy, F. K., & Stern, J. J. (1990). Telemetered EMG study of the antigravity versus propulsive actions of knee and elbow muscles in the slow loris (Nycticebus coucang). In Jouffroy, F., Stack, M., & Niemitz, C. (Eds.), Gravity, posture and locomotion in primates (pp. 221236) (2nd ed.). Florence: Sedicesimo.Google Scholar
Jouffroy, F. K., Renous, S., & Gasc, J. P. (1983). Etude cineradiographique des deplacements du membre anterieur du Potto de Bosman (Perodicticus potto, P .L.S. Muller, 1766) au cours de la marche quadrupede sur une branche horizontale. Annales Des Sciences Naturelles, Zoologie, Paris, 5(2), 7587.Google Scholar
Jugert, P., Eckstein, K., Noack, P., Kuhn, A., & Benbow, A. (2013). Offline and online civic engagement among adolescents and young adults from three ethnic groups. Journal of Youth and Adolescence, 42(1), 123135.Google Scholar
Jungers, W. L. (1979). Locomotion, limb proportions, and skeletal allometry in lemurs and lorises. Folia Primatologica, 32(1–2), 828.Google Scholar
Jungers, W., Stern, J. J., & Jouffroy, F. (1983). Functional morphology of the quadriceps femoris in primates: A comparative anatomical and experimental analysis. Annales Des Sciences Naturelles, Zoologie, 5, 101116.Google Scholar
Jurke, M. H., Czekala, N. M., & Fitch‐Snyder, H. (1997). Non‐invasive detection and monitoring of estrus, pregnancy and the postpartum period in pygmy loris (Nycticebus pygmaeus) using fecal estrogen metabolites. American Journal of Primatology, 41(2), 103115.Google Scholar
Jurke, M. H., Czekala, N. M., Jurke, S., Hagey, L. R., Lance, V. A., Conley, A. J., & Fitch‐Snyder, H. (1998). Monitoring pregnancy in twinning pygmy loris (Nycticebus pygmaeus) using fecal estrogen metabolites. American Journal of Primatology, 46(2), 173183.Google Scholar
Juste, J., Fa, J. E., Perez Del Val, J., & Castroviejo, J. (1995). Market dynamics of bushmeat species in Equatorial Guinea. Journal of Applied Ecology, 32, 454467.Google Scholar
Kaas, J. H. (2004). Evolution of somatosensory and motor cortex in primates. The Anatomical Record Part A, 281(1), 11481156.Google Scholar
Kamilar, J. M., & Cooper, N. (2013). Phylogenetic signal in primate behaviour, ecology and life history. Philosophical Transactions of the Royal Society B: Biological Sciences, 368(1618), 20120341.Google Scholar
Kanagasuntheram, R., & Jayawardene, F. L. W. (1957). The intrinsic muscles of the hand in the slender loris. Proceedings of the Zoological Society of London, 128, 301312.Google Scholar
Kandel, E., Schwartz, J., & Jessell, T. (2000). Principles of neural science. New York, NY: McGraw-Hill.Google Scholar
Kappeler, P. M. (1991). Patterns of sexual dimorphism in body weight among prosimian primates. Folia Primatologica, 57(3), 132146.Google Scholar
Kappeler, P. M. (1997a). Intrasexual selection and testis size in strepsirhine primates. Behavioral Ecology, 8(1), 1019.CrossRefGoogle Scholar
Kappeler, P. M. (1997b). Intrasexual selection in Mirza coquereli: Evidence for scramble competition polygyny in a solitary primate. Behavioral Ecology and Sociobiology, 41(2), 115127.Google Scholar
Kappeler, P. M. (1999). Convergence and nonconvergence in primate social systems. In Fleagle, J., Janson, C., & Reed, K. (Eds.), Primate communities (pp. 158170). Cambridge: Cambridge University Press.Google Scholar
Kappeler, P. M. (2000). Lemur origins: Rafting by groups of hibernators? Folia Primatologica, 71(6), 422425.CrossRefGoogle ScholarPubMed
Kappeler, P. M., & Erkert, H. G. (2003). On the move around the clock: Correlates and determinants of cathemeral activity in wild redfronted lemurs (Eulemur fulvus rufus). Behavioral Ecology and Sociobiology, 54(4), 359369.Google Scholar
Kappeler, P. M., & van Schaik, C. P. (2002). Evolution of primate social systems. International Journal of Primatology, 23(4), 707740.Google Scholar
Karesh, W. B., Cook, R. A., Bennett, E. L., & Newcomb, J. (2005). Wildlife trade and global disease emergence. Emerging Infectious Diseases, 11(7), 10001002.Google Scholar
Kaudern, W. (1914). Einige Beobachtungen ueber die Zeit der Fortpflanzung der madagassischen Saeugetiere. Arkiv För Zoologi (Uppsala), 9(1), 122.Google Scholar
Kawamura, S., & Kubotera, N. (2004). Ancestral loss of short wave-sensitive cone visual pigment in lorisiform prosimians, contrasting with its strict conservation in other prosimians. Journal of Molecular Evolution, 58(3), 314321.Google Scholar
Kay, R. F. (1975). The functional adaptations of primate molar teeth. American Journal of Physical Anthropology, 43(2), 195215.Google Scholar
Kay, R. F. (1984). On the use of anatomical features to infer foraging behavior in extinct primates. In Rodman, P. S. & Cant, C. G.H. (Eds.), Adaptations for foraging in non-human primates (pp. 2153). New York, NY: Columbia University Press.CrossRefGoogle Scholar
Kay, R. F., & Hylander, W. L. (1978). The dental structure of mammalian folivores with special reference to Primates and Phalangeroidea (Marsupialia). In Montgomery, G. G. (Ed.), The ecology of arboreal folivores (pp. 173191). Washington, DC: Smithsonian Institution Press.Google Scholar
Kay, R. F., & Kirk, E. C. (2000). Osteological evidence for the evolution of activity pattern and visual acuity in primates. American Journal of Physical Anthropology, 113(2), 235262.3.0.CO;2-9>CrossRefGoogle ScholarPubMed
Kay, R. F., & Sheine, W. S. (1979). On the relationship between chitin particle size and digestibility in the primate Galago senegalensis. American Journal of Physical Anthropology, 50(3), 301308.CrossRefGoogle Scholar
Kay, R. F., Ross, C., & Williams, B. A. (1997). Anthropoid origins. Science, 275(5301), 797804.Google Scholar
Kay, R. F., Campbell, V. M., Rossie, J. B., Colbert, M. W., & Rowe, T. B. (2004). Olfactory fossa of Tremacebus harringtoni (Platyrrhini, early Miocene, Sacanana, Argentina): Implications for activity pattern. The Anatomical Record Part A, 281(1), 11571172.Google Scholar
Kays, R. W. (1999). A hoistable arboreal mammal trap. Wildlife Society Bulletin (1973–2006), 27(2), 298300.Google Scholar
Kelber, A., & Roth, L. S. (2006). Nocturnal colour vision: Not as rare as we might think. Journal of Experimental Biology, 209(5), 781788.Google Scholar
Kelber, A., Yovanovich, C., & Olsson, P. (2017). Thresholds and noise limitations of colour vision in dim light. Philosophical Transactions of the Royal Society B: Biological Sciences, 372(1717), 20160065.Google Scholar
Kellert, S. R. (1984). Attitudes toward animals: Age-related development among children. In Fox, M. W. & Mickley, L. D. (Eds.), Advances in animal welfare science 1984 (pp. 4360). Dordrecht: Springer.Google Scholar
Kellert, S. R., & Berry, J. K. (1987). Attitudes, knowledge, and behaviors toward wildlife as affected by gender. Wildlife Society Bulletin (1973–2006), 15(3), 363371.Google Scholar
Kemble, E. D., & Lewis, C. (1982). Effects of vibrissal amputation on cricket predation in northern grasshopper mice (Onychomys leucogaster). Bulletin of the Psychonomic Society, 20(5), 275276.Google Scholar
Kenward, R. E. (2000). A manual for wildlife radio tagging. London: Academic Press.Google Scholar
Kenyon, M., Streicher, U., Loung, H., Tran, T., Tran, M., Vo, B., & Cronin, A. (2014). Survival of reintroduced pygmy slow loris Nycticebus pygmaeus in South Vietnam. Endangered Species Research, 25(2), 185195.Google Scholar
Keylock, N. (2002). The importance of hunting for bushmeat to a rural community in Equatorial Guinea (Master’s dissertation). University of London, London, UK.Google Scholar
Khan, M. A. R. (2015). Wildlife of Bangladesh: Checklist and guide. Dhaka: Chayabithi Publisher.Google Scholar
Kietzmann, J. H., Hermkens, K., McCarthy, I. P., & Silvestre, B. S. (2011). Social media? Get serious! Understanding the functional building blocks of social media. Business Horizons, 54(3), 241251.Google Scholar
Kimura, T. (1985). Bipedal and quadrupedal walking of primates: Comparative dynamics. In Kondon, S. (Ed.), Primate morphophysiology, locomotor analyses and human bipedalism (pp. 81104). Tokyo: University of Tokyo Press.Google Scholar
Kimura, T. (1992). Hindlimb dominance during primate high-speed locomotion. Primates, 33(4), 465476.CrossRefGoogle Scholar
Kimura, T., Okada, M., & Ishida, H. (1979). Kinesiological characteristics of primate walking: Its significance in human walking. In Morbeck, M. E., Preuschoft, H., & Gomberg, N. (Eds.), Environment, behavior, and morphology: Dynamic interactions in primates (pp. 297311). New York, NY: Gustav Fischer.Google Scholar
King, J. O., & King, D. T. (1994). In my experience: Use of a long-distance night vision device for wildlife studies. Wildlife Society Bulletin (1973–2006), 22(1), 121125.Google Scholar
Kingdon, J. (1971). East African mammals: An atlas of evolution in Africa, vol. 1. London: Academic Press.Google Scholar
Kingdon, J. (1997). The Kingdon field guide to African mammals. San Diego, CA: Academic Press.Google Scholar
Kingdon, J. (2015). The Kingdon field guide to African mammals. London: Bloomsbury Publishing.Google Scholar
Kingston, A. K., Boyer, D. M., Patel, B. A., Larson, S. G., & Stern, J. T. Jr (2010). Hallucal grasping in Nycticebus coucang: Further implications for the functional significance of a large peroneal process. Journal of Human Evolution, 58(1), 3342.CrossRefGoogle ScholarPubMed
Kirk, E. C. (2003). Evolution of the primate visual system (unpublished doctoral dissertation). Duke University, Durham, NC, USA.Google Scholar
Kirk, E. C. (2004). Comparative morphology of the eye in primates. The Anatomical Record Part A, 281(1), 10951103.Google Scholar
Kirk, E. C. (2006a). Effects of activity pattern on eye size and orbital aperture size in primates. Journal of Human Evolution, 51(2), 159170.Google Scholar
Kirk, E. C. (2006b). Visual influences on primate encephalization. Journal of Human Evolution, 51(1), 7690.Google Scholar
Kirk, E. C. (2013). Characteristics of crown primates. Nature Education Knowledge, 4(8), 3.Google Scholar
Kirk, E. C., & Kay, R. (2004). The evolution of high visual acuity in the Anthropoidea. In Ross, C. & Kay, R. (Eds.), Anthropoid origins: New visions (pp. 539602). New York, NY: Kluwer Academic/Plenum.Google Scholar
Kirk, E. C., & Simons, E. L. (2001). Diets of fossil primates from the Fayum Depression of Egypt: A quantitative analysis of molar shearing. Journal of Human Evolution, 40(3), 203229.Google Scholar
Kitson, H., & Nekaris, K. A. I. (2017). Instagram-fuelled illegal slow loris trade uncovered in Marmaris, Turkey. Oryx, 51, 4.Google Scholar
Klerman, E. B., Wang, W., Duffy, J. F., Dijk, D. J., Czeisler, C. A., & Kronauer, R. E. (2013). Survival analysis indicates that age-related decline in sleep continuity occurs exclusively during NREM sleep. Neurobiology of Aging, 34(1), 309318.Google Scholar
Kling, K. J., & Hopkins, M. E. (2015). Are we making the grade? Practices and reported efficacy measures of primate conservation education programs. American Journal of Primatology, 77(4), 434448.Google Scholar
Klouček, O. (2017). CITES – základní informace. Ministerstvo životního prostředí. Retrieved 21 December 2017, from www.mzp.cz/C1257458002F0DC7/cz/cites_obchod_ohrozenymi_druhy/$FILE/ODOIMZ-CITES_ZAKLADNI_INFO_171221.pdf.Google Scholar
Kobbe, S., Ganzhorn, J. U., & Dausmann, K. H. (2011). Extreme individual flexibility of heterothermy in free-ranging Malagasy mouse lemurs (Microcebus griseorufus). Journal of Comparative Physiology B, 181(1), 165173.Google Scholar
Kochanny, C. O., Delgiudice, G. D., & Fieberg, J. (2009). Comparing global positioning system and very high frequency telemetry home ranges of white‐tailed deer. Journal of Wildlife Management, 73(5), 779787.Google Scholar
Kohlberg, L. (1975). The cognitive-developmental approach to moral education. The Phi Delta Kappan, 56(10), 670677.Google Scholar
Köhler, J. (2017). Grey slender loris (Loris lydekkerianus grandis) EEP studbook. Frankfurt: Frankfurt Zoo.Google Scholar
Koli, V. K., & Bhatnagar, C. (2016). Seasonal variation in the activity budget of Indian giant flying squirrel (Petaurista philippensis) in tropical deciduous forest, Rajasthan, India. Folia Zoologica, 65(1), 3846.Google Scholar
Kosinski, M., Matz, S. C., Gosling, S. D., Popov, V., & Stillwell, D. (2015). Facebook as a research tool for the social sciences: Opportunities, challenges, ethical considerations, and practical guidelines. American Psychologist, 70(6), 543556.Google Scholar
Koufos, G. D., Zouros, N., & Mourouzidou, O. (2003). Prodeinotherium bavaricum (Proboscidea, Mammalia) from Lesvos island, Greece: The appearance of deinotheres in the Eastern Mediterranean. Geobios, 36(3), 305315.Google Scholar
Koufos, G. D., Kostopoulos, D. S., & Vlachou, T. D. (2005). Neogene/Quaternary mammalian migrations in eastern Mediterranean. Belgian Journal of Zoology, 135(2), 181.Google Scholar
Krane, S., Itagaki, Y., Nakanishi, K., & Weldon, P. J. (2003). Venom of the slow loris: Sequence similarity of prosimian skin gland protein and Fel d 1 cat allergen. Naturwissenschaften, 90(2), 6062.CrossRefGoogle Scholar
Krause, W. J. (2010). Morphological and histochemical observations on the crural gland-spur apparatus of the echidna (Tachyglossus aculeatus) together with comparative observations on the femoral gland-spur apparatus of the duckbilled platypus (Ornithorhyncus anatinus). Cells Tissues Organs, 191(4), 336354.Google Scholar
Kretschmer, F., Sajgo, S., Kretschmer, V., & Badea, T. C. (2015). A system to measure the optokinetic and optomotor response in mice. Journal of Neuroscience Methods, 256, 91105.Google Scholar
Kuhar, C. W., Bettinger, T. L., Lehnhardt, K., Tracy, O., & Cox, D. (2010). Evaluating for long‐term impact of an environmental education program at the Kalinzu Forest Reserve, Uganda. American Journal of Primatology, 72(5), 407413.Google Scholar
Kumara, H. N., & Radhakrishna, S. (2013). Evaluation of census techniques to estimate the density of slender Loris (Loris lydekkerianus) in Southern India. Current Science, 104, 10831086.Google Scholar
Kumara, H. N., Kumar, S., & Singh, M. (2005). A novel foraging technique observed in slender loris (Loris lydekkerianus malabaricus) feeding on red ants in the Western Ghats, India. Folia Primatologica, 76(2), 116118.Google Scholar
Kumara, H. N., Singh, M., & Kumar, S. (2006). Distribution, habitat correlates, and conservation of Loris lydekkerianus in Karnataka, India. International Journal of Primatology, 27(4), 941969.Google Scholar
Kumara, H. N., Sasi, R., Suganthasakthivel, R., & Srinivas, G. (2011). Distribution, abundance and conservation of primates in Highwavy Mountains of Western Ghats, Tamil Nadu, India. Current Science, 100(7), 10631067.Google Scholar
Kumara, H. N., Singh, M., Irfan-Ullah, M., & Kumar, S. (2013). Status, distribution, and conservation of slender lorises in India. In Masters, J. C., Gamba, M., & Génin, F. (Eds.), Leaping ahead: Advances in prosimian biology (pp. 343352). New York, NY: Springer Science + Business Media.Google Scholar
Kumara, H. N., Sasi, R., Chandran, S., & Radhakrishna, S. (2016). Distribution of the grey slender loris (Loris lyddekerianus Cabrera, 1908) in Tamil Nadu, southern India. Folia Primatologica, 87(5), 291302.Google Scholar
Kumaran, J. V., Khan, F. A. A., Azhar, I., Wee Chen, E., Ali, M. R. M., Ahmad, A., & Yusoff, A. M. (2016). Diversity and conservation status of small mammals in Kelantan, Malaysia. Songklanakarin Journal of Science & Technology, 38(2).Google Scholar
Kümpel, N. F., Milner-Gulland, E. J., Rowcliffe, J. M., & Cowlishaw, G. (2008). Impact of gun-hunting on diurnal primates in continental Equatorial Guinea. International Journal of Primatology, 29(4), 10651082.CrossRefGoogle Scholar
Kunimatsu, Y., Nakatsukasa, M., Sawada, Y., Sakai, T., Hyodo, M., Hyodo, H., …, & Saneyoshi, M. (2007). A new Late Miocene great ape from Kenya and its implications for the origins of African great apes and humans. Proceedings of the National Academy of Sciences, 104(49), 1922019225.Google Scholar
Kunimatsu, Y., Tsujikawa, H., Nakatsukasa, M., Shimizu, D., Ogihara, N., Kikuchi, Y., …, & Ishida, H. (2017). A new species of Mioeuoticus (Lorisiformes, Primates) from the early Middle Miocene of Kenya. Anthropological Science. DOI: http://doi.org/10.1537/ase.170322.Google Scholar
Kutsukake, N. (2013). Heterogeneous ethical structures in field primatology. Ethics in the Field: Contemporary Challenges, 7, 84.Google Scholar
Lacépède, B. G. E. (1800a). Classification des oiseaux et des mammifières. Séances Des Écoles Normales, Recueillies Par Des Sténographes, Et Revues Par Les Professeurs, 9 (appendix 1), 186.Google Scholar
Lacepède, B. G. E.de. (1800b). Séances des écoles normales, recueillies par les sténographes, et revues par les professeurs (8th ed.). Paris: Cercle-Social.Google Scholar
Lacey, E. A., & Sherman, P. W. (2007). The ecology of sociality in rodents. In Wolff, J. O. & Sherman, P. W. (Eds.), Rodent societies: An ecological and evolutionary perspective (pp. 243254). Chicago, IL: University of Chicago Press.Google Scholar
Lambert, J. E. (2007). Seasonality, fallback strategies, and natural selection: A chimpanzee and cercopithecoid model for interpreting the evolution of the hominin diet. In Ungar, P. S. (Ed.), Evolution of the human diet (pp. 324333). New York, NY: Oxford University Press.Google Scholar
Lambert, J. E. (2014). Handbook of the mammals of the world: 3. primates. Journal of Mammalogy, 95(4), 906907.Google Scholar
Lan, D. Y. (1999). Diversity and conservation of slow loris in Yunnan, China. Tigerpaper (Bangkok), 26(4), 1215.Google Scholar
Lancendorfer, K. M., Atkin, J. L., & Reece, B. B. (2008). Animals in advertising: Love dogs? Love the ad!. Journal of Business Research, 61(5), 384391.Google Scholar
Lancia, R. A., Nichols, J. D., & Pollock, K. H. (1994). Estimating the number of animals in wildlife populations. In Bookhout, T. A. (Ed.), Research and management techniques for wildlife and habitats (pp. 215253). Bethesda, MD: The Wildlife Society.Google Scholar
Lanum, J. (1978). The damaging effects of light on the retina: Empirical findings, theoretical and practical implications. Survey of Ophthalmology, 22(4), 221249.Google Scholar
Larney, E., & Larson, S. G. (2004). Compliant walking in primates: Elbow and knee yield in primates compared to other mammals. American Journal of Physical Anthropology, 125(1), 4250.Google Scholar
Larson, S. G. (1998). Unique aspects of quadrupedal locomotion in non-human primates. In Strasser, E., Fleagle, J. G., Rosenberger, A., & McHenry, H. (Eds.), Primate locomotion: Recent advances (pp. 157173). Boston, MA: Springer.CrossRefGoogle Scholar
Larson, S. G., & Demes, B. (2011). Weight support distribution during quadrupedal walking in Ateles and Cebus. American Journal of Physical Anthropology, 144(4), 633642.Google Scholar
Larson, S. G., Schmitt, D., Lemelin, P., & Hamrick, M. (2000). Uniqueness of primate forelimb posture during quadrupedal locomotion. American Journal of Physical Anthropology, 112(1), 87101.Google Scholar
Larson, S. G., Schmitt, D., Lemelin, P., & Hamrick, M. (2001). Limb excursion during quadrupedal walking: How do primates compare to other mammals?. Journal of Zoology, 255(3), 353365.Google Scholar
Laska, M., Freist, P., & Krause, S. (2007). Which senses play a role in non-human primate food selection? A comparison between squirrel monkeys and spider monkeys. American Journal of Primatology, 69(3), 282294.Google Scholar
Lavorgna, A. (2014). Wildlife trafficking in the internet age. Crime Science, 3(1), 5.Google Scholar
Le Gros Clark, W. E. (1956). A Miocene lemuroid skull from East Africa. In Leakey, L. S. B. (Ed.), Fossil mammals of Africa (pp. 19). London: British Museum (Natural History).Google Scholar
Le Gros Clark, W. E. (1959). The antecedents of man. Edinburgh: Edinburgh University Press.Google Scholar
Le Gros Clark, W. E. (1971). The antecedents of man (3rd ed.). Chicago, IL: Quadrangle Books.Google Scholar
Leakey, L. S. (1961). A new lower Pliocene fossil primate from Kenya. Annals and Magazine of Natural History, 4(47), 689696.Google Scholar
Leakey, L. S. (1962). Primates. In Bishop, W. W. (Ed.), The mammalian fauna and geomorphological relations of the Napak Volcanics, Karamoja (pp. 118). Kampala: Records of the Geological Survey of Uganda 1957–1958.Google Scholar
Leakey, M. G., Ungar, P. S., & Walker, A. (1995). A new genus of large primate from the late Oligocene of Lothidok, Turkana District, Kenya. Journal of Human Evolution, 28(6), 519531.CrossRefGoogle Scholar
Leakey, M. G., Grossman, A., Gutiérrez, M., & Fleagle, J. G. (2011). Faunal change in the Turkana Basin during the late Oligocene and Miocene. Evolutionary Anthropology: Issues, News, and Reviews, 20(6), 238253.Google Scholar
Leakey, R. E., & Leakey, M. G. (1986). A second new Miocene hominoid from Kenya. Nature, 324(6093), 146148.Google Scholar
Leat, S. J., Yadav, N. K., & Irving, E. L. (2009). Development of visual acuity and contrast sensitivity in children. Journal of Optometry, 2(1), 1926.CrossRefGoogle Scholar
Leggett, H. C., El Mouden, C., Wild, G., & West, S. (2011). Promiscuity and the evolution of cooperative breeding. Proceedings of the Royal Society B: Biological Sciences, 279(1732), 14051411.Google Scholar
Lehtinen, J. (2013). Distribution of the Javan slow loris (Nycticebus javanicus): Assessing the presence in East Java, Indonesia (unpublished Master’s dissertation). Oxford Brookes University, Oxford, UK.Google Scholar
Leighty, K. A., Valuska, A. J., Grand, A. P., Bettinger, T. L., Mellen, J. D., Ross, S. R., …, & Ogden, J. J. (2015). Impact of visual context on public perceptions of non-human primate performers. PLoS One, 10(2), e0118487.Google Scholar
Leischner, C. L., Crouch, M., Allen, K. L., Marchi, D., Pastor, F., & Hartstone‐Rose, A. (2018). Scaling of primate forearm muscle architecture as it relates to locomotion and posture. The Anatomical Record, 301(3), 484495.Google Scholar
Lemelin, P., & Diogo, R. (2016). Anatomy, function, and evolution of the primate hand musculature. In Kivell, T. L., Lemelin, P., Richmond, B. G., & Schnitt, D. (Eds.), Evolution of the primate hand (pp. 155193). New York, NY: Springer.CrossRefGoogle Scholar
Lemelin, P., & Jungers, W. L. (2007). Body size and scaling of the hands and feet of prosimian primates. American Journal of Physical Anthropology, 133(2), 828840.Google Scholar
Lemelin, P., & Schmitt, D. (1998). The relation between hand morphology and quadrupedalism in primates. American Journal of Physical Anthropology, 105(2), 185197.Google Scholar
Levy, D. (2011). Participatory eco-drama. Green Teacher, 91, 4043.Google Scholar
Lewis, G. E. (1933). Preliminary notice of a new genus of lemuroid from the Siwaliks. American Journal of Science, 152, 134138.Google Scholar
Lewis, P. O. (2001). A likelihood approach to estimating phylogeny from discrete morphological character data. Systematic Biology, 50(6), 913925.CrossRefGoogle ScholarPubMed
Li, W., & Wang, H. (1999). Wildlife trade in Yunnan Province, China at the border with Vietnam. Traffic Bulletin, 18(1), 2130.Google Scholar
Lieberman, D. E., Ross, C. F., & Ravosa, M. J. (2000). The primate cranial base: Ontogeny, function and integration. Yearbook of Physical Anthropology, 43, 117169.Google Scholar
Lillywhite, H. B., Lafrentz, J. R., Lin, Y. C., & Tu, M. C. (2000). The cantilever abilities of snakes. Journal of Herpetology, 34(4), 523528.Google Scholar
Lim, B. L., Muul, I., & Langham, N. P. E. (1971). Preliminary studies of small mammals collected from Penang Island, Malaysia. Federated Malay States Museums Journal, 16, 6174.Google Scholar
Linder, J. M., & Oates, J. F. (2011). Differential impact of bushmeat hunting on monkey species and implications for primate conservation in Korup National Park, Cameroon. Biological Conservation, 144(2), 738745.Google Scholar
Linder, J. M., Sawyer, S. C., & Brashares, J. S. (2013). Primates in trade. In Sterling, E., Bynum, N., & Blair, M. (Eds.), Primate conservation and ecology: A handbook of techniques (pp. 323345). Oxford: Oxford University Press.Google Scholar
Linnaeus, C. (1735). Systema naturae. Leiden: Lugundi Batovorum (Haak).Google Scholar
Linnaeus, C. (1758a). Systema naturae (10th ed.). Stockholm: Holmiae.Google Scholar
Linnaeus, C. (1758b). Systema naturae per regna tria naturae, secundum classes, ordines, genera, species, cum characteribus, differentiis, synonymis, locis. Stockholm: Laurentii Salvii.Google Scholar
Liu, X., Feng, Z., Jiang, L., Li, P., Liao, C., Yang, Y., & You, Z. (2013). Rubber plantation and its relationship with topographical factors in the border region of China, Laos and Myanmar. Journal of Geographical Sciences, 23(6), 10191040.CrossRefGoogle Scholar
Livingstone, S., & Brake, D. R. (2010). On the rapid rise of social networking sites: New findings and policy implications. Children & Society, 24(1), 7583.Google Scholar
López-Torres, S., Schillaci, M. A., & Silcox, M. T. (2015). Life history of the most complete fossil primate skeleton: Exploring growth models for Darwinius. Royal Society Open Science, 2(9), 150340.Google Scholar
López-Torres, S., Selig, K. R., Prufrock, K. A., Lin, D., & Silcox, M. T. (2018). Dental topographic analysis of paromomyid (Plesiadapiformes, Primates) cheek teeth: More than 15 million years of changing surfaces and shifting ecologies. Historical Biology, 30(1–2), 7688.Google Scholar
Lu, X., Wang, Y., & Zhang, Y. (2001). Divergence and phylogeny of mitochondrial cytochrome b gene from species in Nycticebus. Zoological Research, 22(2), 9398.Google Scholar
Lu, X. M., Fu, Y. X., & Zhang, Y. P. (2002). Evolution of mitochondrial cytochrome b pseudogene in genus Nycticebus. Molecular Biology and Evolution, 19(12), 23372341.Google Scholar
Luhrs, A. M., Svensson, M. S., & Nekaris, K. A. I. (2018). Comparative ecology and behaviour of eastern potto Perodicticus ibeanus and central potto P. edwardsi in Angola, Cameroon, Kenya, Nigeria, Rwanda and Uganda. Journal of East African Natural History, 107(1), 1730.CrossRefGoogle Scholar
Lukas, D., & Clutton-Brock, T. H. (2013). The evolution of social monogamy in mammals. Science, 341(6145), 526530.Google Scholar
Lukas, D., & Clutton-Brock, T. (2014). Evolution of social monogamy in primates is not consistently associated with male infanticide. Proceedings of the National Academy of Sciences, 111(17), E1674E1674.Google Scholar
Lukas, D., & Huchard, E. (2014). The evolution of infanticide by males in mammalian societies. Science, 346(6211), 841844.Google Scholar
Lundeen, I. K., & Kirk, E. C. (2019). Internal nasal morphology of the Eocene primate Rooneyia viejaensis and extant Euarchonta: Using µCT scan data to understand and infer patterns of nasal fossa evolution in primates. Journal of Human Evolution, 132: 137173.Google Scholar
Lydekker, R. (1893). Mammalia. Zoological Record, 29, 2425.Google Scholar
Lydekker, R. (1904). On two lorises. Proceedings of the Zoological Society of London, 74: 345346.Google Scholar
Lyon, M. W. Jr. (1906). Notes on the slow lemurs. Proceedings of the United States National Museum, 31, 527539.Google Scholar
Lythgoe, J. (1979). The ecology of vision. Oxford: Oxford University Press.Google Scholar
Ma, S. L., & Wang, Y. X. (1988). The recent distribution, status and conservation of primates in China. Acta Theriologica Sinica, 8(4), 250260.Google Scholar
Ma, S. L., Han, L., Lan, D., Ji, W., & Harris, R. B. (1995). Faunal resources of the Gaoligongshan region of Yunnan, China: Diverse and threatened. Environmental Conservation, 3(22), 250258.Google Scholar
Mace, G. M. (2004). The role of taxonomy in species conservation. Philosophical Transactions of the Royal Society B: Biological Sciences, 359(1444), 711719.Google Scholar
Machado, C. J. (2013). Maternal influences on social and neural development in macaque monkeys. In Clancy, K. B. H., Hinde, K., & Rutherford, J. N. (Eds.), Building babies: Primate development in proximate and ultimate perspective (pp. 259280). New York, NY: Springer.Google Scholar
MacKenzie, D. I. (2005). What are the issues with presence–absence data for wildlife managers?. Journal of Wildlife Management, 69(3), 849860.Google Scholar
MacKenzie, D. I., & Reardon, J. T. (2013). Occupancy methods for conservation management. In Collen, B., Pettorelli, N., Baillie, J. E. M., & Durant, S. M. (Eds.), Biodiversity monitoring and conservation: Bridging the gap between global commitment and local action (pp. 247264). New York, NY: Wiley-Blackwell.Google Scholar
MacKenzie, D. I., & Royle, J. A. (2005). Designing occupancy studies: General advice and allocating survey effort. Journal of Applied Ecology, 42(6), 11051114.CrossRefGoogle Scholar
MacKenzie, D. I., Nichols, J. D., Lachman, G. B., Droege, S., Andrew Royle, J., & Langtimm, C. A. (2002). Estimating site occupancy rates when detection probabilities are less than one. Ecology, 83(8), 22482255.Google Scholar
MacKenzie, D. I., Royle, J. A., Brown, J. A., & Nichols, J. D. (2004). Occupancy estimation and modeling for rare and elusive populations. In Thompson, W. L. (Ed.), Sampling rare or elusive species: Concepts, designs, and techniques for estimating population parameters (pp. 149171). London: Island Press.Google Scholar
MacKenzie, D. I., Nichols, J. D., Royle, J. A., Pollock, K. H., Bailey, L., & Hines, J. E. (2006). Occupancy estimation and modeling: Inferring patterns and dynamics of species occurrence. Burlington, MA: Academic Press.Google Scholar
Maclatchy, L. (2004). The oldest ape. Evolutionary Anthropology: Issues, News, and Reviews, 13(3), 90103.Google Scholar
MacLatchy, L., & Rossie, J. B. (2005). The Napak hominoid: Still Proconsul major. In Lieberman, D. E., Smith, R. J., & Kelley, J. (Eds.), Interpreting the past: Essays on human, primate, and mammal evolution in honour of David Pilbeam (pp. 1528). Boston, MA: Brill Academic Publishers.Google Scholar
MacPhee, R. D. E., & Jacobs, L. L. (1986). Nycticeboides simpsoni and the morphology, adaptations, and relationships of Miocene Siwalik Lorisidae. Rocky Mountain Geology, 24, 131161.Google Scholar
Madani, G., & Nekaris, K. A. I. (2014). Anaphylactic shock following the bite of a wild Kayan slow loris (Nycticebus kayan): Implications for slow loris conservation. Journal of Venomous Animals and Toxins Including Tropical Diseases20, 43.Google Scholar
Maddison, W., & Maddison, D. (2007). Mesquite 2: A modular system for evolutionary analysis. Retrieved 13 September 2019, from http://andrelevy.net/bioinfo/mesquite/curso_mesquite/Mesquite2Manual.pdf.Google Scholar
Maestripieri, D., & Carroll, K. A. (1998). Behavioral and environmental correlates of infant abuse in group-living pigtail macaques. Infant Behavior and Development, 21(4), 603612.Google Scholar
Maier, W., & Ruf, I. (2014). Morphology of the nasal capsule of primates: With special reference to Daubentonia and Homo. The Anatomical Record, 297(11), 19852006.Google Scholar
Maiolino, S., Boyer, D. M., Bloch, J. I., Gilbert, C. C., & Groenke, J. (2012). Evidence for a grooming claw in a North American adapiform primate: Implications for anthropoid origins. PLoS One, 7(1), e29135.Google Scholar
Manduell, K. L., Harrison, M. E., & Thorpe, S. K. (2012). Forest structure and support availability influence orangutan locomotion in Sumatra and Borneo. American Journal of Primatology, 74(12), 11281142.Google Scholar
Manley, G. H. (1966). Reproduction in lorisoid primates. Symposium Zoological Society of London, 15, 493509.Google Scholar
Manley, G. H. (1967). Gestation periods in the LorisidaeInternational Zoo Yearbook7(1), 8081.Google Scholar
Manley, G. H. (1974). Functions of the external genital glands of Perodicticus and Arctocebus. In Doyle, G. A., Martin, R. D., & Walker, A. C. (Eds.), Prosimian Biology (pp. 313329). London: Duckworth.Google Scholar
Mann, T. M., Williams, K. E., Pearce, P. C., & Scott, E. A. M. (2005). A novel method for activity monitoring in small non-human primates. Laboratory Animals, 39(2), 169177.Google Scholar
Marchi, R. (2012). With Facebook, blogs, and fake news, teens reject journalistic ‘objectivity’. Journal of Communication Inquiry, 36(3), 246262.CrossRefGoogle Scholar
Marivaux, L., Welcomme, J. L., Antoine, P. O., Métais, G., Baloch, I. M., Benammi, M., …, & Jaeger, J. J. (2001). A fossil lemur from the Oligocene of Pakistan. Science, 294(5542), 587591.Google Scholar
Marivaux, L., Ramdarshan, A., Essid, E. M., Marzougui, W., Ammar, H. K., Lebrun, R., …, & Vianey-Liaud, M. (2013). Djebelemur, a tiny pre-tooth-combed primate from the Eocene of Tunisia: A glimpse into the origin of crown strepsirhines. PLoS One, 8(12), e80778.Google Scholar
Marshall, A. J., & Wrangham, R. W. (2007). Evolutionary consequences of fallback foods. International Journal of Primatology, 28(6), 1219.Google Scholar
Marshall, C. D., Rozas, K., Kot, B., & Gill, V. A. (2014). Innervation patterns of sea otter (Enhydra lutris) mystacial follicle–sinus complexes. Frontiers in Neuroanatomy, 8, 121.Google Scholar
Marshall, M. N. (1996). Sampling for qualitative research. Family Practice, 13(6), 522526.Google Scholar
Martin, E. B., & Phipps, M. (1996). A review of the wild animal trade in Cambodia. TRAFFIC Bulletin: Wildlife Trade Monitoring Unit, 16, 4560.Google Scholar
Martín, J. (2001). When hiding from predators is costly: Optimization of refuge use in lizards. Etologia, 9, 913.Google Scholar
Martin, P., & Bateson, P. (2007). Measuring behaviour: An introductory guide (3rd ed.). Cambridge: Cambridge University Press.Google Scholar
Martin, R. D. (1972). Review lecture: Adaptive radiation and behaviour of the Malagasy lemurs. Philosophical Transactions of the Royal Society B: Biological Sciences, 264(862), 295352.Google Scholar
Martin, R. D. (1975). The bearing of reproductive behaviour and ontogeny on Strepsirhine phylogeny. In Luckett, W. & Szalay, F. (Eds.), Phylogeny of the primates (pp. 265297). Boston, MA: Springer.Google Scholar
Martin, R. D. (1979). Phylogenetic aspects of prosimian behavior. In Doyle, G. A. and Martin, R. D. (Eds.), The study of prosimian behavior (pp. 4578). New York, NY: Academic Press.Google Scholar
Martin, R. D. (1981). Field studies of primate behaviour. Symposia of the Zoological Society of London, 46, 287336.Google Scholar
Martin, R. D. (1990). Primate origins and evolution: A phylogenetic reconstruction. London: Chapman and Hall.Google Scholar
Martin, R. D. (1993). Primate origins: Plugging the gaps. Nature, 363(6426), 223.CrossRefGoogle ScholarPubMed
Martin, R. D. (1995). Prosimians: From obscurity to extinction?. In Alterman, L., Doyle, G. A., & Izard, M. K. (Eds.), Creatures of the dark: The nocturnal prosimians (pp. 535563). Boston, MA: Springer.Google Scholar
Martin, R. D. (2000). Origins, diversity and relationships of lemurs. International Journal of Primatology, 21(6), 10211049.Google Scholar
Martin, R. D. (2003). Combining the primate record. Nature, 422, 388391.Google Scholar
Martin, R. D. (2010). Foreword. In Burrows, A. M. & Nash, L. T. (Eds.), The Evolution of exudativory in primates (pp. 714). New York, NY: Springer.Google Scholar
Martínez-García, F., Martínez-Ricós, J., Agustín-Pavón, C., Martínez-Hernández, J., Novejarque, A., & Lanuza, E. (2009). Refining the dual olfactory hypothesis: Pheromone reward and odour experience. Behavioural Brain Research, 200(2), 277286.Google Scholar
Masters, J. C., & Brothers, D. J. (2002). Lack of congruence between morphological and molecular data in reconstructing the phylogeny of the Galagonidae. American Journal of Physical Anthropology, 117(1), 7993.CrossRefGoogle ScholarPubMed
Masters, J. C., & Couette, S. (2015). Characterizing cryptic species: A morphometric analysis of craniodental characters in the dwarf galago genus Galagoides. American Journal of Physical Anthropology, 158 (2): 288299.Google Scholar
Masters, J. C., Lumsden, W. H. R., & Young, D. A. (1988). Reproductive and dietary parameters in wild greater galago populations. International Journal of Primatology, 9(6), 573.Google Scholar
Masters, J. C., Anthony, N. M., De Wit, M. J., & Mitchell, A. (2005). Reconstructing the evolutionary history of the Lorisidae using morphological, molecular, and geological data. American Journal of Physical Anthropology, 127(4), 465480.Google Scholar
Masters, J. C., Boniotto, M., Crovella, S., Roos, C., Pozzi, L., & Delpero, M. (2007). Phylogenetic relationships among the Lorisoidea as indicated by craniodental morphology and mitochondrial sequence data. American Journal of Primatology, 69(1), 615.Google Scholar
Masters, J. C., Silvestro, D., Génin, F., & DelPero, M. (2013). Seeing the wood through the trees: The current state of higher systematics in the Strepsirhini. Folia Primatologica, 84(3–5), 201219.CrossRefGoogle ScholarPubMed
Masters, J. C., Génin, F., Couette, S., Groves, C. P., Nash, S. D., Delpero, M., & Pozzi, L. (2017a). A new genus for the eastern dwarf galagos (Primates: Galagidae). Zoological Journal of the Linnean Society, 181(1), 229241.Google Scholar
Masters, J. C., Pozzi, L., & Godfrey, L. (2017b). Evolution of modern Strepsirhines. In Fuentes, A. (Ed.), The international encyclopedia of primatology (pp. 15). Oxford: Wiley-Blackwell.Google Scholar
Mastorakos, G., & Ilias, I. (2003). Maternal and fetal hypothalamic–pituitary–adrenal axes during pregnancy and postpartum. Annals of the New York Academy of Sciences, 997(1), 136149.Google Scholar
Matsui, A., Rakotondraparany, F., Munechika, I., Hasegawa, M., & Horai, S. (2009). Molecular phylogeny and evolution of prosimians based on complete sequences of mitochondrial DNAs. Gene, 441(1–2), 5366.Google Scholar
Matsui, A., Go, Y., & Niimura, Y. (2010). Degeneration of olfactory receptor gene repertories in primates: No direct link to full trichromatic vision. Molecular Biology and Evolution, 27(5), 11921200.Google Scholar
Matsumoto, Y., Hiramatsu, C., Matsushita, Y., Ozawa, N., Ashino, R., Nakata, M., …, & Melin, A. D. (2014). Evolutionary renovation of L/M opsin polymorphism confers a fruit discrimination advantage to ateline New World monkeys. Molecular Ecology, 23(7), 17991812.Google Scholar
Mattson, E. E., & Marshall, C. D. (2016a). Follicle microstructure and innervation vary between pinniped micro-and macrovibrissae. Brain, Behavior and Evolution, 88(1), 4358.Google Scholar
Mattson, E. E., & Marshall, C. D. (2016b). Innervational and microanatomical support for functional compartmentalization within the mystacial vibrissal sensory system of pinnipeds. Integrative and Comparative Biology, 56, E141E141.Google Scholar
Mauro, A. A., & Jayne, C. B. (2016). Perch compliance and experience affect destination choice of brown tree snakes (Boiga irregularis). Zoology, 119(2), 113118.Google Scholar
Mbete, R. A., Banga-Mboko, H., Racey, P., Mfoukou-Ntsakala, A., Nganga, I., Vermeulen, C., …, & Leroy, P. (2011). Household bushmeat consumption in Brazzaville, the Republic of the Congo. Tropical Conservation Science, 4(2), 187202.Google Scholar
McCabe, S., & Nekaris, K. A. I. (2018). The impact of subtle anthropomorphism on gender differences in learning conservation ecology in Indonesian school children. Applied Environmental Education & Communication, 18(1), 112.Google Scholar
McCann, C., Buchanan-Smith, H. M., Jones-Engel, L., Farmer, K. H., Prescott, M. J., Fitch-Snyder, H., & Taylor, S. (2007). IPS International guidelines for the acquisition, care and breeding of non-human primates. n.p.: International Primatological Society, USA.Google Scholar
McCauley, D. J. (2006). Selling out on nature. Nature, 443(7107), 27.CrossRefGoogle ScholarPubMed
McConkey, K. R., & O’Farrill, G. (2016). Loss of seed dispersal before the loss of seed dispersers. Biological Conservation, 201, 3849.Google Scholar
McCrossin, M. L. (1992). New species of bushbaby from the middle Miocene of Maboko Island, KenyaAmerican Journal of Physical Anthropology89(2), 215233.Google Scholar
McCullagh, P., & Nelder, J. A. (1989). Generalized linear models (2nd ed.). London: Chapman & Hall.CrossRefGoogle Scholar
McDonald, L. L. (2004). Sampling rare populations. In Thompson, W. L. (Ed.), Sampling rare or elusive species (pp. 1142). Washington, DC: Island Press.Google Scholar
McKenna, M. C., & Bell, S. K. (1997). Classification of mammals above the species level. New York, NY: Columbia University Press.Google Scholar
McNaughton, M. J. (2004). Educational drama in the teaching of education for sustainability. Environmental Education Research, 10(2), 139155.Google Scholar
Meijide, A., Badu, C. S., Moyano, F., Tiralla, N., Gunawan, D., & Knohl, A. (2018). Impact of forest conversion to oil palm and rubber plantations on microclimate and the role of the 2015 ENSO event. Agricultural and Forest Meteorology, 252, 208219.Google Scholar
Mein, P., & Ginsburg, L. (1997). Les mammifères du gisement miocène inférieur de Li Mae Long, Thaïlande: Systématique, biostratigraphie et paléoenvironnement. Geodiversitas, 19(4), 783844.Google Scholar
Meisami, E., & Bhatnagar, K. P. (1998). Structure and diversity in mammalian accessory olfactory bulb. Microscopy Research and Technique, 43(6), 476499.Google Scholar
Melcón, M. L., & Moss, C. F. (2013). Introduction to special issue,’How nature shaped echolocation in animals’. Frontiers in Physiology, 4, 193.Google Scholar
Melin, A. D., Moritz, G. L., Fosbury, R. A., Kawamura, S., & Dominy, N. J. (2012). Why aye‐ayes see blue. American Journal of Primatology, 74(3), 185192.Google Scholar
Mendoza, S. P., & Mason, W. A. (1997). Attachment relationships in New World primates. Annals of the New York Academy of Sciences, 807(1), 203209.Google Scholar
Menegaz, R. A., & Kirk, E. C. (2009). Septa and processes: Convergent evolution of the orbit in haplorhine primates and strigiform birds. Journal of Human Evolution, 57(6), 672687.Google Scholar
Miard, P., Nekaris, K. A. I., & Ramlee, H. (2017). Hiding in the dark: Local ecological knowledge about slow loris in Sarawak sheds light on relationships between human populations and wild animals. Human Ecology, 45(6), 823831.Google Scholar
Militz, T. A., & Foale, S. (2017). The ‘Nemo effect’: Perception and reality of Finding Nemo’s impact on marine aquarium fisheries. Fish and Fisheries, 18(3), 596606.Google Scholar
Miller, L. J., Pisacane, C. B., & Vicino, G. A. (2016). Relationship between behavioural diversity and faecal glucocorticoid metabolites: A case study with cheetahs (Acinonyx jubatus). Animal Welfare, 25(3), 325329.Google Scholar
Miller, R. A. (1943). Functional and morphological adaptations in the forelimbs of the slow lemurs. American Journal of Anatomy, 73(2), 153183.Google Scholar
Millspaugh, J., & Marzluff, J. M. (2001). Radio tracking and animal populations. San Diego, CA: Academic Press.Google Scholar
Milne Edwards, A. (1867). Note sur une nouvelle espèce du genre Nycticebus provenant de Siam et de Cochinchine. Nouvelles Archives du Musèum d’Histoire Naturelle de Paris, 3(2): 913.Google Scholar
Milne Edwards, A. (1874). Note sur le potto de Bosman ou Perodicticus potto. Bulletin Des Nouvelles Archives Du Muséum D’Histoire Naturelle De Paris, 10(3), 111114.Google Scholar
Milton, K. (1999). Nutritional characteristics of wild primate foods: Do the diets of our closest living relatives have lessons for us?. Nutrition, 15(6), 488498.Google Scholar
Mitchell, A., Gottfried, J., Barthel, M., & Shearer, E. (2018). The modern news consumer: News attitudes and practices in the digital era. Washington, DC: Pew Research Center.Google Scholar
Mittermeier, R. A., Wallis, J., Rylands, A. B., Ganzhorn, J. U., Oates, J. F., Williamson, E. A., …, & Supriatna, J. (2012). Primates in peril: The world’s 25 most endangered primates 2008–2010. Primate Conservation, 24(1), 158.Google Scholar
Mittermeier, R. A., Rylands, A. B., & Wilson, D. E. (2013). Handbook of the mammals of the world, vol. 3. Barcelona: Lynx Edicions.Google Scholar
Mivart, St G. (1864). Notes on the crania and dentition of the Lemuridae. Proceedings of the Zoological Society of London 1864, 32(1), 611648.Google Scholar
Mollah, A. R., Kundu, D. K., & Rahman, M. M. (2004). Site-level field appraisal for protected area co-management: Satchari Reserve Forest. Dhaka: Nature Conservation Management (NACOM).Google Scholar
Molur, S., Brandon‐Jones, D., Dittus, W., Eudey, A., Kumar, A., Singh, M., …, & Feeroz, S. E. (2003). Status of South Asian primates: Conservation assessment and management plan (C.A.M.P.) workshop report 2003. Coimbatore: Zoo Outreach Organisation/CBSG‐South Asia.Google Scholar
Monroe, M. C., Ballard, H. L., Oxarart, A., Sturtevant, V. E., Jakes, P. J., & Evans, E. R. (2016). Agencies, educators, communities and wildfire: Partnerships to enhance environmental education for youth. Environmental Education Research, 22(8), 10981114.Google Scholar
Moore, R. S. (2012). Ethics, ecology and evolution of Indonesian slow lorises (Nycticebus spp.) rescued from the pet trade (unpublished doctoral dissertation). Oxford Brookes University, Oxford, UK.Google Scholar
Moore, R. S., Wihermanto, , & Nekaris, K. A. I. (2014). Compassionate conservation, rehabilitation and translocation of Indonesian slow lorises. Endangered Species Research, 26(2), 93102.Google Scholar
Moore, R. S., Cabana, F., & Nekaris, K. A. I. (2015). Factors influencing stereotypic behaviours of animals rescued from Asian animal markets: A slow loris case study. Applied Animal Behaviour Science, 166, 131136.Google Scholar
Morgan, B. (2014). Tricks of the trade: The slender loris and other wildlife in supernatural rituals in Bangalore (unpublished Master’s dissertation). Oxford Brookes University, Oxford, UK.Google Scholar
Morgan, K. N., & Tromborg, C. T. (2007). Sources of stress in captivity. Applied Animal Behaviour Science, 102(3–4), 262302.Google Scholar
Morgan, M. A. (2015). Exotic addition. Duke Law Journal Online, 65(1), 1.Google Scholar
Moritz, G. L. (2015). Primate origins through the lens of functional and degenerate cone opsins (unpublished doctoral dissertation). Dartmouth College, Hanover, NH, USA.Google Scholar
Moritz, G. L., Ong, P. S., Perry, G. H., & Dominy, N. J. (2017). Functional preservation and variation in the cone opsin genes of nocturnal tarsiers. Philosophical Transactions of the Royal Society B: Biological Sciences, 372(1717), 20160075.Google Scholar
Morse, P. E., Chester, S. G., Boyer, D. M., Smith, T., Smith, R., Gigase, P., & Bloch, J. I. (2019). New fossils, systematics, and biogeography of the oldest known crown primate Teilhardina from the earliest Eocene of Asia, Europe, and North America. Journal of Human Evolution, 128, 103131.Google Scholar
Muchlinski, M. N. (2008). The relationship between the infraorbital foramen, infraorbital nerve, and maxillary mechanoreception: Implications for interpreting the paleoecology of fossil mammals based on infraorbital foramen size. The Anatomical Record, 291(10), 12211226.CrossRefGoogle ScholarPubMed
Muchlinski, M. N. (2010a). A comparative analysis of vibrissa count and infraorbital foramen area in primates and other mammals. Journal of Human Evolution, 58(6), 447473.CrossRefGoogle ScholarPubMed
Muchlinski, M. N. (2010b). Ecological correlates of infraorbital foramen area in primates. American Journal of Physical Anthropology, 141(1), 131141.Google Scholar
Muchlinski, M. N., & Kirk, E. C. (2017). A comparative analysis of infraorbital foramen size in Paleogene euarchontans. Journal of Human Evolution, 105, 5768.Google Scholar
Muchlinski, M. N., Durham, E. L., Smith, T. D., & Burrows, A. M. (2013). Comparative histomorphology of intrinsic vibrissa musculature among primates: Implications for the evolution of sensory ecology and ‘face touch’. American Journal of Physical Anthropology, 150(2), 301312.Google Scholar
Muchlinski, M. N., Wible, J. R., Corfe, I., Sullivan, M., & Grant, R. A. (2018). Good vibrations: The evolution of whisking in small mammals. The Anatomical Record. DOI: http://doi.org/10.1002/ar.23989.Google Scholar
Mukul, S. A., Sohel, M. S. I., Herbohn, J., Inostroza, L., & König, H. (2017). Integrating ecosystem services supply potential from future land-use scenarios in protected area management: A Bangladesh case study. Ecosystem Services, 26, 355364.Google Scholar
Müller, A. E. (1999). The social organisation of the fat-tailed dwarf lemur, Cheirogaleus medius (Lemuriformes; Primates) (unpublished doctoral dissertation). Universität Züric, Zurich, Switzerland.Google Scholar
Müller, A. E., & Thalmann, U. (2000). Origin and evolution of primate social organisation: A reconstruction. Biological Reviews of the Cambridge Philosophical Society, 75(3), 405435.Google Scholar
Müller, A. E., Soligo, C., & Thalmann, U. (2007a). New views on the origin of primate social organization. In Ravosa, M. J. & Dagosto, M. (Eds.), Primate origins: Adaptations and evolution (pp. 677701). Boston, MA: Springer.Google Scholar
Müller, B., Goodman, S. M., & Peichl, L. (2007b). Cone photoreceptor diversity in the retinas of fruit bats (Megachiroptera). Brain, Behavior and Evolution, 70(2), 90104.Google Scholar
Müller, E. F. (1979). Energy metabolism, thermoregulation and water budget in the slow loris (Nycticebus coucang, Boddaert 1785). Comparative Biochemistry and Physiology Part A: Physiology, 64(1), 109119.Google Scholar
Müller, E. F., Nieschalk, U., & Meier, B. (1985). Thermoregulation in the slender loris (Loris tardigradus). Folia Primatologica, 44(3–4), 216226.Google Scholar
Müller, P. L. S. (1773). Des Ritters Carl von Linné vollständiges Natursystem : Nach der zwölften lateinischen Ausgabe, und nach Anleitung des holländischen Houttuynischen Werks/mit einer ausführlichen Erklärung, ausgefertiget von Philipp Ludwig Statius Müller. Nuremberg: Gabriel Nicolaus Raspe.Google Scholar
Müller, P. L. S. (1776). Des Ritters Carl von Linné Königlich Schwedischen Leibarztes &c. &c. vollständigen Natursystems Supplements- und Register-Band über alle sechs Theile oder Classen des Thierreichs. Mit einer ausführlichen Erklärung. Nebst drey Kupfertafeln. Nuremberg: Gabriel Nicolaus Raspe.Google Scholar
Munds, R. A., Ali, R., Nijman, V., Nekaris, K. A. I., & Goossens, B. (2013a). Living together in the night: Abundance and habitat use of sympatric and allopatric populations of slow lorises and tarsiers. Endangered Species Research, 22(3), 269277.Google Scholar
Munds, R. A., Nekaris, K. A. I., & Ford, S. M. (2013b). Taxonomy of the Bornean slow loris, with new species Nycticebus kayan (Primates, Lorisidae). American Journal of Primatology, 75(1), 4656.Google Scholar
Munds, R. A., Titus, C. L., Eggert, L. S., & Blomquist, G. E. (2018). Using a multi-gene approach to infer the complicated phylogeny and evolutionary history of lorises (Order Primates: Family Lorisidae). Molecular Phylogenetics and Evolution, 127, 556567.Google Scholar
Murie, J., & Mivart, S. G. (1869). On the anatomy of the Lemuroidea. Transactions of the Zoological Society of London, 7(1), 1113.Google Scholar
Musing, L., Suzuki, K., & Nekaris, K. A. I. (2015). Crossing international borders: The trade of slow lorises, Nycticebus spp., as pets in Japan. Asian Primates, 5, 1224.Google Scholar
Muzaffar, S. B., Islam, M. A., Feeroz, M. M., Kabir, M., Begum, S., Mahmud, M. S., …, & Hasan, M. K. (2007). Habitat characteristics of the endangered hoolock Gibbons of Bangladesh: The role of plant species richness. Biotropica, 39(4), 539545.Google Scholar
Naarendorp, F., Esdaille, T. M., Banden, S. M., Andrews-Labenski, J., Gross, O. P., & Pugh, E. N. (2010). Dark light, rod saturation, and the absolute and incremental sensitivity of mouse cone vision. Journal of Neuroscience, 30(37), 1249512507.Google Scholar
Nakano, Y. (2002). The effects of substratum inclination on locomotor patterns in primates. Zeitschrift Für Morphologie Und Anthropologie, 83(2–3), 189199.Google Scholar
Nakano, Y., Ishida, H., & Hirasaki, E. (1996). The change of the locomotor pattern caused by the inclination of the substrata in a Japanese macaque. Primate Research, 12(2), 7987.CrossRefGoogle Scholar
Nakano, Y., Hirasaki, E., & Kumakura, H. (2006). Patterns of vertical climbing in primates. In Ishida, H., Tuttle, R. H., Pickford, M., Ogihara, N., & Nakatsukasa, M. (Eds.), Human Origins and Environmental Backgrounds (pp. 97104). Boston, MA: Springer.Google Scholar
Nakashima, Y., Nakabayashi, M., & Sukor, J. A. (2013). Space use, habitat selection, and day-beds of the common palm civet (Paradoxurus hermaphroditus) in human-modified habitats in Sabah, Borneo. Journal of Mammalogy, 94(5), 11691178.Google Scholar
Napier, J. R., & Napier, P. H. (1967). A handbook of living primates. New York, NY: Academic Press.Google Scholar
Nash, L. T. (1986a). Dietary, behavioral, and morphological aspects of gummivory in primates. American Journal of Physical Anthropology, 29(S7), 113137.CrossRefGoogle Scholar
Nash, L. T. (1986b). Influence of moonlight level on traveling and calling patterns in two sympatric species of Galago in Kenya. In Taub, D. M. & King, F. A. (Eds.), Current perspectives in primate social dynamics (pp. 357367). New York, NY: Van Nostrand Reinhold.Google Scholar
Nash, L. T. (1989). Galagos and gummivory. Human Evolution, 4(2–3), 199206.CrossRefGoogle Scholar
Nash, L. T. (1993). Juveniles in nongregarious primates. In Pereira, M. E. & Fairbanks, L. A. (Eds.), Juvenile primates: Life history, development, and behavior (pp. 119137). Oxford: Oxford University Press.Google Scholar
Nash, L. T. (2007). Moonlight and behavior in nocturnal and cathemeral primates, especially Lepilemur leucopus: Illuminating possible anti-predator efforts. In Gursky, S. L. & Nekaris, K. A.I. (Eds.), Primate anti-predator strategies: Developments in Primatology – Progress and Prospects (pp. 173205). Boston, MA: Springer.CrossRefGoogle Scholar
Nash, L. T., & Burrows, A. M. (2010). Introduction: Advances and remaining sticky issues in the understanding of exudativory in primates. In Burrows, A. M. & Nash, L. T. (Eds.), The evolution of exudativory in primates (pp. 124). New York, NY: Springer.Google Scholar
Nash, L. T., Bearder, S. K., & Olson, T. R. (1990). Synopsis of Galago species characteristics. International Journal of Primatology, 11(6), 629629.Google Scholar
Nash, S. V. (1993). Sold for a song: The trade in Southeast Asian non-CITES birds. Cambridge: TRAFFIC International.Google Scholar
Nathans, J., Thomas, D., & Hogness, D. S. (1986). Molecular genetics of human color vision: The genes encoding blue, green, and red pigments. Science, 232(4747), 193202.Google Scholar
National Research Council (NRC). (2003). Nutrient requirements of non-human primates. New York, NY: National Academies Press.Google Scholar
Nayak, U. V. (1933). A contribution to our study of the Lemuroidea: A comparative study of the Lorisinae and Galaginae (unpublished doctoral dissertation). University College London, London, UK.Google Scholar
Neagu, A. N., & Petraru, O. M. (2015). ‘Aquatic’ vs. ‘terrestrial’ eye design: A functional ecomorphological aproach. Analele S, tiinţifice ale Universitaţii,, Alexandru Ioan Cuza’ din Ias, i, s. Biologie animal, 61, 101115.Google Scholar
Nei, M., Niimura, Y., & Nozawa, M. (2008). The evolution of animal chemosensory receptor gene repertoires: Roles of chance and necessity. Nature Reviews Genetics, 9(12), 951.Google Scholar
Neilson, E., Nijman, V., & Nekaris, K. A. I. (2013). Conservation assessments of arboreal mammals in difficult terrain: Occupancy modeling of pileated gibbons (Hylobates pileatus). International Journal of Primatology, 34(4), 823835.Google Scholar
Nekaris, K. A. I. (2001). Activity budget and positional behavior of the Mysore slender loris (Loris tardigradus lydekkerianus): Implications for slow climbing locomotion. Folia Primatologica, 72(4), 228241.Google Scholar
Nekaris, K. A. I. (2003a). Observations of mating, birthing and parental behaviour in three subspecies of slender loris (Loris tardigradus and Loris lydekkerianus) in India and Sri Lanka. Folia Primatologica, 74(5–6), 312336.Google Scholar
Nekaris, K. A. I. (2003b). Spacing system of the Mysore slender loris (Loris lydekkerianus lydekkerianus). American Journal of Physical Anthropology, 121(1), 8696.Google Scholar
Nekaris, K. A. I. (2005a). Visual predation in the slender loris. Journal of Human Evolution, 49, 289300.CrossRefGoogle ScholarPubMed
Nekaris, K. A. I. (2005b). Foraging behaviour of the slender loris (Loris lydekkerianus lydekkerianus): Implications for theories of primate origins. Journal of Human Evolution, 49(3), 289300.Google Scholar
Nekaris, K. A. I. (2013). Family Lorisidae (Angwantibos, pottos, and lorises). In Mittermeier, R. A., Rylands, A. B., & Wilson, D. E. (Eds.), Handbook of the mammals of the world, vol. 3 (pp. 210235). Barcelona: Lynx Edicions.Google Scholar
Nekaris, K. A. I. (2014). Extreme primates: Ecology and evolution of Asian lorises. Evolutionary Anthropology: Issues, News, and Reviews, 23(5), 177187.Google Scholar
Nekaris, K. A. I. (2016). The Little Fireface Project: Community conservation of Asia’s slow lorises via ecology, education and empowerment. In Waller, M. (Ed.), Ethnoprimatology in the 21st century (pp. 259272). Zurich: Springer.Google Scholar
Nekaris, A., & Bearder, S. K. (2007). The Lorisiform primates of Asia and mainland Africa. In Campbell, C., Fuents, A., Mackinnon, K., Panger, M., & Bearder, S. (Eds.), Primates in perspective (pp. 2445). New York, NY: Oxford University Press.Google Scholar
Nekaris, K. A. I., & Bearder, S. K. (2011). The lorisiform primates of Asia and mainland Africa: Diversity shrouded in darkness. In Campbell, C., Fuentes, A., MacKinnon, K., Bearder, S., & Stumpf, R. (Eds.), Primates in perspective (2nd ed.) (pp. 3454). Oxford: Oxford University Press.Google Scholar
Nekaris, K. A. I., & de Silva Wijeyeratne, G. (2009). The primates of Sri Lanka. Colombo: Sri Lanka Tourism Promotion Bureau.Google Scholar
Nekaris, K., & Geerah, D. R. (2018). Novel use of pure ultrasonic communication by a wild nocturnal primate, the Javan slow loris (Nycticebus javanicus). American Journal of Physical Anthropology, 165, 186187.Google Scholar
Nekaris, K. A. I., & Jaffe, S. (2007). Unexpected diversity of slow lorises (Nycticebus spp.) within the Javan pet trade: Implications for slow loris taxonomy. Contributions to Zoology, 76(3), 187196.Google Scholar
Nekaris, K. A. I., & Jayewardene, J. (2003). Pilot study and conservation status of the slender loris (Loris tardigradus and L. lydekkerianus) in Sri Lanka. Primate Conservation, 19, 8390.Google Scholar
Nekaris, K. A. I., & Jayewardene, J. (2004a). Survey of the slender loris (Primates, Lorisidae Gray, 1821: Loris tardigradus Linnaeus, 1758 and Loris lydekkerianus cabrera, 1908) in Sri Lanka. Journal of Zoology, 262(4), 327338.Google Scholar
Nekaris, K. A. I., & Jayewardene, J. (2004b). Distribution of slender lorises in four ecological zones in Sri Lanka. Journal of Zoology, 262, 112.Google Scholar
Nekaris, K. A. I., & Munds, R. (2010). Using facial markings to unmask diversity: The slow lorises (Primates: Lorisidae: Nycticebus spp.) of Indonesia. In: Gursky, S. & Supriatna, J. (Eds.), The Primates of Indonesia (pp. 383396). New York, NY: Springer.Google Scholar
Nekaris, K. A. I., & Nijman, V. (2007). CITES proposal highlights rarity of Asian nocturnal primates (Lorisidae: Nycticebus). Folia Primatologica, 78(4), 211214.Google Scholar
Nekaris, K. A. I., & Nijman, V. (2018). Successful prosecution of slow loris traders in Indonesia. Oryx, 52(3), 411.Google Scholar
Nekaris, K. A. I., & Rasmussen, D. T. (2003). Diet and feeding behavior of Mysore slender lorises. International Journal of Primatology, 24(1), 3346.Google Scholar
Nekaris, K. A. I., & Starr, C. R. (2015). Conservation and ecology of the neglected slow loris: Priorities and prospects. Endangered Species Research, 28(1), 8795.Google Scholar
Nekaris, K. A. I., & Stengel, C. J. (2013). Where are they? Distribution and microhabitat use of fragments by red slender loris (Loris tardigradus tardigradus) in Sri Lanka. In Marsh, L. K. & Chapman, C. A. (Eds.), Primates in fragments: Complexity and resilience (pp. 371384). New York, NY: Springer.Google Scholar
Nekaris, K. A. I., & Stevens, N. J. (2007). Not all lorises are slow: Rapid arboreal locomotion in Loris tardigradus of southwestern Sri Lanka. American Journal of Primatology, 69(1), 113121.Google Scholar
Nekaris, A., & Streicher, U. (2008a). Nycticebus coucang. The IUCN Red List of Threatened Species 2008. Retrieved 1 November 2018, from http://dx.doi.org/10.2305/IUCN.UK.2008.RLTS.T39759A10263403.en.Google Scholar
Nekaris, A., & Streicher, U. (2008b). Nycticebus menagensis. The IUCN Red List of Threatened Species 2008. Retrieved 1 November 2018, from http://dx.doi.org/10.2305/IUCN.UK.2008.RLTS.T39760A10263652.en.Google Scholar
Nekaris, K. A. I., Liyanage, W. K. D. D., & Gamage, S. (2005). Relationship between forest structure and floristic composition and population density of the southwestern Ceylon slender loris (Loris tardigradus tardigradus) in Masmullah Forest, Sri Lanka. Mammalia, 69(2), 110.Google Scholar
Nekaris, K. A. I., Pimley, E. R., & Ablard, K. M. (2007). Predator defense by slender lorises and pottos. In Gursky-Doyen, S. & Nekaris, K. A. I. (Eds.), Primate anti-predator strategies (pp. 222240). New York, NY: Springer.Google Scholar
Nekaris, K. A. I., Blackham, G. V., & Nijman, V. (2008a). Conservation implications of low encounter rates of five nocturnal primate species (Nycticebus spp.) in Asia. Biodiversity and Conservation, 17(4), 733747.CrossRefGoogle Scholar
Nekaris, K. A. I., Singh, M., & Chhangani, A. K. (2008b). Loris lydekkerianus ssp. nordicus. IUCN Red List of Threatened Species. Version 012.2. Retrieved 1 June 2014, from www.iucnredlist.org.Google Scholar
Nekaris, K. A. I., Starr, C.R., Collins, R.L., & Wilson, A. (2010a). Comparative ecology of exudate feeding by lorises (Nycticebus, Loris) and pottos (Perodicticus, Arctocebus). In Burrows, A. & Nash, L. (Eds.), The evolution of exudativory in primates: Developments in primatology – progress and prospects (pp. 155168). New York, NY: Springer.Google Scholar
Nekaris, K. A. I., Shepherd, C. R., Starr, C. R., & Nijman, V. (2010b). Exploring cultural drivers for wildlife trade via an ethnoprimatological approach: A case study of slender and slow lorises (Loris and Nycticebus) in South and Southeast Asia. American Journal of Primatology, 72(10), 877886.Google Scholar
Nekaris, K. A. I., Campbell, N., Coggins, T. G., Rode, E. J., & Nijman, V. (2013a). Tickled to death: Analysing public perceptions of ‘cute’videos of threatened species (slow lorises –Nycticebus spp.) on Web 2.0 sites. PLoS One, 8(7), e69215.Google Scholar
Nekaris, K. A. I., Moore, R. S., Rode, E. J., & Fry, B. G. (2013b). Mad, bad and dangerous to know: The biochemistry, ecology and evolution of slow loris venom. Journal of Venomous Animals and Toxins Including Tropical Diseases, 19(1), 21.Google Scholar
Nekaris, K. A. I., Jaffe, S. M., & Donati, G. (2013c). Forest fragmentation imperils red slender lorises (Loris tardigradus tardigradus) in south-western Sri Lanka. In Masters, J., Gamba, M., & Génin, F. (Eds.), Leaping ahead: Advances in prosimian biology (pp. 8996). New York, NY: Springer.Google Scholar
Nekaris, K. A. I., Pambudi, J. A. A., Susanto, D., Ahmad, R. D., & Nijman, V. (2014). Densities, distribution and detectability of a small nocturnal primate (Javan slow loris Nycticebus javanicus) in a montane rainforest. Endangered Species Research, 24(2), 95103.CrossRefGoogle Scholar
Nekaris, K. A. I., Musing, L., Vazquez, A. G., & Donati, G. (2016). Is tickling torture? Assessing welfare towards slow lorises (Nycticebus spp.) within Web 2.0 videos. Folia Primatologica, 86(6), 534551.Google Scholar
Nekaris, K. A. I., Poindexter, S., Reinhardt, K. D., Sigaud, M., Cabana, F., Wirdateti, W., & Nijman, V. (2017). Coexistence between Javan slow lorises (Nycticebus javanicus) and humans in a dynamic agroforestry landscape in West Java, Indonesia. International Journal of Primatology, 38(2), 303320.Google Scholar
Nekaris, K. A. I., Weldon, A., Imron, M. A., Maynard, K. Q., Nijman, V., Poindexter, S. A., & Morcatty, T. Q. (2019). Venom in furs: Facial masks as aposematic signals in a venomous mammal. Toxins, 11, 93.Google Scholar
Neri-Arboleda, I., Stott, P., & Arboleda, N. P. (2002). Home ranges, spatial movements and habitat associations of the Philippine tarsier (Tarsius syrichta) in Corella, Bohol. Journal of Zoology, 257(3), 387402.CrossRefGoogle Scholar
Nett, E. M., & Ravosa, M. J. (2019). Ontogeny of orbit orientation in primates. The Anatomical Record, 302, 20932104.Google Scholar
Nevo, O., & Heymann, E. W. (2015). Led by the nose: Olfaction in primate feeding ecology. Evolutionary Anthropology: Issues, News, and Reviews, 24(4), 137148.Google Scholar
Newing, H. (2010). Conducting research in conservation: Social science methods and practice. Abingdon: Routledge.Google Scholar
Ng, J., & Nemora, (2007). Tiger Trade Revisited in Sumatra, Indonesia. Petaling Jaya: TRAFFIC.Google Scholar
Ni, X., Gebo, D. L., Dagosto, M., Meng, J., Tafforeau, P., Flynn, J. J., & Beard, K. C. (2013). The oldest known primate skeleton and early haplorhine evolution. Nature, 498(7452), 60.Google Scholar
Nichols, J. D., & Karanth, K. U. (2002). Statistical concepts: Assessing spatial distributions. In Karanth, K. U. & Nichols, J. D. (Eds.), Monitoring tigers and their prey: A manual for wildlife managers, researchers and conservationists (pp. 2938). Bangalore: Centre for Wildlife Studies.Google Scholar
Nichols, J. D., & Williams, B. K. (2006). Monitoring for conservation. Trends in Ecology & Evolution, 21(12), 668673.Google Scholar
Nieschalk, U. (1991). Fortbewegung und Funktionsmorphologie von Loris tardigradus und anderen kleinen Halbaffen in Anpassung an unterschiedliche Habitate (unpublished doctoral dissertation). Ruhr-Universität Bochum, Bochum, Germany.Google Scholar
Nieschalk, U., & Demes, B. (1993). Biomechanical determinants of reduction of the second ray in Lorisinae. In Preuschoft, H. & Chivers, D. (Eds.), Hands of primates (pp. 225234). Vienna: Springer.Google Scholar
Niimura, Y. (2012). Olfactory receptor multigene family in vertebrates: From the viewpoint of evolutionary genomics. Current Genomics, 13(2), 103114.Google Scholar
Niimura, Y., Matsui, A., & Touhara, K. (2018). Acceleration of olfactory receptor gene loss in primate evolution: Possible link to anatomical change in sensory systems and dietary transition. Molecular Biology and Evolution, 35(6), 14371450.Google Scholar
Nijman, V. (2005). In full swing: Assessment of trade in orang-utans and gibbons on Java and Bali, Indonesia. Petaling Jaya: TRAFFIC Southeast Asia.Google Scholar
Nijman, V., & Healy, A. (2016). Present-day international primate trade in historical context. In Wich, S. & Marshall, A. (Eds.), An introduction to primate conservation (pp. 129142). Oxford: Oxford University Press.Google Scholar
Nijman, V., & Nekaris, K. A. I. (2010). Checkerboard patterns, interspecific competition, and extinction: Lessons from distribution patterns of tarsiers (Tarsius) and slow lorises (Nycticebus) in insular Southeast Asia. International Journal of Primatology, 31(6), 11471160.Google Scholar
Nijman, V., & Nekaris, K. A. I. (2014a). Trade in wildlife for medicinal and decorative purposes in Bali, Indonesia.TRAFFIC Bulletin, 14, 3136.Google Scholar
Nijman, V., & Nekaris, K. A. I. (2014b). Traditions, taboos and trade in slow lorises in Sundanese communities in southern Java, Indonesia. Endangered Species Research, 25(1), 7988.CrossRefGoogle Scholar
Nijman, V., & Shepherd, C. R. (2007). Trade in non-native, CITES-listed, wildlife in Asia, as exemplified by the trade in freshwater turtles and tortoises (Chelonidae) in Thailand. Contributions to Zoology, 76(3), 207211.Google Scholar
Nijman, V., Nekaris, K. A. I., Donati, G., Bruford, M., & Fa, J. (2011). Primate conservation: Measuring and mitigating trade in primates. Endangered Species Research, 13(2), 159161.Google Scholar
Nijman, V., Shepherd, C. R., & Nekaris, K. A. I. (2014). Trade in Bengal slow lorises in Mong La, Myanmar, on the China border. Primate Conservation, 2014(28), 139143.Google Scholar
Nijman, V., Bergin, D., & van Lavieren, E. (2015). Barbary macaques exploited as photo-props in Marrakesh’s punishment square. Swara. July–September, 38–41.Google Scholar
Nijman, V., Spaan, D., Rode‐Margono, E. J., & Nekaris, K. A. I. (2017). Changes in the primate trade in Indonesian wildlife markets over a 25‐year period: Fewer apes and langurs, more macaques, and slow lorises. American Journal of Primatology, 79(11), e22517.Google Scholar
Noble, V. E., Kowalski, E. M., & Ravosa, M. J. (2000). Orbit orientation and the function of the mammalian postorbital bar. Journal of Zoology, 250(3), 405418.Google Scholar
Norconk, M. A., Wright, B. W., Conklin-Brittain, N. L., & Vinyard, C. J. (2009). Mechanical and nutritional properties of food as factors in platyrrhine dietary adaptations. In Garber, P. A., Estrada, A., Bicca-Marques, J. C., Heymann, E. W., & Strier, K. B. (Eds.), South American primates: Developments in primatology – progress and prospects (pp. 279319). New York, NY: Springer.CrossRefGoogle Scholar
Nowack, J., & Dausmann, K. H. (2015). Can heterothermy facilitate the colonization of new habitats?. Mammal Review, 45(2), 117127.Google Scholar
Nowack, J., Mzilikazi, N., & Dausmann, K. H. (2010). Torpor on demand: Heterothermy in the non-lemur primate Galago moholi. PLoS One, 5(5), e10797.Google Scholar
Nowack, J., Mzilikazi, N., & Dausmann, K. H. (2013). Torpor as an emergency solution in Galago moholi: Heterothermy is triggered by different constraints. Journal of Comparative Physiology B, 183(4), 547556.CrossRefGoogle ScholarPubMed
Nummela, S., Pihlström, H., Puolamäki, K., Fortelius, M., Hemilä, S., & Reuter, T. (2013). Exploring the mammalian sensory space: Co-operations and trade-offs among senses. Journal of Comparative Physiology A, 199(12), 10771092.Google Scholar
Nunn, C. L. (2000). Social evolution in primates: The relative roles of ecology and intersexual conflict. In van Schaik, C. P. (Ed.), Infanticide by males and its implications (pp. 388419). Cambridge: Cambridge University Press.Google Scholar
Nussinovitch, A. (2009). Plant gum exudates of the world: Sources, distribution, properties, and applications. Chicago, IL: CRC Press.Google Scholar
Nyakatura, J. A., & Heymann, E. W. (2010). Effects of support size and orientation on symmetric gaits in free-ranging tamarins of Amazonian Peru: Implications for the functional significance of primate gait sequence patterns. Journal of Human Evolution, 58(3), 242251.CrossRefGoogle ScholarPubMed
Nyakatura, J. A., Fischer, M. S., & Schmidt, M. (2008). Gait parameter adjustments of cotton‐top tamarins (Saguinus oedipus, Callitrichidae) to locomotion on inclined arboreal substrates. American Journal of Physical Anthropology, 135(1), 1326.Google Scholar
O’Brien, T., Strindberg, S., & Wilkie, D. (2012). Measuring conservation effectiveness: Occupancy-related metrics for willdlife. New York: Wildlife Conservation Society.Google Scholar
Oates, J. F. (1984). The niche of the potto, Perodicticus potto. International Journal of Primatology, 5(1), 51.Google Scholar
Oates, J. F. (2011). Primates of West Africa: A field guide and natural history. Arlington, VA: Conservation International.Google Scholar
Oates, J. F., & Ambrose, L. (2013). Arctocebus calabarensis: Calabar angwantibo. In Butynski, T. M., Kingdon, J., & Kalina, J. (Eds.), Mammals of Africa, vol. 2 (pp. 400402). London: Bloomsbury Publishing.Google Scholar
Oates, J. F., & Bearder, S. (2008). Arctocebus calabarensis. The IUCN Red List of Threatened Species. Retrieved 13 September 2019, from www.iucnredlist.org/species/2054/9211259.Google Scholar
O’Dea, J. D. (1990). The mammalian rete mirabile and oxygen availability. Comparative Biochemistry and Physiology Part A: Physiology, 95(1), 2325.Google Scholar
OfCom. (2017). Children’s and parents’ media use and attitudes. Retrieved 13 September 2019, from www.ofcom.org.uk/research-and-data/media-literacy-research/childrens/children-parents-2017.Google Scholar
OfCom. (2018). Adults’ media use and attitudes. Retrieved 13 September 2019, from www.ofcom.org.uk/research-and-data/media-literacy-research/adults/adults-media-use-and-attitudes.Google Scholar
Off, E. C. (2003). Galago (Galago matschiei and Galagoides thomasi) locomotion and habitat use in Kibale National Forest, Uganda (unpublished Master’s disseration). Northern Illinois University, DeKalb, IL, USA.Google Scholar
Oftedal, O. T., & Allen, M. E. (1997). The feeding and nutrition of omnivores with emphasis on primates. In Kleiman, D. G., Allen, M. E., & Thompson, K. V. (Eds.), Wild mammals in captivity: Principles and techniques (pp. 148157). Chicago, IL: University of Chicago Press.Google Scholar
Olejniczak, A. J., & Grine, F. E. (2006). Assessment of the accuracy of dental enamel thickness measurements using microfocal X‐ray computed tomography. The Anatomical Record Part A, 288(3), 263275.Google Scholar
Olender, T., Lancet, D., & Nebert, D. W. (2008). Update on the olfactory receptor (OR) gene superfamily. Human Genomics, 3(1), 87.Google Scholar
Olson, T. R. & Nash, L. T. (2002–2003). Galago (Galagonidae) body measurements and museum collections data. African Primates, 6, 5053.Google Scholar
Olupot, W., McNeilage, A. J., & Plumptre, A. J. (2009). An analysis of socioeconomics of bushmeat hunting at major hunting sites in Uganda. Wildlife Conservation Society (WCS) Working Paper 38.Google Scholar
O’Mara, M. T., Gordon, A. D., Catlett, K. K., Terranova, C. J., & Schwartz, G. T. (2012). Growth and the development of sexual size dimorphism in lorises and galagos. American Journal of Physical Anthropology, 147(1), 1120.Google Scholar
Onderdonk, D. A., & Chapman, C. A. (2000). Coping with forest fragmentation: The primates of Kibale National Park, Uganda. International Journal of Primatology, 21(4), 587611.Google Scholar
Oonincx, D. G. A. B., & Van der Poel, A. F. B. (2011). Effects of diet on the chemical composition of migratory locusts (Locusta migratoria). Zoo Biology, 30(1), 916.CrossRefGoogle ScholarPubMed
Opie, C. F. (2013). The evolution of social systems in human and non-human primates (unpublished doctoral dissertation). Oxford University, Oxford, UK.Google Scholar
Opie, C. F., Atkinson, Q. D., & Shultz, S. (2012). The evolutionary history of primate mating systems. Communicative & Integrative Biology, 5(5), 458461.Google Scholar
Opie, C. F., Atkinson, Q. D., Dunbar, R. I., & Shultz, S. (2013). Male infanticide leads to social monogamy in primates. Proceedings of the National Academy of Sciences, 110(33), 1332813332.CrossRefGoogle ScholarPubMed
Opie, C., Atkinson, Q. D., Dunbar, R. I. M., & Shultz, S. (2014). Reply to Lukas and Clutton-Brock: Infanticide still drives primate monogamyProceedings of the National Academy of Sciences, 111(17), E1675CrossRefGoogle ScholarPubMed
Organ, J. M., Muchlinski, M. N., & Deane, A. S. (2011). Mechanoreceptivity of prehensile tail skin varies between ateline and cebine primates. The Anatomical Record, 294(12), 20642072.Google Scholar
Organisciak, D. T., & Vaughan, D. K. (2010). Retinal light damage: Mechanisms and protection. Progress in Retinal and Eye Research, 29(2), 113134.Google Scholar
Orme, D., Freckleton, R., Thomas, G., & Petzoldt, T. (2013). The caper package: Comparative analysis of phylogenetics and evolution in R. R Package Version 5(2), 1–36.Google Scholar
Orr, I. C. (1974). Puppet theatre in Asia. Asian Folklore Studies, 33(1), 6984.Google Scholar
Osterberg, P., & Nekaris, K. A. I. (2015). The use of animals as photo props to attract tourists in Thailand: A case study of the slow loris Nycticebus spp. TRAFFIC Bulletin, 27(1), 1318.Google Scholar
Otis, D. L., Burnham, K. P., White, G. C., & Anderson, D. R. (1978). Statistical inference from capture data on closed animal populations. Wildlife Monographs, 62, 3135.Google Scholar
Overdorff, D. J. (1996). Ecological correlates to activity and habitat use of two prosimian primates: Eulemur rubriventer and Eulemur fulvus rufus in Madagascar. American Journal of Primatology, 40(4), 327342.3.0.CO;2-#>CrossRefGoogle ScholarPubMed
Owen, K., Murphy, D., & Parsons, C. (2009). ZATPAC: A model consortium evaluates teen programs. Zoo Biology, 28(5), 429446.CrossRefGoogle Scholar
Oxnard, C. E., Crompton, R. H., & Lieberman, S. S. (1990). Animal lifestyles and anatomies: The case of the prosimian primates. Washington, DC: University of Washington Press.Google Scholar
Padua, S. M. (2010). Primate conservation: Integrating communities through environmental education programs. American Journal of Primatology, 72(5), 450453.CrossRefGoogle ScholarPubMed
Palagi, E., & Norscia, I. (2009). Multimodal signaling in wild Lemur catta: Economic design and territorial function of urine marking. American Journal of Physical Anthropology, 139(2), 182192.CrossRefGoogle ScholarPubMed
Palsbøll, P. J., Berube, M., & Allendorf, F. W. (2007). Identification of management units using population genetic data. Trends in Ecology & Evolution, 22(1), 1116.CrossRefGoogle ScholarPubMed
Pan, D., Chen, J. H., Groves, C., Wang, Y. X., Narushima, E., Fitch-Snyder, H., …, & Chemnick, L. (2007). Mitochondrial control region and population genetic patterns of Nycticebus bengalensis and N. pygmaeus. International Journal of Primatology, 28(4), 791799.Google Scholar
Pantel, S., & Awang Anak, N. (2010). A preliminary assessment of pangolin trade in Sabah. Petaling Jaya: TRAFFIC Southeast Asia.Google Scholar
Papailiou, A., Sullivan, E., & Cameron, J. L. (2008). Behaviors in rhesus monkeys (Macaca mulatta) associated with activity counts measured by accelerometer. American Journal of Primatology, 70(2), 185190.Google Scholar
Papworth, S. K., Nghiem, T. P. L., Chimalakonda, D., Posa, M. R. C., Wijedasa, L. S., Bickford, D., & Carrasco, L. R. (2015). Quantifying the role of online news in linking conservation research to Facebook and Twitter. Conservation Biology, 29(3), 825833.Google Scholar
Paradis, E., Claude, J., & Strimmer, K. (2004). APE: Analyses of phylogenetics and evolution in R language. Bioinformatics, 20(2), 289290.Google Scholar
Pariente, G. (1976). Les différents aspects de la limite du tapetum lucidum chez les prosimiens. Vision Research, 16(4), 387391.Google Scholar
Pariente, G. (1979). The role of vision in prosimian behavior. In Doyle, G. A. & Martin, R. D. (Eds.), The study of prosimian behavior (pp. 411459). New York, NY: Academic Press.Google Scholar
Pariente, G. (1980). Quantitative and qualitative study of the light available in the natural biotope of Malagasy prosimians. In Charles- Dominque, P., Cooper, H. M., Hladik, A., Hladik, C. M., Pages, E., Pariente, G., …, & Schilling, A. (Eds.), Nocturnal Malagasy primates: Ecology, physiology, and behavior (pp. 117134). New York, NY: Academic Press.Google Scholar
Passamani, M. (1998). Activity budget of Geoffroy’s marmoset (Callithrix geoffroyi) in an Atlantic forest in southeastern Brazil. American Journal of Primatology, 46(4), 333340.3.0.CO;2-7>CrossRefGoogle Scholar
Patel, B. A., Boyer, D. M., Perchalski, B. A., Ryan, T. M., St. Clair, E. M., Winchester, J. M., & Seiffert, E. R. (2017). New fossils and the paleobiology of Karanisia clarki from the late Eocene of Egypt. American Journal of Physical Anthropology, 162 (S64), 310311.Google Scholar
Paulli, S. (1900). Über die Pneumaticität des Schädels bei den Säugethieren: II. Über die Morphologie des Siebbeins der Pneumaticität bei den Ungulaten und Probosciden. Gegenbaurs Morphol Jahrb, 28, 179251.Google Scholar
Peck, M., Thorn, J., Mariscal, A., Baird, A., Tirira, D., & Kniveton, D. (2011). Focusing conservation efforts for the critically endangered brown-headed spider monkey (Ateles fusciceps) using remote sensing, modeling, and playback survey methods. International Journal of Primatology, 32(1), 134148.Google Scholar
Peichl, L. (2005). Diversity of mammalian photoreceptor properties: Adaptations to habitat and lifestyle?. The Anatomical Record Part A, 287(1), 10011012.CrossRefGoogle ScholarPubMed
Peichl, L., Kaiser, A., Rakotondraparany, F., Dubielzig, R. R., Goodman, S. M., & Kappeler, P. M. (2019). Diversity of photoreceptor arrangements in nocturnal, cathemeral and diurnal Malagasy lemurs. Journal of Comparative Neurology, 527(1), 1337.CrossRefGoogle ScholarPubMed
Peng, M. L., Tsai, C. Y., Chien, C. L., Hsiao, J. C. J., Huang, S. Y., Lee, C. J., …, & Tseng, K. W. (2012). The influence of low-powered family LED lighting on eyes in mice experimental model. Life Sciences Journal, 9(1), 477482.Google Scholar
Perelman, P., Johnson, W. E., Roos, C., Seuánez, H. N., Horvath, J. E., Moreira, M. A., …, & Schneider, M. P. C. (2011). A molecular phylogeny of living primates. PLoS Genetics, 7(3), e1001342.Google Scholar
Perera, M. S. J. (2008). A review of the distribution of grey slender loris (Loris lydekkerianus) in Sri Lanka. Primate Conservation, 23(1), 8997.Google Scholar
Perera, M. S. J., Rodrigo, R. K., Perera, W. P. N., Samarawickrema, V. A. P., Asela, M. D. C., Chandranimal, D., & Perera, P. G. D. R. (2009). Extension of range of the grey slender loris (Loris lydekkerianus) towards south east of Sri Lanka. Journal of the Wildlife and Nature Protection Society of Sri Lanka, 25, 57.Google Scholar
Peres, C. A. (2000). Identifying keystone plant resources in tropical forests: The case of gums from Parkia pods. Journal of Tropical Ecology, 16(2), 287317.Google Scholar
Perkin, A. (2004). Galagos of the coastal forests and eastern Arc Mtns. of Tanzania: Notes and records. Tanzania Forest Conservation Group, Technical Paper 8.Google Scholar
Perrigo, G., & Bronson, F. H. (1985). Sex differences in the energy allocation strategies of house mice. Behavioral Ecology and Sociobiology, 17(3), 297302.Google Scholar
Perry, G. H., Martin, R. D., & Verrelli, B. C. (2007). Signatures of functional constraint at aye-aye opsin genes: The potential of adaptive color vision in a nocturnal primate. Molecular Biology and Evolution, 24(9), 19631970.Google Scholar
Peters, W. H. C. (1876). Űber die von dem verstorbenen Professor Dr Rheinhold Buchholz in Westafrika gesammelten Säugethiere. Monatsberichte der Königlichen Preussischen Akademie der Wissenschaften zu Berlin (Session of 7 August 1876), 469–485.Google Scholar
Peterson, B. E., Goldreich, D., & Merzenich, M. M. (1998). Optical imaging and electrophysiology of rat barrel cortex: I. Responses to small single-vibrissa deflections. Cerebral Cortex, 8(2), 173183.Google Scholar
Petter, J. J., Schilling, A., & Pariente, G. (1975). Observations on the behaviour and ecology of Phaner furcifer. In Tatersall, I. & Sussman, R. W. (Eds.), Lemur biology (pp. 209218). New York, NY: Plenum Press.CrossRefGoogle Scholar
Pettigrew, J. D. (1978). Comparison of the retinotopic organization of the visual wulst in nocturnal and diurnal raptors, with a note on the evolution of frontal vision. In Cool, S. J. & Smith, E. L. (Eds.), Frontiers of visual science (pp. 328335). New York, NY: Springer-Verlag.Google Scholar
Pettigrew, J. D., Dreher, B., Hopkins, C. S., McCall, M. J., & Brown, M. (1988). Peak density and distribution of ganglion cells in the retinae of microchiropteran bats: Implications for visual acuity (Part 1 of 2). Brain, Behavior and Evolution, 32(1), 3947.Google Scholar
Pettigrew, J. D., Jamieson, B. G. M., Robson, S. K., Hall, L. S., McNally, K. I., & Cooper, H. M. (1989). Phylogenetic relations between microbats, megabats and primates (Mammalia: Chiroptera and Primates). Philosophical Transactions of the Royal Society B: Biological Sciences, 325(1229), 489559.Google ScholarPubMed
Phelps, J., Webb, E. L., Bickford, D., Nijman, V., & Sodhi, N. S. (2010). Boosting cites. Science, 330(6012), 17521753.Google Scholar
Phillips, E. M., & Walker, A. (2000). A new species of fossil lorisid from the Miocene of East Africa. Primates, 41(4), 367372.Google Scholar
Phillips, E. M., & Walker, A. (2002). Fossil lorisoids. In Hartwig, W. (Ed.), The primate fossil record. (pp. 8295). Cambridge: Cambridge University Press.Google Scholar
Pickford, M. (2012). Lorisine primate from the Late Miocene of Kenya. Journal of Biological Research: Bollettino Della Società Italiana Di Biologia Sperimentale, 85(1), 4752.Google Scholar
Pickford, M. (2015a). Late Eocene Potamogalidae and Tenrecidae (Mammalia) from the Sperrgebiet, Namibia. Communications of the Geological Survey of Namibia, 16, 121159.Google Scholar
Pickford, M. (2015b). Bothriogenys (Anthracotheriidae) from the Bartonian of Eoridge, Namibia. Communications of the Geological Survey of Namibia, 16, 215222.Google Scholar
Pickford, M. (2015c). Late Eocene lorisiform primate from Eocliff, Sperrgebiet, Namibia. Communications of the Geological Survey of Namibia, 16, 194199.Google Scholar
Pickford, M., & Kunimatsu, Y. (2005). Catarrhines from the Middle Miocene (ca. 14.5 Ma) of Kipsaraman, Tugen Hills, Kenya. Anthropological Science, 113(2), 189224.Google Scholar
Pickford, M., Senut, B., Mocke, H., Mourer-Chauviré, C., Rage, J. C., & Mein, P. (2014). Eocene aridity in southwestern Africa: Timing of onset and biological consequences. Transactions of the Royal Society of South Africa, 69(3), 139144.Google Scholar
Piep, M., Radespiel, U., Zimmermann, E., Schmidt, S., & Siemers, B. M. (2008). The sensory basis of prey detection in captive-born grey mouse lemurs, Microcebus murinus. Animal Behaviour, 75(3), 871878.CrossRefGoogle Scholar
Pihlström, H. (2012). The size of major mammalian sensory organs as measured from cranial characters, and their relation to the biology and evolution of mammals (unpublished doctoral dissertation). University of Helsinki, Helsinki, Finland.Google Scholar
Pihlström, H., Fortelius, M., Hemilä, S., Forsman, R., & Reuter, T. (2005). Scaling of mammalian ethmoid bones can predict olfactory organ size and performance. Proceedings of the Royal Society B: Biological Sciences, 272(1566), 957962.Google Scholar
Pimley, E. R. (2003). Behavioural ecology and genetics of two nocturnal prosimians: Pottos (Perodicticus potto edwardsi) and Allen’s bushbabies (Galago alleni cameronensis) (unpublished doctoral dissertation). University of Cambridge, Cambridge, UK.Google Scholar
Pimley, E. (2009). A survey of nocturnal primates (Strepsirrhini: Galaginae, Perodictinae) in southern Nigeria. African Journal of Ecology, 47(4), 784787.Google Scholar
Pimley, E. R., & Bearder, S. K. (2013). Perodicticus potto: potto. In Butynski, T. M., Kingdon, J., & Kalina, J. (Eds.), Mammals of Africa (vol. 2, pp. 393399). London: Bloomsbury Publishing.Google Scholar
Pimley, E. R., Bearder, S. K., & Dixson, A. F. (2005a). Home range analysis of Perodicticus potto edwardsi and Sciurocheirus cameronensis. International Journal of Primatology, 26(1), 191205.Google Scholar
Pimley, E. R., Bearder, S. K., & Dixson, A. F. (2005b). Social organization of the Milne‐Edward’s potto. American Journal of Primatology, 66(4), 317330.Google Scholar
Pirie, A. (1959). Crystals of riboflavin making up the tapetum lucidum in the eye of a lemur. Nature, 183(4666), 985.Google Scholar
Plesker, R., & Schulze, H. (2015). Necropsy findings in slender Lorises (Loris lydekkerianus). Vietnamese Journal of Primatology, 2, 4955.Google Scholar
Pliosungnoen, M., Gale, G., & Savini, T. (2010). Density and microhabitat use of Bengal slow loris in primary forest and non‐native plantation forest. American Journal of Primatology, 72(12), 11081117.Google Scholar
Pocock, R. I. (1914). On the facial vibrissae of mammalia. Proceedings of the Zoological Society of London, 84(3), 889912.Google Scholar
Poindexter, S., & Nekaris, K. A. I. (2017a). Lorisiformes. In Fuentes, A. (Ed.), The international encyclopedia of primatology (pp. 13). Oxford: Wiley-Blackwell.Google Scholar
Poindexter, S. A., & Nekaris, K. A. I. (2017b). Vertical clingers and gougers: Rapid acquisition of adult limb proportions facilitates feeding behaviours in young Javan slow lorises (Nycticebus javanicus). Mammalian Biology, 87, 4049.Google Scholar
Poindexter, S., Khoa, D. D., & Nekaris, K. A. I. (2017). Ranging patterns of reintroduced pygmy slow lorises (Nycticebus pygmaeus) in Cuc Phuong National Park, Vietnam. Vietnamese Journal of Primatology, 2(5), 3779.Google Scholar
Pollock, K. H., Nichols, J. D., Simons, T. R., Farnsworth, G. L., Bailey, L. L., & Sauer, J. R. (2002). Large scale wildlife monitoring studies: Statistical methods for design and analysis. Environmetrics, 13(2), 105119.Google Scholar
Porter, C. A., Page, S. L., Czelusniak, J., Schneider, H., Schneider, M. P. C., Sampaio, I., & Goodman, M. (1997). Phylogeny and evolution of selected primates as determined by sequences of the ε-globin locus and 5′ flanking regions. International Journal of Primatology, 18(2), 261295.Google Scholar
Porter, L., & Golan, G. (2006). From subservient chickens to brawny men. Journal of Interactive Advertising, 6(2), 2633.Google Scholar
Power, D. J., & Phillips-Wren, G. (2011). Impact of social media and Web 2.0 on decision-making. Journal of Decision Systems, 20(3), 249261.Google Scholar
Power, M. L. (2010). Nutritional and digestive challenges to being a gum-feeding primate. In Burrows, A. M. & Nash, L. T. (Eds.), The evolution of exudativory in primates (pp. 2544). New York, NY: Springer.Google Scholar
Pozzi, L. (2016). The role of forest expansion and contraction in species diversification among galagos (Primates: Galagidae). Journal of Biogeography, 43(10), 19301941.Google Scholar
Pozzi, L., Disotell, T. R., & Masters, J. C. (2014a). A multilocus phylogeny reveals deep lineages within African galagids (Primates: Galagidae). BMC Evolutionary Biology, 14(1), 72.Google Scholar
Pozzi, L., Hodgson, J. A., Burrell, A. S., Sterner, K. N., Raaum, R. L., & Disotell, T. R. (2014b). Primate phylogenetic relationships and divergence dates inferred from complete mitochondrial genomes. Molecular Phylogeny and Evolution, 75, 165183.Google Scholar
Pozzi, L., Nekaris, K. A. I., Perkin, A., Bearder, S. K., Pimley, E. R., Schulze, H., …, & Roos, C. (2015). Remarkable ancient divergences amongst neglected lorisiform primates. Zoological Journal of the Linnean Society, 175(3), 661674.Google Scholar
Prasetyo, D., & Sugardjito, J. (2011). Nest density as determinants for habitat utilizations of Bornean orangutan (Pongo pygmaeus wurmbii) in degraded forests of Gunung Palung National Park, West Kalimantan. Biodiversitas Journal of Biological Diversity, 12(3), 164170.Google Scholar
Preuschoft, H. (2002). What does ‘arboreal locomotion’ mean exactly and what are the relationships between’ climbing’, environment and morphology?. Zeitschrift Für Morphologie Und Anthropologie, 83, 171188.Google Scholar
Preuschoft, H., Godinot, M., Beard, C., Nieschalk, U., & Jouffroy, F. (1993). Biomechanical considerations to explain important morphological characters of primate hands. In Preuschoft, H. & Chivers, D. (Eds.), Hands of primates (pp. 245256). Vienna: Springer.Google Scholar
Preuschoft, H., Witte, H., & Fischer, M. (1995). Locomotion in nocturnal prosimians. In Alterman, L., Doyle, G., & Izard, M. (Eds.), Creatures of the dark: The nocturnal prosimians (pp. 453472). Boston, MA: Springer.CrossRefGoogle Scholar
Prufrock, K. A., Boyer, D. M., & Silcox, M. T. (2016a). The first major primate extinction: An evaluation of paleoecological dynamics of North American stem primates using a homology free measure of tooth shape. American Journal of Physical Anthropology, 159(4), 683697.Google Scholar
Prufrock, K. A., López-Torres, S., Silcox, M. T., & Boyer, D. M. (2016b). Surfaces and spaces: Troubleshooting the study of dietary niche space overlap between North American stem primates and rodents. Surface Topography: Metrology and Properties, 4(2), 024005.Google Scholar
Pullen, S. L. (2001). Behavioural and genetic studies of the mating system in a nocturnal primate: The lesser galago (Galago moholi) (unpublished doctoral dissertation). University of Cambridge, Cambridge, UK.Google Scholar
Puri, R. K. (2011). Documenting local environmental knowledge and change. In Newing, H. (Ed.), Conducting research in conservation: A social science perspective (pp. 146169). London: Routledge.Google Scholar
Putri, P. R., Moore, R., Andayani, N., & Sanchez, K. (2015). Effects of environmental enrichment on the social behaviour of Javan slow loris, Nycticebus javanicus. TAPROBANICA: The Journal of Asian Biodiversity, 7(1)129137.Google Scholar
Qi, G. Q. (1979). Pliocene mammals fauna of Lufeng, Yunnan. Vertebrata Palasiatica, 17(1), 1624.Google Scholar
Quade, J., Cater, J. M., Ojha, T. P., Adam, J., & Mark Harrison, T. (1995). Late Miocene environmental change in Nepal and the northern Indian subcontinent: Stable isotopic evidence from paleosols. Geological Society of America Bulletin, 107(12), 13811397.Google Scholar
Quan, G. Q., Jin, J. Y., Huang, J. S., & Zhou, Y. F. (1987). A new record of primate species in China. Acta Theriologica sinica, 7(2), 158.Google Scholar
Quinn, A., & Wilson, D. E. (2004). Daubentonia madagascariensis. Mammalian Species, 740, 116.Google Scholar
R Core Team (2016). A language and environment for statistical computing. Vienna: R Foundation for Statistical Computing.Google Scholar
Radespiel, U., Ehresmann, P., & Zimmermann, E. (2003). Species‐specific usage of sleeping sites in two sympatric mouse lemur species (Microcebus murinus and M. ravelobensis) in northwestern Madagascar. American Journal of Primatology, 59(4), 139151.Google Scholar
Radhakrishna, S., & Singh, M. (2002a). Home range and ranging pattern in the slender loris (Loris tardigradus lydekkerianus). Primates, 43(3), 237.Google Scholar
Radhakrishna, S., & Singh, M. (2002b). Social behaviour of the slender loris (Loris tardigradus lydekkerianus). Folia Primatologica73(4), 181196.Google Scholar
Radhakrishna, S., & Singh, M. (2002c). Conserving the slender loris (Loris lydekkerianus lydekkerianus). In Proceedings of the National Seminar on Conservation of Eastern Ghats Tirupati (pp. 181196). Hyderabad: Environment Protection Training Research Institute.Google Scholar
Radhakrishna, S., & Sinha, A. (2004). Population survey and conservation of the Bengal slow loris in Assam and Meghalaya, Northeastern India. New York, NY: Margot Marsh Biodiversity Foundation.Google Scholar
Radhakrishna, S., Kumara, H. N., & Sasi, R. (2011). Distribution patterns of slender loris subspecies (Loris lydekkerianus) in Kerala, southern India. International Journal of Primatology, 32(4), 10071019.Google Scholar
Radinsky, L. B. (1968). A new approach to mammalian cranial analysis, illustrated by examples of prosimian primates. Journal of Morphology, 124(2), 167179.Google Scholar
Raichlen, D. A., Pontzer, H., Shapiro, L. J., & Sockol, M. D. (2009). Understanding hind limb weight support in chimpanzees with implications for the evolution of primate locomotion. American Journal of Physical Anthropology, 138(4), 395402.Google Scholar
Rasmussen, D. T. (1986). Life history and behavior of slow lorises and slender lorises (unpublished doctoral dissertation). Duke University, Durham NC, USA.Google Scholar
Rasmussen, D. T. (1990a). Primate origins: Lessons from a neotropical marsupial. American Journal of Primatology, 22(4), 263277.Google Scholar
Rasmussen, D. T. (1990b). The phylogenetic position of Mahgarita stevensi: Protoanthropoid or lemuroid?. International Journal of Primatology, 11(5), 439469.Google Scholar
Rasmussen, D. T. (1994). The different meanings of a tarsioid–anthropoid clade and a new model of anthropoid origin. In Fleagle, J. G. & Kay, R. F. (Eds.), Anthropoid origins (pp. 335360). New York, NY: Plenum.CrossRefGoogle Scholar
Rasmussen, D. T., & Izard, M. K. (1988). Scaling of growth and life history traits relative to body size, brain size, and metabolic rate in lorises and galagos (Lorisidae, Primates). American Journal of Physical Anthropology75(3), 357367.Google Scholar
Rasmussen, D. T., & Nekaris, K. (1998). Evolutionary history of lorisiform primates. Folia Primatologica, 69(suppl. 1), 250285.Google Scholar
Rasmussen, T., & Simons, E. L. (1988). New specimens of Oligopithecus savagei, early Oligocene primate from the Fayum, Egypt. Folia Primatologica, 51(4), 182208.Google Scholar
Rasmussen, D. T., & Simons, E. L. (1992). Paleobiology of the oligopithecines, the earliest known anthropoid primates. International Journal of Primatology, 13(5), 477508.Google Scholar
Rasmussen, D. T., & Sussman, R. W. (2007). Parallelisms among primates and possums. In Ravosa, M. J. & Degasto, M. (Eds.), Primate origins and adaptations (pp. 775803). Boston, MA: Springer.Google Scholar
Rasoloharijaona, S., Rakotosamimanana, B., & Zimmermann, E. (2000). Infanticide by a male Milne-Edwards’ sportive lemur (Lepilemur edwardsi) in Ampijoroa, NW-Madagascar. International Journal of Primatology, 21(1), 4145.Google Scholar
Ratajszczak, R. (1998). Taxonomy, distribution and status of the lesser slow loris Nycticebus pygmaeus and their implications for captive management. Folia Primatologica, 69(suppl. 1), 171174.Google Scholar
Rattenborg, N. C., de la Iglesia, H. O., Kempenaers, B., Lesku, J. A., Meerlo, P., & Scriba, M. F. (2017). Sleep research goes wild: New methods and approaches to investigate the ecology, evolution and functions of sleep. Philosophical Transactions of the Royal Society B: Biological Sciences, 372(1734), 20160251.Google Scholar
Raubenheimer, D., & Rothman, J. M. (2013). Nutritional ecology of entomophagy in humans and other primates. Annual Review of Entomology, 58, 141160.Google Scholar
Ravosa, M. J. (1988). Browridge development in Cercopithecidae: A test of two models. American Journal of Physical Anthropology, 76(4), 535555.Google Scholar
Ravosa, M. J. (1991a). Interspecific perspective on mechanical and nonmechanical models of primate circumorbital morphology. American Journal of Physical Anthropology, 86(3), 369396.Google Scholar
Ravosa, M. J. (1991b). Ontogenetic perspective on mechanical and nonmechanical models of primate circumorbital morphology. American Journal of Physical Anthropology, 85(1), 95112.Google Scholar
Ravosa, M. J. (1998). Cranial allometry and geographic variation in slow lorises (Nycticebus). American Journal of Primatology, 45(3), 225243.Google Scholar
Ravosa, M. J. (2007). Cranial ontogeny, diet, and ecogeographic variation in African lorises. American Journal of Primatology, 69(1), 5973.Google Scholar
Ravosa, M. J., & Savakova, D. G. (2004). Euprimate origins: The eyes have it. Journal of Human Evolution, 46(3), 357.Google Scholar
Ravosa, M. J., Noble, V. E., Hylander, W. L., Johnson, K. R., & Kowalski, E. M. (2000). Masticatory stress, orbital orientation and the evolution of the primate postorbital bar. Journal of Human Evolution, 38(5), 667693.Google Scholar
Ravosa, M. J., Savakova, D. G., Johnson, K. R., & Hylander, W. L. (2006). Primate origins and the function of the circumorbital region: What’s load got to do with it? In Ravosa, M. J. & Dagosto, M. (Eds.), Primate origins: Adaptations and evolution (pp. 285328). New York, NY: Springer.Google Scholar
Ravosa, M. J., Daniel, A. N., & Costley, D. B. (2010). Allometry and evolution in the galago skull. Folia Primatologica, 81(4), 177196.Google Scholar
Rawlins, D. R., & Handasyde, K. A. (2002). The feeding ecology of the striped possum Dactylopsila trivirgata (Marsupialia: Petauridae) in far north Queensland, Australia. Journal of Zoology, 257(2), 195206.Google Scholar
Reddy, C. S., Jha, C. S., Dadhwal, V. K., Krishna, P. H., Pasha, S. V., Satish, K. V., …, & Diwakar, P. G. (2016). Quantification and monitoring of deforestation in India over eight decades (1930–2013). Biodiversity and Conservation, 25(1), 93116.Google Scholar
Reichard, U. H., Barelli, C., Hirai, H., & Nowak, M. G. (2016). The evolution of gibbons and siamang. In Reichard, U. H., Hirai, H., & Barelli, C. (Eds.), Evolution of gibbons and siamang: Phylogeny, morphology, and cognition (pp. 342). New York, NY: Springer Nature.Google Scholar
Reid, K. B., Colomb, M., Petry, F., & Loos, M. (2002). Complement component C1 and the collectins–first-line defense molecules in innate and acquired immunity. Trends in Immunology, 23(3), 115117.Google Scholar
Reinhardt, K.D. (2019). Ecophysiology of a wild nocturnal primate, the Javan slow loris (Nycticebus javanicus). PhD Dissertation. Oxford Brookes University.Google Scholar
Reinhardt, K. D., Wirdateti, , and Nekaris, K. A. I. (2016). Climate-mediated activity of the Javan slow loris, Nycticebus javanicus. AIMS Environmental Sciences, 3(2), 249260.Google Scholar
Reinhardt, K. D., Vyazovskiy, V. V., Hernandez-Aguilar, R. A., Wirdateti, , & Nekaris, K. A. I. (2019). Environment shapes activity and sleep patterns in a wild mammal. Scientific Reports, 9 , 9939.Google Scholar
Remis, M. J., & Robinson, C. A. J. (2012). Reductions in primate abundance and diversity in a multiuse protected area: Synergistic impacts of hunting and logging in a Congo basin forest. American Journal of Primatology, 74(7), 602612.Google Scholar
Resheff, Y. S., Rotics, S., Harel, R., Spiegel, O., & Nathan, R. (2014). AcceleRater: A web application for supervised learning of behavioral modes from acceleration measurements. Movement Ecology, 2(1), 27.Google Scholar
Reuter, M., Piller, W. E., Harzhauser, M., Mandic, O., Berning, B., Rögl, F., …, & Hamedani, A. (2009). The Oligo-/Miocene Qom Formation (Iran): Evidence for an early Burdigalian restriction of the Tethyan Seaway and closure of its Iranian gateways. International Journal of Earth Sciences, 98(3), 627650.Google Scholar
Rexstad, E., & Burnham, K. P. (1991). User’s guide for interactive program CAPTURE: Abundance estimation of closed animal populations. Fort Collins, CO: Colorado State University.Google Scholar
Richard, A. (1985). Primates in Nature. New York, NY: W. H. Freeman and Company.Google Scholar
Ripley, S. (1967). The leaping of langurs: A problem in the study of locomotor adaptation. American Journal of Physical Anthropology, 26(2), 149170.Google Scholar
Rist, J., Milner-Gulland, E. J., Cowlishaw, G. U. Y., & Rowcliffe, M. (2010). Hunter reporting of catch per unit effort as a monitoring tool in a bushmeat‐harvesting system. Conservation Biology, 24(2), 489499.Google Scholar
Roberts, R. G., Flannery, T. F., Ayliffe, L. K., Yoshida, H., Olley, J. M., Prideaux, G. J., …, & Smith, B. L. (2001). New ages for the last Australian megafauna: Continent-wide extinction about 46,000 years ago. Science, 292(5523), 18881892.Google Scholar
Robjent, H. (2005). Nocturnal primates of the Ituri forest DRC: A preliminary study involving distance sampling (unpublished Master’s dissertation). Oxford Brookes University, Oxford, UK.Google Scholar
Robson, C. (2011). Real world research (3rd ed.). Chichester: Wiley.Google Scholar
Rode-Margono, E. J., & Nekaris, K. A. I. (2014). Impact of climate and moonlight on a venomous mammal, the Javan slow loris (Nycticebus javanicus Geoffroy, 1812). Contributions to Zoology, 83(4), 217225.Google Scholar
Rode-Margono, J., & Nekaris, K. (2015). Cabinet of curiosities: Venom systems and their ecological function in mammals, with a focus on primates. Toxins, 7(7), 26392658.Google Scholar
Rode-Margono, E. J., Nijman, V., Wirdateti, N. K., & Nekaris, K. A. I. (2014). Ethology of the critically endangered Javan slow loris Nycticebus javanicus E. Geoffroy Saint-Hilaire in West Java. Asian Primates, 4, 2741.Google Scholar
Rogers, L. D., & Nekaris, K. A. I. (2011). Behaviour and habitat use of the Bengal slow loris Nycticebus bengalensis in the dry dipterocarp forests of Phnom Samkos Wildlife Sanctuary, Cambodia. Cambodian Journal of Natural History, 2001(2), 104112.Google Scholar
Rögl, F. (1997). Palaeogeographic considerations for Mediterranean and Paratethys seaways (Oligocene to Miocene). Annalen Des Naturhistorischen Museums in Wien, 99, 279310.Google Scholar
Rögl, F. (1999). Mediterranean and Paratethys: Facts and hypotheses of an Oligocene to Miocene paleogeography (short overview). Geologica Carpathica, 50(4), 339349.Google Scholar
Roos, C. (2003). Molekulare Phylogenie der Halbaffen, Schlankaffen, und Gibbons (dissertation). University of Munich, Munich, Germany.Google Scholar
Roos, C., Schmitz, J., & Zischler, H. (2004). Primate jumping genes elucidate strepsirrhine phylogeny. Proceedings of the National Academy of Sciences, 101(29), 1065010654.Google Scholar
Roos, C., Vu, N. T., Walter, L., & Nadler, T. (2007). Molecular systematics of Indochinese primates. Vietnamese Journal of Primatology, 1(1), 4153.Google Scholar
Roos, C., Boonratana, R., Supriatna, J., Fellowes, J. R., Groves, C. P., Nash, S. D., …, & Mittermeier, R. A. (2014). An updated taxonomy and conservation status review of Asian primates. Asian Primates Journal, 4, 238.Google Scholar
Ropert-Coudert, Y., Wilson, R. P., Grémillet, D., Kato, A., Lewis, S., & Ryan, P. G. (2006). Electrocardiogram recordings in free-ranging gannets reveal minimum difference in heart rate during flapping versus gliding flight. Marine Ecology Progress Series, 328, 275284.Google Scholar
Rose, K. D., Walker, A., & Jacobs, L. L. (1981). Function of the mandibular tooth comb in living and extinct mammals. Nature, 289(5798), 583585.Google Scholar
Rose, K. D., Rana, R. S., Sahni, A., & Smith, T. (2007). A new adapoid primate from the early Eocene of India. Contributions from the Museum of Paleontology, the University of Michigan, 31, 379385.Google Scholar
Rose, K. D., Chew, A. E., Dunn, R. H., Kraus, M. J., Fricke, H. C., & Zack, S. P. (2012). Earliest Eocene mammalian fauna from the Paleocene–Eocene Thermal Maximum at Sand Creek Divide, Southern Bighorn Basin, Wyoming. University of Michigan Papers on Paleontology, 36, 1122.Google Scholar
Rose, K. D., Dunn, R. H., Kumar, K., Perry, J. M., Prufrock, K. A., Rana, R. S., & Smith, T. (2018). New fossils from Tadkeshwar Mine (Gujarat, India) increase primate diversity from the early Eocene Cambay Shale. Journal of Human Evolution, 122, 93107.Google Scholar
Rose, L. M. (1994). Sex differences in diet and foraging behavior in white-faced capuchins (Cebus capucinus). International Journal of Primatology, 15(1), 95114.Google Scholar
Rose, R. W., Nevison, C. M., & Dixson, A. F. (1997). Testes weight, body weight and mating systems in marsupials and monotremes. Journal of Zoology, 243(3), 523531.Google Scholar
Rosenberger, A. L. (1978). Loss of incisor enamel in marmosets. Journal of Mammalogy, 59(1), 207208.Google Scholar
Rosenberger, A. L. (2010). Adaptive profile versus adaptive specialization: Fossils and gummivory in early primate evolution. In Burrows, A. M. & Nash, L. T. (Eds.), The evolution of exudativory in primates (pp. 273296). New York, NY: Springer.Google Scholar
Rosenberger, A. L. (2013). Fallback foods, preferred foods, adaptive zones, and primate origins. American Journal of Primatology, 75(9), 883890.Google Scholar
Rosenberger, A. L., & Strasser, E. (1985). Toothcomb origins: Support for the grooming hypothesis. Primates, 26(1), 7384.Google Scholar
Ross, C. F. (1992). Basal metabolic rate, body weight and diet in primates: An evaluation of the evidence. Folia Primatologica, 58(1), 723.Google Scholar
Ross, C. F. (1995). Allometric and functional influences on primate orbit orientation and the origins of the Anthropoidea. Journal of Human Evolution, 29(3), 201227.Google Scholar
Ross, C. F. (1996). Adaptive explanation for the origins of the Anthropoidea (Primates). American Journal of Primatology, 40(3), 205230.Google Scholar
Ross, C. F. (2000). Into the light: The origin of Anthropoidea. Annual Review of Anthropology, 29(1), 147194.Google Scholar
Ross, C. F., & Kirk, E. C. (2007). Evolution of eye size and shape in primates. Journal of Human Evolution, 52(3), 294313.Google Scholar
Ross, C. F., & Ravosa, M. J. (1993). Basicranial flexion, relative brain size, and facial kyphosis in non-human primates. American Journal of Physical Anthropology, 91(3), 305324.Google Scholar
Ross, C. F., Henneberg, M., Ravosa, M. J., & Richard, S. (2004). Curvilinear, geometric and phylogenetic modeling of basicranial flexion: Is it adaptive, is it constrained?. Journal of Human Evolution, 46(2), 185213.Google Scholar
Ross, S. R., Vreeman, V. M., & Lonsdorf, E. V. (2011). Specific image characteristics influence attitudes about chimpanzee conservation and use as pets. PLoS One, 6(7), e22050.Google Scholar
Rossie, J. B. (2006). Ontogeny and homology of the paranasal sinuses in Platyrrhini (Mammalia: Primates). Journal of Morphology, 267(1), 140.Google Scholar
Rossie, J. B., & MacLatchy, L. (2006). A new pliopithecoid genus from the early Miocene of Uganda. Journal of Human Evolution, 50(5), 568586.Google Scholar
Roth, L. S., Balkenius, A., & Kelber, A. (2008). The absolute threshold of colour vision in the horse. PLoS One, 3(11), e3711.Google Scholar
Rothman, J. M., Raubenheimer, D., Bryer, M. A., Takahashi, M., & Gilbert, C. C. (2014). Nutritional contributions of insects to primate diets: Implications for primate evolution. Journal of Human Evolution, 71, 5969.Google Scholar
Rowe, N. (1996). The Pictorial Guide to the Living Primates. New York, NY: Pogonias Press.Google Scholar
Rowe, N., & Myers, M. (2016). All the World’s Primates. Charlestown, RI: Pogonias Press.Google Scholar
Rowe, N., & Myers, M. (2017). All the World’s Primates. Charlestown, RI: Primate Conservation Inc.Google Scholar
Rowe, T. B., & Shepherd, G. M. (2016). Role of ortho‐retronasal olfaction in mammalian cortical evolution. Journal of Comparative Neurology, 524(3), 471495.Google Scholar
Rudd, R. L., & Stevens, G. L. (1994). A long-term capture–mark–release (CMR) and removal study of small mammals on Malaysian sub-montane rain forest. Wasmann Journal of Ecology, 50, 96150.Google Scholar
Ruf, T., & Geiser, F. (2015). Daily torpor and hibernation in birds and mammals. Biological Reviews, 90(3), 891926.Google Scholar
Ruf, T., Streicher, U., Stalder, G. L., Nadler, T., & Walzer, C. (2015). Hibernation in the pygmy slow loris (Nycticebus pygmaeus): Multiday torpor in primates is not restricted to Madagascar. Scientific Reports, 5, 17392.Google Scholar
Runestad, J. A. (1997). Postcranial adaptations for climbing in Loridae (Primates). Journal of Zoology, 242(2), 261290.Google Scholar
Sajeva, M., Augugliaro, C., Smith, M. J., & Oddo, E. (2013). Regulating internet trade in CITES species. Conservation Biology, 27(2), 429430.Google Scholar
Sakamoto, K. Q., Sato, K., Ishizuka, M., Watanuki, Y., Takahashi, A., Daunt, F., & Wanless, S. (2009). Can ethograms be automatically generated using body acceleration data from free-ranging birds?. PLoS One, 4(4), e5379.Google Scholar
Sallam, H. M., & Seiffert, E. R. (2016). New phiomorph rodents from the latest Eocene of Egypt, and the impact of Bayesian ‘clock’-based phylogenetic methods on estimates of basal hystricognath relationships and biochronology. PeerJ, 4, e1717.Google Scholar
Saltzstein, H. D., Roazzi, A., & Dias, M. D. G. (2003). The moral choices children attribute to adults and to peers: Implications for moral acquisition. European Journal of Psychology of Education, 18(3), 295307.Google Scholar
Sanderson, I. T. (1940). The mammals of the North Gameroons Forest Area: Being the results of the Percy Sladen Expedition to the Mamfe Division of the British Cameroons. The Transactions of the Zoological Society of London, 24(7), 623726.Google Scholar
Santini, L., Rojas, D., & Donati, G. (2015). Evolving through day and night: Origin and diversification of activity pattern in modern primates. Behavioral Ecology, 26(3), 789796.CrossRefGoogle Scholar
Sarmiento, E. (1998). The validity of ‘Pseudopotto martini. African Primates, 3, 4445.Google Scholar
Sasi, R., & Kumara, H. N. (2014). Distribution and relative abundance of the slender loris Loris lydekkerianus in southern Kerala, India. Primate Conservation, 2014(28), 165171.Google Scholar
Sauther, M. L., & Nash, L. T. (1987). Effect of reproductive state and body size on food consumption in captive Galago senegalensis braccatus. American Journal of Physical Anthropology, 73(1), 8188.Google Scholar
Savage, A., Guillen, R., Lamilla, I., & Soto, L. (2010). Developing an effective community conservation program for cotton‐top tamarins (Saguinus oedipus) in Colombia. American Journal of Primatology, 72(5), 379390.Google Scholar
Sawada, Y., Pickford, M., Senut, B., Itaya, T., Hyodo, M., Miura, T., …, & Fujii, H. (2002). The age of Orrorin tugenensis, an early hominid from the Tugen Hills, Kenya. Comptes Rendus Palevol, 1(5), 293303.Google Scholar
Schilling, A. (1979). Olfactory communication in prosimians. The Study of Prosimian Behavior, 461, 542.Google Scholar
Schilling, A. (2000). Organes des sens et communication chex le microcebe. Primatologie, 3, 85143.Google Scholar
Schiviz, A. N., Ruf, T., Kuebber‐Heiss, A., Schubert, C., & Ahnelt, P. K. (2008). Retinal cone topography of artiodactyl mammals: Influence of body height and habitat. Journal of Comparative Neurology, 507(3), 13361350.Google Scholar
Schmid, J. (2000). Daily torpor in the gray mouse lemur (Microcebus murinus) in Madagascar: Energetic consequences and biological significance. Oecologia, 123(2), 175183.Google Scholar
Schmid, J., & Kappeler, P. M. (2005). Physiological adaptations to seasonality in primates. In Brockmn, D. K. & Van Schaik, C. P. (Eds.), Seasonality in Primates (pp. 129156). Cambridge: Cambridge University Press.Google Scholar
Schmidt, D. A., Kerley, M. S., Porter, J. H., & Dempsey, J. L. (2005). Structural and nonstructural carbohydrate, fat, and protein composition of commercially available, whole produce. Zoo Biology, 24(4), 359373.Google Scholar
Schmitt, D. (1999). Compliant walking in primates. Journal of Zoology, 248(2), 149160.Google Scholar
Schmitt, D. (2003a). Evolutionary implications of the unusual walking mechanics of the common marmoset (C. jacchus). American Journal of Physical Anthropology, 122(1), 2837.Google Scholar
Schmitt, D. (2003b). Insights into the evolution of human bipedalism from experimental studies of humans and other primates. Journal of Experimental Biology, 206(9), 14371448.Google Scholar
Schmitt, D. (2012). Primate locomotor evolution: Biomechanical studies of primate locomotion and their implications for understanding primate neuroethology. In Platt, M. L. & Ghazanfar, A. A. (Eds.), Primate neuroethology (pp. 3163). Oxford: Oxford University Press.Google Scholar
Schmitt, D., & Hanna, J. B. (2004). Substrate alters forelimb to hindlimb peak force ratios in primates. Journal of Human Evolution, 46(3), 237252.Google Scholar
Schmitt, D., & Lemelin, P. (2002). Origins of primate locomotion: Gait mechanics of the woolly opossum. American Journal of Physical Anthropology, 118(3), 231238.Google Scholar
Schmitt, D., & Lemelin, P. (2004). Locomotor mechanics of the slender loris (Loris tardigradus). Journal of Human Evolution, 47(1–2), 8594.Google Scholar
Schmitt, D., Cartmill, M., Griffin, T. M., Hanna, J. B., & Lemelin, P. (2006). Adaptive value of ambling gaits in primates and other mammals. Journal of Experimental Biology, 209(11), 20422049.Google Scholar
Schoener, T. W. (1971). Theory of feeding strategies. Annual Review of Ecology and Systematics, 2(1), 369404.Google Scholar
Schoonaert, K., D’Août, K., Samuel, D., Talloen, W., Nauwelaerts, S., Kivell, T. L., & Aerts, P. (2016). Gait characteristics and spatio‐temporal variables of climbing in bonobos (Pan paniscus). American Journal of Primatology, 78(11), 11651177.Google Scholar
Schrauf, R. W., & Sanchez, J. (2008). Using freelisting to identify, assess, and characterize age differences in shared cultural domains. The Journals of Gerontology Series B: Psychological Sciences and Social Sciences, 63(6), S385S393.Google Scholar
Schreuder, H. T. (1994). Simplicity versus efficiency in sampling designs and estimation. Environmental Monitoring and Assessment, 33(3), 237245.Google Scholar
Schülke, O. (2003). To breed or not to breed: Food competition and other factors involved in female breeding decisions in the pair-living nocturnal fork-marked lemur (Phaner furcifer). Behavioral Ecology and Sociobiology, 55(1), 1121.Google Scholar
Schülke, O., & Kappeler, P. M. (2003). So near and yet so far: Territorial pairs but low cohesion between pair partners in a nocturnal lemur, Phaner furcifer. Animal Behaviour, 65(2), 331343.Google Scholar
Schultz, A. H. (1940). The size of the orbit and of the eye in primates. American Journal of Physical Anthropology, 26(1), 389408.Google Scholar
Schulze, H. (2019). Conservation database for lorises (Loris, Nycticebus) and pottos (Arctocebus, Perodicticus), prosimian primates. Retrieved 13 September 2019, from www.loris-conservation.org/database.Google Scholar
Schulze, H., & Groves, C. P. (2004). Asian lorises: Taxonomic problems caused by illegal trade.In Nadler, T., Streicher, U., & Ha, T. L. (Eds.), Conservation of primates in Vietnam (pp. 3336). Frankfurt: Frankfurt Zoological Society.Google Scholar
Schulze, H., & Meier, B. (1995a). Behavior of captive Loris tardigradus nordicus: A qualitative description, including some information about morphological bases of behavior. In Alterman, L., Doyle, G. A., & Izard, M. K. (Eds.), Creatures of the dark: The nocturnal prosimians (pp. 221249). Boston, MA: Springer.Google Scholar
Schulze, H., & Meier, B. (1995b). The subspecies of Loris tardigradus and their conservation status: A review. In Alterman, L., Doyle, G. A., & Izard, M. K. (Eds.), Creatures of the dark: The nocturnal prosimians (pp. 193209). Boston, MA: Springer.Google Scholar
Schwartz, J. H. (1992). Phylogenetic relationships of African and Asian lorisids. In Matano, S., Tuttle, R. H., Ishida, H., & Goodman, M. (Eds.), Topics in Primatology 3: Evolutionary Biology, Reproduction Endocrinology, and Virology (pp. 6581). Tokyo: Tokyo University Press.Google Scholar
Schwartz, J. H. (1996). Pseudopotto martini: A new genus and species of extant lorisiform primate. Anthropological Papers of the American Museum of Natural History, 78, 114.Google Scholar
Schwartz, J. H., & Beutel, J. C. (1995). Species diversity in lorisids: A preliminary analysis of Arctocebus, Perodicticus, and Nycticebus. In Alterman, L., Doyle, G. A., & Izard, M. K. (Eds.), Creatures of the dark: The nocturnal prosimians (pp.171192). Boston, MA: Springer.Google Scholar
Schwartz, J. H., & Tattersall, I. (1985). Evolutionary relationships of living lemurs and lorises(Mammalia, Primates) and their potential affinities with European Adapidae. Anthropological Papers of the American Museum of Natural History, 60, 1100.Google Scholar
Schwartz, K., & Flesness, N. (2014). On the origin of species information systems: An evolutionary perspective. In McGregor Reid, G. and Moore, G. (Eds.), History of zoos and aquariums: From royal gifts to biodiversity conservation (pp. 100106). Chester: North of England Zoological Society.Google Scholar
Schwarz, E. (1931). On the African short-tailed lemurs or pottos. Annals and Magazine of Natural History, 8(45), 249256.Google Scholar
Schwitzer, C., & Kaumanns, W. (2003). Foraging patterns of free-ranging and captive primates: Implications for captive feeding regimes. In Fidgett, A., Clauss, M., Gansloßer, U., Hatt, J.-M., & Nijboer, J. (Eds.), Zoo Animal Nutrition II (pp. 247265) Fürth: Filander Verlag.Google Scholar
Scordato, E. S., & Drea, C. M. (2007). Scents and sensibility: Information content of olfactory signals in the ringtailed lemur, Lemur catta. Animal Behaviour, 73(2), 301314.Google Scholar
Scott, P. W., Stevenson, M. F., & Cooper, J. E. (1999). Zoo standards review groups recommendations on revised secretary of states’s standards of modern zoo practice: Concordance. London: Department of the Environment, Transport and the Regions.Google Scholar
Seba, A. (1734 [2016]). Locupletissimi rerum naturalium thesauri. Republished as Cabinet of Natural Curiosities. Köln: Taschen.Google Scholar
Seiffert, E. R. (2007). Early evolution and biogeography of lorisiform strepsirrhines. American Journal of Primatology, 69(1), 2735.Google Scholar
Seiffert, E. R. (2012). Early primate evolution in Afro‐Arabia. Evolutionary Anthropology: Issues, News, and Reviews, 21(6), 239253.Google Scholar
Seiffert, E. R., Simons, E. L., & Attia, Y. (2003). Fossil evidence for an ancient divergence of lorises and galagos. Nature, 422(6930), 421424.Google Scholar
Seiffert, E. R., Simons, E. L., Ryan, T. M., & Attia, Y. (2005). Additional remains of Wadilemur elegans, a primitive stem galagid from the late Eocene of Egypt. Proceedings of the National Academy of Sciences, 102(32), 1139611401.Google Scholar
Seiffert, E. R., Perry, J. M., Simons, E. L., & Boyer, D. M. (2009). Convergent evolution of anthropoid-like adaptations in Eocene adapiform primates. Nature, 461(7267), 1118.Google Scholar
Seiffert, E. R., Simons, E. L., Boyer, D. M., Perry, J. M., Ryan, T. M., & Sallam, H. M. (2010). A fossil primate of uncertain affinities from the earliest late Eocene of Egypt. Proceedings of the National Academy of Sciences, 107(21), 97129717.Google Scholar
Seiffert, E. R., Costeur, L., & Boyer, D. M. (2015). Primate tarsal bones from Egerkingen, Switzerland, attributable to the middle Eocene adapiform Caenopithecus lemuroides. PeerJ, 3, e1036.Google Scholar
Seiffert, E. R., Boyer, D. M., Fleagle, J. G., Gunnell, G. F., Heesy, C. P., Perry, J. M., & Sallam, H. M. (2018). New adapiform primate fossils from the late Eocene of Egypt. Historical Biology, 30(1–2), 204226.Google Scholar
Selig, K. R., López-Torres, S., Hartstone-Rose, A., Nash, L. T., Burrows, A. M., & Silcox, M. T. (2019). A novel method for assessing enamel thickness distribution in the anterior dentition as a signal for gouging and other extractive foraging behaviors in gummivorous mammals. Folia Primatologica. DOI: 10.1159/000502819.Google Scholar
Sellers, W. I. (1996). A biomechanical investigation into the absence of leaping in the locomotor repertoire of the slender loris (Loris tardigradus). Folia Primatologica, 67(1), 114.Google Scholar
Sellers, W. I., & Crompton, R. H. (2004). Automatic monitoring of primate locomotor behaviour using accelerometers. Folia Primatologica, 75(4), 279293.Google Scholar
Sellers, W. I., Varley, J. S., & Waters, S. S. (1998). Remote monitoring of locomotion using accelerometers: A pilot study. Folia Primatologica, 69(Suppl. 1), 8285.Google Scholar
Senthilkumar, K., Mathialagan, P., Manivannan, C., Jayathangaraj, M. G., & Gomathinayagam, S. (2016). A study on the tolerance level of farmers toward human–wildlife conflict in the forest buffer zones of Tamil Nadu. Veterinary World, 9(7), 747.Google Scholar
Shahrul, A. M. S., & Ibrahim, J. (2006). Terrestrial Vertebrate Species of Penang National Park: Potential for Ecotourism. Kuala Lumpur: Percetakan Nasional Berhad.Google Scholar
Shapiro, L. J. (2007). Morphological and functional differentiation in the lumbar spine of lorisids and galagids. American Journal of Primatology, 69(1), 86102.Google Scholar
Shapiro, L. J., & Simons, C. V. (2002). Functional aspects of strepsirrhine lumbar vertebral bodies and spinous processes. Journal of Human Evolution, 42(6), 753783.Google Scholar
Sharpe, D. J., & Goldingay, R. L. (2007). Home range of the Australian squirrel glider, Petaurus norfolcensis (Diprotodontia). Journal of Mammalogy, 88(6), 15151522.Google Scholar
Shea, B. T. (1988). Phylogeny and skull form in the hominoid primates. In Schwartz, J. H. (Ed.), Orang-utan biology (pp. 233245). Oxford: Oxford University Press.Google Scholar
Shea, B. T. (1995). Ontogenetic scaling and size correction in the comparative study of primate adaptations. Anthropologie, 33(1–2), 116.Google Scholar
Shea, B. T., & Russell, M. D. (1986). On skull form and the supraorbital torus in primates. Current Anthropology, 27(3), 257259.Google Scholar
Sheng, H. L. (1999). Wild mammals in China. Beijing: China Forestry Press.Google Scholar
Shepard, E. L., Wilson, R. P., Quintana, F., Laich, A. G., Liebsch, N., Albareda, D. A., …, & Newman, C. (2008). Identification of animal movement patterns using tri-axial accelerometry. Endangered Species Research, 10, 4760.Google Scholar
Shepherd, C. R. (2006). The bird trade in Medan, North Sumatra: An overview. Birding Asia, 5, 1624.Google Scholar
Shepherd, C. R. (2010). Illegal primate trade in Indonesia exemplified by surveys carried out over a decade in North Sumatra. Endangered Species Research, 11(3), 201205.Google Scholar
Shepherd, C. R., & Nijman, V. (2007a). An assessment of wildlife trade at Mong La market on the Myanmar–China border. Traffic Bulletin, 21(2), 8588.Google Scholar
Shepherd, C. R., & Nijman, V. (2007b). An overview of the regulation of the freshwater turtle and tortoise pet trade in Jakarta, Indonesia. Petaling Jaya: TRAFFIC Southeast Asia.Google Scholar
Shepherd, C. R., Sukumaran, J., & Wich, S. A. (2004). Open season: An analysis of the pet trade in Medan, Sumatra 1997–2001. Petaling Jaya: TRAFFIC Southeast Asia.Google Scholar
Sherman, G. D., Haidt, J., & Coan, J. A. (2009). Viewing cute images increases behavioral carefulness. Emotion, 9(2), 282.Google Scholar
Shim, J. P., Dekleva, S., Guo, C., & Mittleman, D. (2011). Twitter, Google, iPhone/iPad, and Facebook (TGIF) and smart technology environments: How well do educators communicate with students via TGIF?. Communications of the Association for Information Systems, 29(1), 35.Google Scholar
Shine, R., Branch, W. R., Webb, J. K., Harlow, P. S., & Shine, T. (2006). Sexual dimorphism, reproductive biology, and dietary habits of psammophiine snakes (Colubridae) from southern Africa. Copeia, 2006(4), 650664.Google Scholar
Siemers, B. M., Goerlitz, H. R., Robsomanitrandrasana, E., Piep, M., Ramanamanjato, J. B., Rakotondravony, D., …, & Ganzhorn, J. U. (2007). Sensory basis of food detection in wild Microcebus murinus. International Journal of Primatology, 28(2), 291.Google Scholar
Sigé, B., Jaeger, J. J., Sudre, J., & Vianey-Liaud, M. (1990). Altiatlasius koulchii n. gen. et sp., primate omomyidé du Paléocène supérieur du Maroc, et les origines des euprimates. Palaeontographica Abteilung A, 214, 3156.Google Scholar
Silcox, M. T., & López-Torres, S. (2017). Major questions in the study of primate origins. Annual Review of Earth and Planetary Sciences, 45, 113137.Google Scholar
Silcox, M. T., Bloch, J. I., Boyer, D. M., Chester, S. G., & López‐Torres, S. (2017). The evolutionary radiation of plesiadapiforms. Evolutionary Anthropology: Issues, News, and Reviews, 26(2), 7494.Google Scholar
Silverman, D. (2011). Interpreting qualitative data (4th ed.). London: Sage.Google Scholar
Simonis, P., Rattal, M., Oualim, E. M., Mouhse, A., & Vigneron, J. P. (2014). Radiative contribution to thermal conductance in animal furs and other woolly insulators. Optics Express, 22(2), 19401951.Google Scholar
Simons, E. L. (1989). Description of two genera and species of Late Eocene Anthropoidea from Egypt. Proceedings of the National Academy of Sciences, 86(24), 99569960.Google Scholar
Simons, E. L. (1997). Discovery of the smallest Fayum Egyptian primates (Anchomomyini, Adapidae). Proceedings of the National Academy of Sciences, 94(1), 180184.Google Scholar
Simons, E. L., & Rasmussen, D. T. (1989). Cranial morphology of Aegyptopithecus and Tarsius and the question of the tarsier–anthropoidean clade. American Journal of Physical Anthropology, 79(1), 123.Google Scholar
Simons, E. L., & Rasmussen, D. T. (1994a). A remarkable cranium of Plesiopithecus teras (Primates, Prosimii) from the Eocene of Egypt. Proceedings of the National Academy of Sciences, 91(21), 99469950.Google Scholar
Simons, E. L., & Rasmussen, T. (1994b). A whole new world of ancestors: Eocene anthropoideans from Africa. Evolutionary Anthropology: Issues, News, and Reviews, 3(4), 128139.Google Scholar
Simpson, G. G. (1967). The Teritary lorisiform primates of Africa. Bulletin of the Museum of Comparative Zoology, 136, 3962.Google Scholar
Simpson, G. M., Fuller, G., Lukas, K. E., Kuhar, C. W., Fitch‐Snyder, H., Taylor, J., & Dennis, P. M. (2018). Sources of morbidity in lorises and pottos in North American zoos: A retrospective review, 1980–2010. Zoo Biology, 37(4), 245257.Google Scholar
Singh, M., Lindburg, D. G., Udhayan, A., Kumar, M. A., & Kumara, H. N. (1999). Status survey of slender loris Loris tardigradus lydekkerianus in Dindigul, Tamil Nadu, India. Oryx, 33(1), 3137.Google Scholar
Singh, M., Kumar, M. A., Kumara, H. N., & Mohnot, S. M. (2000). Distribution and conservation of slender lorises (Loris tardigradus lydekkerianus) in southern Andhra Pradesh, South India. International Journal of Primatology, 21(4), 721730.Google Scholar
Singh, Y. (2015). Man–animal conflict: A study of human–wildlife conflicts in eastern Vidarbha Region of Maharashtra. Indian Streams Research Journal, 5(8), 2135.Google Scholar
Skinner, D., Metcalf, C. A., Seager, J. R., De Swardt, J. S., & Laubscher, J. A. (1991). An evaluation of an education programme on HIV infection using puppetry and street theatre. AIDS Care, 3(3), 317329.Google Scholar
Slik, J. F., Franklin, J., Arroyo-Rodríguez, V., Field, R., Aguilar, S., Aguirre, N., …, & Avella, A. (2018). Phylogenetic classification of the world’s tropical forests. Proceedings of the National Academy of Sciences, 115(8), 18371842.Google Scholar
Smeenk, C., Godfrey, L. R., & Williams, F. L. (2006). The early specimens of the potto Perodicticus potto (Statius Muller, 1776) in the National Museum of Natural History, Leiden, with the selection of a neotype. Zoologische Mededelingen, 80(4), 139.Google Scholar
Smith, A. C. (2010). Exudativory in primates: Interspecific patterns. In Burrows, A. M. & Nash, L. T. (Eds.), The evolution of exudativory in primates (pp. 4588). New York, NY: Springer.Google Scholar
Smith, J. A. (1860). Notice of the ‘Angwátibo’ of Old Calabar. Proceedings of the Royal Physical Society, Edinburgh, 2, 172–92.Google Scholar
Smith, J. J., & Borgatti, S. P. (1997). Salience counts-and so does accuracy: Correcting and updating a measure for free-list-item salience. Journal of Linguistic Anthropology, 7(2), 208209.Google Scholar
Smith, K. F., Behrens, M., Schloegel, L. M., Marano, N., Burgiel, S., & Daszak, P. (2009). Reducing the risks of the wildlife trade. Science, 324(5927), 594595.Google Scholar
Smith, R. J., & Jungers, W. L. (1997). Body mass in comparative primatology. Journal of Human Evolution, 32(6), 523559.Google Scholar
Smith, T. D., & Bhatnagar, K. P. (2004). Microsmatic primates: Reconsidering how and when size matters. The Anatomical Record Part B, 279(1), 2431.Google Scholar
Smith, T. D., & Rossie, J. B. (2006). Primate olfaction: Anatomy and evolution. In Brewer, W. J., Castle, D., & Pantelis, C. (Eds.), Olfaction and the brain (pp. 135166). Cambridge: Cambridge University Press.Google Scholar
Smith, T. D., & Rossie, J. B. (2008). Nasal fossa of mouse and dwarf lemurs (Primates, Cheirogaleidae). The Anatomical Record, 291(8), 895915.Google Scholar
Smith, T. D., Rose, K. D., & Gingerich, P. D. (2006). Rapid Asia–Europe–North America geographic dispersal of earliest Eocene primate Teilhardina during the Paleocene–Eocene thermal maximum. Proceedings of the National Academy of Sciences, 103(30), 1122311227.Google Scholar
Smith, T. D., Alport, L. J., Burrows, A. M., Bhatnagar, K. P., Dennis, J. C., Tuladhar, P., & Morrison, E. E. (2007a). Perinatal size and maturation of the olfactory and vomeronasal neuroepithelia in lorisoids and lemuroids. American Journal of Primatology, 69, 7485.Google Scholar
Smith, T. D., Rossie, J. B., & Bhatnagar, K. P. (2007b). Evolution of the nose and nasal skeleton in primates. Evolutionary Anthropology: Issues, News, and Reviews, 16(4), 132146.Google Scholar
Smith, T. D., Eiting, T. P., & Rossie, J. B. (2011). Distribution of olfactory and nonolfactory surface area in the nasal fossa of Microcebus murinus: Implications for microcomputed tomography and airflow studies. The Anatomical Record, 294(7), 12171225.Google Scholar
Smith, T. D., Eiting, T. P., Bonar, C. J., & Craven, B. A. (2014). Nasal morphometry in marmosets: Loss and redistribution of olfactory surface area. The Anatomical Record, 297(11), 20932104.Google Scholar
Smith, T. D., Eiting, T. P., & Bhatnagar, K. P. (2015a). Anatomy of the nasal passages in mammals. In Doty, R. L. (Ed.), Handbook of olfaction and gustation (pp. 3961) (3rd ed.). Hoboken, NJ: Wiley.Google Scholar
Smith, T. D., Muchlinski, M. N., Bhatnagar, K. P., Durham, E. L., Bonar, C. J., & Burrows, A. M. (2015b). The vomeronasal organ of Lemur catta. American Journal of Primatology, 77(2), 229238.Google Scholar
Smith, T. D., Martell, M. C., Rossie, J. B., Bonar, C. J., & Deleon, V. B. (2016). Ontogeny and microanatomy of the nasal turbinals in lemuriformes. The Anatomical Record, 299(11), 14921510.Google Scholar
Smithers, R. H. N. (1971). The mammals of Botswana (unpublished doctoral dissertation). University of Pretoria, Pretoria, South Africa.Google Scholar
Snijders, L., Weme, L. E. N., de Goede, P., Savage, J. L., van Oers, K., & Naguib, M. (2017). Context‐dependent effects of radio transmitter attachment on a small passerine. Journal of Avian Biology, 48(5), 650659.Google Scholar
Sobel, N., Prabhakaran, V., Desmond, J. E., Glover, G. H., Goode, R. L., Sullivan, E. V., & Gabrieli, J. D. E. (1998). Sniffing and smelling: Separate subsystems in the human olfactory cortex. Nature, 392(6673), 282.Google Scholar
Sodhi, N. S., Posa, M. R. C., Lee, T. M., Bickford, D., Koh, L. P., & Brook, B. W. (2010). The state and conservation of Southeast Asian biodiversity. Biodiversity and Conservation, 19(2), 317328.Google Scholar
Sokal, R. R., & Rohlf, F. J.. (2012). Biometry: The principles and practice of statistics in biological research (4th ed.). New York, NY: W. H. Freeman and Company.Google Scholar
Sokolov, V. E., Ushakova, N. A., Telitsyna, A. Y., & Meshik, V. A. (1993). Some peculiarities of reproduction of lesser slow loris Nycticebus pygmaeus coucang in Moscow Zoo. Izvestiya Akademii Nauk, Seriya Biogicheskaya, 2, 219–26.Google Scholar
Sonnini, C. S. (1803). Histoire Naturelle Générale et Particulière. Paris: F. Dufart.Google Scholar
Southern, H. N. (1955). Nocturnal animals. Scientific American, 193(4), 8899.Google Scholar
Southgate, D., Westoby, N., & Page, G. (2010). Creative determinants of viral video viewing. International Journal of Advertising, 29(3), 349368.Google Scholar
Species360. (2018). Zoological Information Management System (ZIMS). Retrieved 13 September 2019, from http://zims.Species360.org.Google Scholar
Spellerberg, I. F., & Fedor, P. J. (2003). A tribute to Claude Shannon (1916–2001) and a plea for more rigorous use of species richness, species diversity and the ‘Shannon–Wiener’ Index. Global Ecology and Biogeography, 12(3), 177179.Google Scholar
Spoor, F., Garland, T., Krovitz, G., Ryan, T. M., Silcox, M. T., & Walker, A. (2007). The primate semicircular canal system and locomotion. Proceedings of the National Academy of Sciences, 104(26), 1080810812.Google Scholar
Springer, M. S., Meredith, R. W., Gatesy, J., Emerling, C. A., Park, J., Rabosky, D. L., …, & Murphy, W. J. (2012). Macroevolutionary dynamics and historical biogeography of primate diversification inferred from a species supermatrix. PLoS One, 7(11), e49521.Google Scholar
Sri Lanka Environment Office (SLEO). (2009). Ministry of Environment (Book 3 Part 1). Retrieved from www.environmentmin.gov.lk/web/pdf/annual_reports/Book-3-Part-1.pdf.Google Scholar
Starr, C. R. (2011a). The conservation and ecology of the pygmy slow loris (Nycticebus pygmaeus) in Eastern Cambodia (unpublished doctoral dissertation). University of Queensland, Queensland, Australia.Google Scholar
Starr, C. (2011b). The conservation and ecology of the pygmy slow loris (Nycticebus pygmaeus) in Eastern Cambodia (unpublished doctoral dissertation). Oxford Brookes University, Oxford, UK.Google Scholar
Starr, C., & Nekaris, K. A. I. (2013). Obligate exudativory characterizes the diet of the pygmy slow loris Nycticebus pygmaeus. American Journal of Primatology, 75(10), 10541061.Google Scholar
Starr, C., Nekaris, K. A. I., Streicher, U., & Leung, L. (2010). Traditional use of slow lorises Nycticebus bengalensis and N. pygmaeus in Cambodia: An impediment to their conservation. Endangered Species Research, 12(1), 1723.Google Scholar
Starr, C., Nekaris, K. A. I., Streicher, U., & Leung, L. K. P. (2011). Field surveys of the Vulnerable pygmy slow loris Nycticebus pygmaeus using local knowledge in Mondulkiri Province, Cambodia. Oryx, 45(1), 135142.Google Scholar
Starr, C., Nekaris, K. A. I., & Leung, K. P. (2012a). A comparison of three survey methods for detecting the elusive pygmy slow loris Nycticebus pygmaeus in Eastern Cambodia. Cambodian Journal of Natural History, 2012(2), 123130.Google Scholar
Starr, C., Nekaris, K. A. I., & Leung, L. (2012b). Hiding from the moonlight: Luminosity and temperature affect activity of Asian nocturnal primates in a highly seasonal forest. PLoS One, 7(4), e36396.Google Scholar
Statistical Year Book Bangladesh. (2017). Statistics & Informatics Division (Sid), Ministry of Planning Government of the People’s Republic of Bangladesh. Dhaka: Bangladesh Bureau of Statistics.Google Scholar
Stauffer, H. B., Ralph, C. J., & Miller, S. L. (2004). Ranking habitat for marbled murrelets: New conservation approach for species with uncertain detection. Ecological Applications, 14(5), 13741383.Google Scholar
Steel, E. A. (1994). Study of the value and volume of bushmeat commerce in Gabon. Libreville: WWF Programme pour le Gabon.Google Scholar
Sterling, E. J. (1993). Patterns of range use and social organization in aye-ayes (Daubentonia madagascariensis) on Nosy Mangabe. In Kappeler, P. M. & Ganzhorn, J. U. (Eds.), Lemur social systems and their ecological basis (pp. 110). Boston, MA: Springer.Google Scholar
Sterling, E. J., Nguyen, N. G. A., & Fashing, P. J. (2000). Spatial patterning in nocturnal prosimians: A review of methods and relevance to studies of sociality. American Journal of Primatology, 51(1), 319.Google Scholar
Stevens, C. E., & Hume, I. D. (2001). Comparative physiology of the vertebrate digestive system (2nd ed.). New York, NY: Cambridge University Press.Google Scholar
Stevens, N. J. (2006). Stability, limb coordination and substrate type: The ecorelevance of gait sequence pattern in primates. Journal of Experimental Zoology Part A: Comparative Experimental Biology, 305(11), 953963.Google Scholar
Stevens, N. J., & Heesy, C. P. (2006). Malagasy primate origins: Phylogenies, fossils, and biogeographic reconstructions. Folia Primatologica, 77(6), 419433.Google Scholar
Stevens, N. J., & Heesy, C. P. (2012). Head posture and visual orientation in Loris tardigradus during locomotion on oblique supports. In Masters, J., Gamba, M., & Génin, F. (Eds.), Leaping ahead (pp. 97104). New York, NY: Springer.Google Scholar
Stone, W., & Rehn, J. A. G. (1902). A collection of mammals from Sumatra, with a review of the genera Nycticebus and Tragulus. Proceedings of the Academy of Natural Sciences of Philadelphia, 54, 127142.Google Scholar
Straus, W. L. Jr (1942). Rudimentary digits in primates. The Quarterly Review of Biology, 17(3), 228243.Google Scholar
Strehlow, H. (2004). Mammals in the Berlin Aquarium at the time of Alfred Edmund Brehm. Milu, 11(3), 229239.Google Scholar
Streicher, U. (2003). Seasonal variation in fur coloring and markings in the Pygmy loris (Nycticebus pygmaeus) and its taxonomic significance. Zoologische Garten, 73, 368373.Google Scholar
Streicher, U. (2004). Aspects of ecology and conservation of the pygmy loris Nycticebus pygmaeus in Vietnam (doctoral dissertation). Ludwig-Maximilians-University of Munich, Munich, Germany.Google Scholar
Streicher, U. (2005). Seasonal body weight changes in pygmy lorises Nycticebus pygmaeus. Verhandlungsber: Zootierkrk, 42, 144145.Google Scholar
Streicher, U. (2009). Diet and feeding behaviour of pygmy lorises (Nycticebus pygmaeus) in Vietnam. Vietnamese Journal of Primatology, 3, 3744.Google Scholar
Streicher, U., & Nadler, T. (2003). Re-introduction of pygmy lorises in Vietnam. Reintroduction News, 23, 3740.Google Scholar
Streicher, U., Singh, M., Timmins, R. J., & Brockelman, W. (2008a). Nycticebus bengalensis. IUCN Red List of Threatened Species 2008. Retrieved 15 October 2018, from http://dx.doi.org/10.2305/IUCN.UK.2008.RLTS.T39758A10263081.en.Google Scholar
Streicher, U., Vu, T. N., Nadler, T., Timmins, R. J., & Nekaris, A. (2008b). Nycticebus pygmaeus. IUCN Red List of Threatened Species v.2010.2. Retrieved 12 July 2010, from http://iucnredlist.org.Google Scholar
Streicher, U., Wilson, A., Collins, R. L., & Nekaris, K. A. I. (2012a). Exudates and animal prey characterize slow loris (Nycticebus pygmaeus, N. coucang and N. javanicus) diet in captivity and after release into the wild. In Masters, J., Gamba, M., & Génin, F. (Eds.), Leaping ahead (pp. 165172). New York, NY: Springer.Google Scholar
Streicher, U., Collins, R., Wilson, A., & Nekaris, K. A. I. (2012b). Observations on the feeding preferences of slow lorises (N. pygmaeus, N. javanicus, N. coucang) confiscated from the trade. In Masters, J., Genin, F., & Crompton, R. (Eds.), Leaping ahead (pp. 165172). New York, NY: Springer.Google Scholar
Streicher, U., Nowack, J., Stalder, G., Walzer, C., Nadler, T., & Ruf, T. (2017). Hibernation in pygmy lorises (Nycticebus pygmaeus): What does it mean?. Vietnamese Journal of Primatology, 2, 5157.Google Scholar
Strier, K. (2003). Primate behavioural ecology. Abingdon: Routledge.Google Scholar
Su, B., Wang, W., & Zhang, Y. P. (1998). Protein polymorphism and genetic divergence in slow loris (genus Nycticebus). Primates, 39(1), 7984.Google Scholar
Sumner, P., & Mollon, J. D. (2003). Colors of primate pelage and skin: Objective assessment of conspicuousness. American Journal of Primatology, 59(2), 6791.Google Scholar
Supriatna, J., & Ario, A. (2015). Primates as flagships for conserving biodiversity and parks in Indonesia: Lessons learned from West Java and North Sumatra. Primate Conservation, 2015(29), 123132.Google Scholar
Sussman, R. W. (1974). Ecological distinctions in sympatric species of Lemur. In Martin, R. D., Doyle, G. A., & Walker, A. C. (Eds.), Prosimian biology (pp. 7518). London: Duckworth.Google Scholar
Sussman, R. W. (1991). Primate origins and the evolution of angiosperms. American Journal of Primatology, 23(4), 209223.Google Scholar
Sussman, R. W. (1999). Primate ecology and social structure. Volume 1: lorises, lemurs and tarsiers. Needham Heights, MA: Pearson Custom Publishing.Google Scholar
Sussman, R. W., Rasmussen, D. T., & Raven, P. H. (2013). Rethinking primate origins again. American Journal of Primatology, 75(2), 95106.Google Scholar
Suwa, G., Kono, R. T., Katoh, S., Asfaw, B., & Beyene, Y. (2007). A new species of great ape from the late Miocene epoch in Ethiopia. Nature, 448(7156), 921.Google Scholar
Svensson, M. S., & Friant, S. C. (2014). Threats from trading and hunting of pottos and angwantibos in Africa resemble those faced by slow lorises in Asia. Endangered Species Research, 23(2), 107114.Google Scholar
Svensson, M. S., Ingram, D. J., Nekaris, K. A. I., & Nijman, V. (2016). Trade and ethnozoological use of African lorisiformes in the last 20 years. Hystrix, 26(2), 153161.Google Scholar
Svensson, M. S., Bersacola, E., Mills, M. S., Munds, R. A., Nijman, V., Perkin, A., …, & Bearder, S. K. (2017). A giant among dwarfs: A new species of galago (Primates: Galagidae) from Angola. American Journal of Physical Anthropology, 163(1), 3043.Google Scholar
Svensson, M. S., Nekaris, K. A. I., Bearder, S. K., Bettridge, C. M., Butynski, T. M., Cheyne, S. M., …, & Nijman, V. (2018). Sleep patterns, daytime predation, and the evolution of diurnal sleep site selection in lorisiforms. American Journal of Physical Anthropology, 166(3), 563577.Google Scholar
Swapna, N., Gupta, A., & Radhakrishna, S. (2008). Distribution survey of Bengal slow loris Nycticebus bengalensis in Tripura, northeastern India. Asian Primates Journal, 1(1), 3740.Google Scholar
Swapna, N., Radhakrishna, S., Gupta, A. K., & Kumar, A. (2009). Exudativory in the Bengal slow loris (Nycticebus bengalensis) in Trishna Wildlife Sanctuary, Tripura, Northeast India. American Journal of Primatology, 71, 19.Google Scholar
Swapna, N., Radhakrishna, S., Gupta, A. K., & Kumar, A. (2010). Exudativory in the Bengal slow loris (Nycticebus bengalensis) in Trishna Wildlife Sanctuary, Tripura, northeast India. American Journal of Primatology, 72(2), 113121.Google Scholar
Swindler, D. R. (2002). Primate dentition. Cambridge: Cambridge University Press.Google Scholar
Sykes, S. M., Robison, W. G., Waxler, M., & Kuwabara, T. (1981). Damage to the monkey retina by broad-spectrum fluorescent light. Investigative Ophthalmology & Visual Science, 20(4), 425434.Google Scholar
Szalay, F. S., & Delson, E. (1979). Evolutionary history of the primates. New York, NY: Academic Press.Google Scholar
Szalay, F. S., & Seligsohn, D. (1977). Why did the strepsirhine tooth comb evolve?. Folia Primatologica, 27(1), 7582.Google Scholar
Tabuce, R., Marivaux, L., Lebrun, R., Adaci, M., Bensalah, M., Fabre, P. H., …, & Mahboubi, M. (2009). Anthropoid versus strepsirhine status of the African Eocene primates Algeripithecus and Azibius: Craniodental evidence. Proceedings of the Royal Society B: Biological Sciences, 276(1676), 40874094.Google Scholar
Tan, C. L. (1994). Survey of Nycticebus pygmaeus in southern Vietnam In Handbook and Abstracts of 15th IPS Congress, Bali (pp. 136). n.p.: International Primatological Society.Google Scholar
Tan, C. L., & Drake, J. H. (2001). Evidence of tree gouging and exudate eating in pygmy slow lorises (Nycticebus pygmaeus). Folia Primatologica, 72(1), 3739.Google Scholar
Tan, Y., Yoder, A. D., Yamashita, N., & Li, W. H. (2005). Evidence from opsin genes rejects nocturnality in ancestral primates. Proceedings of the National Academy of Sciences, 102(41), 1471214716.Google Scholar
Tandler, B., Pinkstaff, C. A., & Phillips, C. J. (2006). Interlobular excretory ducts of mammalian salivary glands: Structural and histochemical review. The Anatomical Record Part A, 288(5), 498526.Google Scholar
Tattersall, I. (2005). Mechanisms of faunal origin and diversity in island environments: The case of Madagascar’s mammals. Hellenic Journal of Geosciences, 41, 3546.Google Scholar
Tattersall, I. (2006). Origin of the Malagasy strepsirhine primates. In Gould, L. & Sauther, M. L. (Eds.), Lemurs: Ecology and adaptation (pp. 317). New York, NY: Springer.Google Scholar
Tattersall, I. (2008). Vicariance vs. dispersal in the origin of the Malagasy mammal fauna. In Fleagle, J. G. & Gilbert, C. C. (Eds.), Elwyn Simons: A search for origins (pp. 397408). New York, NY: Springer.Google Scholar
Taylor, G., Scharlemann, J. P. W., Rowcliffe, M., Kümpel, N., Harfoot, M. B. J., Fa, J. E., …, & Coad, L. M. (2015). Synthesising bushmeat research effort in West and Central Africa: A new regional database. Biological Conservation, 181, 199205.Google Scholar
Taylor, N., & Signal, T. D. (2005). Empathy and attitude to animals. Anthrozöos, 18(1), 1827.Google Scholar
Thạch, H. M., Le, M. D., , N. B., Panariello, A., Sethi, G., Sterling, E. J., & Blair, M. E. (2018). Slow loris trade in Vietnam: Exploring diverse knowledge and values. Folia Primatologica, 89(1), 4562.Google Scholar
Thalmann, U. (2001). Food resource characteristics in two nocturnal lemurs with different social behavior: Avahi occidentalis and Lepilemur edwardsi. International Journal of Primatology, 22(2), 287324.Google Scholar
Than, N. T. (2001). Awareness of Vietnamese primary schoolteachers on environmental education. International Research in Geographical and Environmental Education, 10(4), 429444.Google Scholar
Thomas, O. (1908). The nomenclature of certain lorises. Annals and Magazine of Natural History Series, 8(1), 467469.Google Scholar
Thomas, O. (1910). A new potto from British East Africa. In Proceedings of the Zoological Society of London: January–March General Meeting for Science Business (pp. 536–537).Google Scholar
Thomas, O. (1921). Two new species of slow-loris. Journal of Natural History, 8(48), 627628.Google Scholar
Thompson, E. C., & Juan, Z. (2006). Comparative cultural salience: Measures using free-list data. Field Methods, 18(4), 398412.Google Scholar
Thompson, S. D. (1982). Spatial utilization and foraging behavior of the desert woodrat, Neotoma lepida lepida. Journal of Mammalogy, 63(4), 570581.Google Scholar
Tilden, C. D., & Oftedal, O. T. (1997). Milk composition reflects pattern of material care in prosimian primates. American Journal of Primatology, 41(3), 195211.Google Scholar
Timm, R. M., & Birney, E. C. (1992). Systematic notes on the Philippine slow loris, Nycticebus coucang menagensis (Lydekker, 1893) (Primates: Lorisidae). International Journal of Primatology, 13(6), 679686.Google Scholar
Toussaint, S., Herrel, A., Ross, C. F., Aujard, F., & Pouydebat, E. (2015). Substrate diameter and orientation in the context of food type in the gray mouse lemur, Microcebus murinus: Implications for the origins of grasping in primates. International Journal of Primatology, 36(3), 583604.Google Scholar
Trent, B. K., Tucker, M. E., & Lockard, J. S. (1977). Activity changes with illumination in slow loris Nycticebus coucang. Applied Animal Ethology, 3(3), 281286.Google Scholar
Trouessart, E.-L. (1898–1899). Catalogus mammalium tam viventium quam fossilium. Berolini: R. Friedlander & Sohn.Google Scholar
Tyre, A. J., Tenhumberg, B., Field, S. A., Niejalke, D., Parris, K., & Possingham, H. P. (2003). Improving precision and reducing bias in biological surveys: Estimating false‐negative error rates. Ecological Applications, 13(6), 17901801.Google Scholar
Vagell, R., Vagell, V. J., Jacobs, R. L., Gordon, J., & Baden, A. (2019). SMARTA: Automated testing apparatus for visual discrimination tasks. Behavior research Methods. DOI: https://doi.org/10.3758/s13428-018-1113-9.Google Scholar
Valenta, K., Burke, R. J., Styler, S. A., Jackson, D. A., Melin, A. D., & Lehman, S. M. (2013). Colour and odour drive fruit selection and seed dispersal by mouse lemurs. Scientific Reports, 3, 2424.Google Scholar
Valkenburg, P. M., Peter, J., & Schouten, A. P. (2006). Friend networking sites and their relationship to adolescents’ well-being and social self-esteem. CyberPsychology & Behavior, 9(5), 584590.Google Scholar
Van Campen, F., & Van der Hoeven, J. (1859). Ontleedkundig onderzoek van den potto van Bosman. Amsterdam: Van Der Post.Google Scholar
van der Linden, S. (2017). The nature of viral altruism and how to make it stick. Nature Human Behaviour, 1(41), 13.Google Scholar
van der Ploeg, J., Cauillan-Cureg, M., van Weerd, M., & Persoon, G. (2011). ‘Why must we protect crocodiles?’ Explaining the value of the Philippine crocodile to rural communities. Journal of Integrative Environmental Sciences, 8(4), 287298.Google Scholar
van der Sandt, L. (2016). Towards a successful translocation of captive slow lorises (Nycticebus spp.) in Borneo: A review and recommendations. bioRxiv: 078535.Google Scholar
van Geffen, K. G., van Grunsven, R. H., van Ruijven, J., Berendse, F., & Veenendaal, E. M. (2014). Artificial light at night causes diapause inhibition and sex‐specific life history changes in a moth. Ecology and Evolution, 4(11), 20822089.Google Scholar
Van Hees, V. T., Sabia, S., Anderson, K. N., Denton, S. J., Oliver, J., Catt, M., …, & Singh-Manoux, A. (2015). A novel, open access method to assess sleep duration using a wrist-worn accelerometer. PLoS One, 10(11), e0142533.Google Scholar
Van Oort, B. E., Tyler, N. J., Storeheier, P. V., & Stokkan, K. A. (2004). The performance and validation of a data logger for long-term determination of activity in free-ranging reindeer, Rangifer tarandus L. Applied Animal Behaviour Science, 89(3–4), 299308.Google Scholar
Van Schaik, C. P., & Van Hooff, J. (1983). On the ultimate causes of primate social systems. Behaviour, 85(1), 91117.Google Scholar
Van Schaik, C. P., Wich, S. A., Utami, S. S., & Odom, K. (2005). A simple alternative to line transects of nests for estimating orangutan densities. Primates, 46(4), 249254.Google Scholar
Van Someren, E. J., Swaab, D. F., Colenda, C. C., Cohen, W., McCall, W. V., & Rosenquist, P. B. (1999). Bright light therapy: Improved sensitivity to its effects on rest–activity rhythms in Alzheimer patients by application of nonparametric methods. Chronobiology International, 16(4), 505518.Google Scholar
Van Valen, L., & Sloan, R. E. (1965). The earliest primates. Science, 150(3697), 743745.Google Scholar
Van Valkenburgh, B., Smith, T. D., & Craven, B. A. (2014). Tour of a labyrinth: Exploring the vertebrate nose. The Anatomical Record, 297(11), 19751984.Google Scholar
van Vliet, N., & Mbazza, P. (2011). Recognizing the multiple reasons for bushmeat consumption in urban areas: A necessary step toward the sustainable use of wildlife for food in Central Africa. Human Dimensions of Wildlife, 16(1), 4554.Google Scholar
Vázquez, A. G., Musing, L., Nekaris, K. A. I., & Martínez, I. J. (2016). El efecto de los vídeos de YouTube en la percepción de especies exóticas como mascotas potenciales. Chronica Naturae, 6, 1323.Google Scholar
Veilleux, C. C. (2016). Genetics of primate color vision. In Fuentes, A. (Ed.), The international encyclopedia of primatology (pp. 462464). Oxford: Wiley-Blackwell.Google Scholar
Veilleux, C. C., & Cummings, M. E. (2012). Nocturnal light environments and species ecology: Implications for nocturnal color vision in forests. Journal of Experimental Biology, 215(23), 40854096.Google Scholar
Veilleux, C. C., & Kirk, E. C. (2009). Visual acuity in the cathemeral strepsirrhine Eulemur macaco flavifrons. American Journal of Primatology, 71(4), 343352.Google Scholar
Veilleux, C. C., & Kirk, E. C. (2014). Visual acuity in mammals: Effects of eye size and ecology. Brain, Behavior and Evolution, 83(1), 4353.Google Scholar
Veilleux, C. C., Louis, E. E. Jr, & Bolnick, D. A. (2013). Nocturnal light environments influence color vision and signatures of selection on the OPN1SW opsin gene in nocturnal lemurs. Molecular Biology and Evolution, 30(6), 14201437.Google Scholar
Veilleux, C. C., Jacobs, R. L., Cummings, M. E., Louis, E. E., & Bolnick, D. A. (2014). Opsin genes and visual ecology in a nocturnal folivorous lemur. International Journal of Primatology, 35(1), 88107.Google Scholar
Verbeken, D., Dierckx, S., & Dewettinck, K. (2003). Exudate gums: Occurrence, production, and applications. Applied Microbiology and Biotechnology, 63(1), 1021.Google Scholar
Vereecke, E. E., & Wunderlich, R. E. (2016). Experimental research on hand use and function in primates. In Kivell, T., Lemelin, P., Richmond, B., & Schmitt, D. (Eds.), The evolution of the primate hand: Anatomical, developmental, functional, and paleontological evidence (pp. 259284). New York, NY: Springer.Google Scholar
Veríssimo, D., & Wan, A. K. (2018). Characterizing efforts to reduce consumer demand for wildlife products. Conservation Biology. DOI: http://doi.org/10.1111/cobi.13227.Google Scholar
Veríssimo, D., Vaughan, G., Ridout, M., Waterman, C., MacMillan, D., & Smith, R. J. (2017). Increased conservation marketing effort has major fundraising benefits for even the least popular species. Biological Conservation, 211, 95101.Google Scholar
Vijay, V., Pimm, S. L., Jenkins, C. N., & Smith, S. J. (2016). The impacts of oil palm on recent deforestation and biodiversity loss. PLoS One, 11(7), e0159668.Google Scholar
Vilensky, J. A., & Larson, S. G. (1989). Primate locomotion: Utilization and control of symmetrical gaits. Annual Review of Anthropology, 18(1), 1735.Google Scholar
Vinyard, C. J. (2007). Interspecific analysis of covariance structure in the masticatory apparatus of galagos. American Journal of Primatology, 69(1), 4658.Google Scholar
Vinyard, C. J., Wall, C. E., Williams, S. H., & Hylander, W. L. (2003). Comparative functional analysis of skull morphology of tree‐gouging primates. American Journal of Physical Anthropology, 120(2), 153170.Google Scholar
Vinyard, C. J., Ravosa, M. J., Williams, S. H., Wall, C. E., Johnson, K. R., & Hylander, W. L. (2007). Jaw-muscle function and the origin of primates. In Ravosa, M. J. & Dagosto, M. (Eds.), Primate origins: Adaptations and evolution (pp. 179231). Boston, MA: Springer.Google Scholar
Visualisation Sciences Group (2009). Avizo. Burlington, VT: Mercury Computer Systems.Google Scholar
Von Haller, A. (1756). Icones anatomicae. Gottingen: A. Vandenhoeck.Google Scholar
Voskamp, A., Rode, E. J., Coudrat, C. N., Wilson, R. J., & Nekaris, K. A. I. (2014). Modelling the habitat use and distribution of the threatened Javan slow loris Nycticebus javanicus. Endangered Species Research, 23(3), 277286.Google Scholar
Vosmaer, A. (1770). Description d’une espèce de paresseux pentadactyle, jusqu’ici inconnu, qui se trouve au Bengale. Amsterdam: Pierre Maijer.Google Scholar
Vuarin, P., & Henry, P. Y. (2014). Field evidence for a proximate role of food shortage in the regulation of hibernation and daily torpor: A review. Journal of Comparative Physiology B, 184(6), 683697.Google Scholar
Vuarin, P., Dammhahn, M., & Henry, P. Y. (2013). Individual flexibility in energy saving: Body size and condition constrain torpor use. Functional Ecology, 27(3), 793799.Google Scholar
Walker, A. (1969a). The locomotion of the lorises, with special reference to the potto. African Journal of Ecology, 7(1), 15.Google Scholar
Walker, A. (1969b). True affinities of Propotto leakeyi Simpson 1967. Nature, 223(5206), 647.Google Scholar
Walker, A. (1972). The dissemination and segregation of early primates in relation to continental configuration. In Bishop, W. W. & Miller, J. A. (Eds.), Calibration of hominid evolution, (pp. 192218). Edinburgh: Scottish Academic Press.Google Scholar
Walker, A. (1974). A review of the Miocene Lorisidae of East Africa. In Martin, R. D., Doyle, G. A., & Walker, A. C. (Eds.), Prosimian biology (pp. 435447). London: Duckworth.Google Scholar
Walker, A. (1978). Prosimian primates. In Maglio, V. J. & Cooke, H. B. S. (Eds.), Evolution of African mammals (pp. 9099). Cambridge, MA: Harvard University Press.Google Scholar
Walker, A. (1979). Prosimian locomotor behavior. In Doyle, G. A. & Martin, R. D. (Eds.), The study of prosimian behavior (pp. 543565). New York, NY: Academic Press.Google Scholar
Walker, A., Ryan, T. M., Silcox, M. T., Simons, E. L., & Spoor, F. (2008). The semicircular canal system and locomotion: The case of extinct lemuroids and lorisoids. Evolutionary Anthropology: Issues, News, and Reviews, 17(3), 135145.Google Scholar
Walls, G. L. (1942). The vertebrate eye and its adaptive radiation. Bloomfield Hills, MI: Cranbrook Press.Google Scholar
Walston, J., Davidson, P., & Soriyun, M. (2001). A wildlife survey of southern Mondulkiri Province, Cambodia. Phnom Penh: Wildlife Conservation Society Cambodia Programme.Google Scholar
Wang, W., Su, B., Lan, H., Liu, R., Zhu, C., Nie, W., …, & Zhang, Y. (1996). Interspecific differentiation of the slow lorises (genus Nycticebus) inferred from ribosomal DNA restriction maps. Zoological Research, 17(1), 8993.Google Scholar
Wang, Z., Liu, L., Li, Q., & Sun, R. (2000). The characteristics of nonshivering thermogenesis and cellular respiration in lesser slow loris (Nycticebus pygmaeus). Acta Theriologica Sinica, 20(1), 1320.Google Scholar
Wang, Z. K., & Sun, R. Y. (1995). Characteristics of the resting metabolism and temperature regulation in lesser slow loris (Nycticebus pygmaeus). Acta Zoologica Sinica, 41(2), 149157.Google Scholar
Ward, S., Brown, B., Hill, A., Kelley, J., & Downs, W. (1999). Equatorius: A new hominoid genus from the middle Miocene of Kenya. Science, 285(5432), 13821386.Google Scholar
Warrant, E. (2004). Vision in the dimmest habitats on earth. Journal of Comparative Physiology A, 190(10), 765789.Google Scholar
Warren, R. D., & Crompton, R. H. (1998). Diet, body size and the energy costs of locomotion in saltatory primates. Folia Primatologica, 69(suppl. 1), 86100.Google Scholar
Warter, S., & Rumpler, Y. (1998). Chromosomal variability in the genus Nycticebus. Folia Primatologica, 69(5), 296299.Google Scholar
Watanuki, Y., & Nakayama, Y. (1993). Age difference in activity pattern of Japanese monkeys: Effects of temperature, snow, and diet. Primates, 34(4), 419430.Google Scholar
Watson, S. L., Ward, J. P., Davis, K. B., & Stavisky, R. C. (1999). Scent-marking and cortisol response in the small-eared bushbaby (Otolemur garnettii). Physiology & Behavior, 66(4), 695699.Google Scholar
Weber, D., Hintermann, U., & Zangger, A. (2004). Scale and trends in species richness: Considerations for monitoring biological diversity for political purposes. Global Ecology and Biogeography, 13(2), 97104.Google Scholar
Webster, S. (2015). The trade in slow lorises (Nycticebus spp.): Optimising social media reporting policies and its use as a conservation and awareness tool (unpublished Master’s dissertation). Oxford Brookes University, Oxford, UK.Google Scholar
Weisenseel, K., Chapman, C. A., & Chapman, L. J. (1993). Nocturnal primates of Kibale Forest: Effects of selective logging on prosimian densities. Primates, 34(4), 445450.Google Scholar
Weisenseel, K., Izard, M. K., Nash, L. T., Ange, R. L., & Poorman-Allen, P. (1998). A comparison of reproduction in two species of NycticebusFolia Primatologica69(suppl. 1), 321324.Google Scholar
Weller, S. (2007). Cultural consensus theory: Applications and frequently asked questions. Field Methods, 19(4), 339368.Google Scholar
Weller, S., & Romney, A. (1988). Systematic data collection. Thousand Oaks, CA: Sage.Google Scholar
Wells, K., Pfeiffer, M., Lakim, M. B., & Linsenmair, K. E. (2004). Use of arboreal and terrestrial space by a small mammal community in a tropical rain forest in Borneo, Malaysia. Journal of Biogeography, 31(4), 641652.Google Scholar
Wells, N. M., & Lekies, K. S. (2006). Nature and the life course: Pathways from childhood nature experiences to adult environmentalism. Children Youth and Environments, 16(1), 124.Google Scholar
Westphal, M. I., Browne, M., MacKinnon, K., & Noble, I. (2008). The link between international trade and the global distribution of invasive alien species. Biological Invasions, 10(4), 391398.Google Scholar
White, G. C., & Garrott, R. A. (1990). Analysis of wildlife radio tracking data. San Diego, CA: Academic Press.Google Scholar
Whittow, G. C., & Liat, L. B. (1977). Body temperature and oxygen consumption of two Malaysian prosimians. Primates, 18(2), 471474.Google Scholar
Whittow, G. G., Scammell, C. C., Manuel, J. K., Rand, D., & Leong, M. (1977). Temperature regulation in a hypometabolic primate, the slow loris (Nycticebus coucang). Archives Internationales De Physiologie Et De Biochimie, 85(1), 139151.Google Scholar
Whytock, R. C., Buij, R., Virani, M. Z., & Morgan, B. J. (2016). Do large birds experience previously undetected levels of hunting pressure in the forests of Central and West Africa?. Oryx, 50(1), 7683.Google Scholar
Wible, J. R. (2011). On the treeshrew skull (Mammalia, Placentalia, Scandentia). Annals of Carnegie Museum, 79(3), 149231.Google Scholar
Wiens, F. (2002). Behavior and ecology of wild slow lorises (Nycticebus coucang): Social organization, infant care system, and diet (unpublished doctoral dissertation). University of Bayreuth, Bayreuth, Germany.Google Scholar
Wiens, F., & Zitzmann, A. (2003a). Social dependence of infant slow lorises to learn diet. International Journal of Primatology, 24(5), 10071021.Google Scholar
Wiens, F., & Zitzmann, A. (2003b). Social structure of the solitary slow loris Nycticebus coucang (Lorisidae). Journal of Zoology, 261(1), 3546.Google Scholar
Wiens, F., Zitzmann, A., & Hussein, N. A. (2006). Fast food for slow lorises: Is low metabolism related to secondary compounds in high-energy plant diet?. Journal of Mammalogy, 87(4), 790798.Google Scholar
Wikler, K. C., & Rakic, P. (1990). Distribution of photoreceptor subtypes in the retina of diurnal and nocturnal primates. Journal of Neuroscience, 10(10), 33903401.Google Scholar
Wilde, H. (1972). Anaphylactic shock following bite by a ‘slow loris,’ Nycticebus coucang. American Journal of Tropical Medicine and Hygiene, 21(5), 592594.Google Scholar
Willcox, A. S., & Nambu, D. M. (2007). Wildlife hunting practices and bushmeat dynamics of the Banyangi and Mbo people of southwestern Cameroon. Biological Conservation, 134, 251261.Google Scholar
Williams, B. A., Kay, R. F., Kirk, E. C., & Ross, C. F. (2010). Darwinius masillae is a strepsirrhine: A reply to Franzen et al.(2009). Journal of Human Evolution, 59(5), 567573.Google Scholar
Williams, B. K., Nichols, J. D., & Conroy, M. J. (2002). Analysis and management of animal populations. New York, NY: Academic Press.Google Scholar
Williams, E., Cabana, F., & Nekaris, K. A. I. (2015). Improving diet and activity of insectivorous primates in captivity: Naturalizing the diet of Northern Ceylon gray slender loris, Loris lydekkerianus nordicus. Zoo Biology, 34(5), 473482.Google Scholar
Williams, S. H., Wall, C. E., Vinyard, C. J., & Hylander, W. L. (2002). A biomechanical analysis of skull form in gum-harvesting galagids. Folia Primatologica, 73(4), 197209.Google Scholar
Wilson, D. E., & Mittermeier, R. A. (2015). Handbook of the Mammals of the World, vol. 5. Barcelona: Lynx Edicions.Google Scholar
Wilson, E. O. (1984). Biophilia: The human bond with other species. London: Harvard University Press.Google Scholar
Woerndl, M., Papagiannidis, S., Bourlakis, M., & Li, F. (2008). Internet-induced marketing techniques: Critical factors in viral marketing campaigns. International Journal of Business Science and Applied Management, 3(1), 3345.Google Scholar
Wolf, I. D., & Croft, D. B. (2012). Observation techniques that minimize impacts on wildlife and maximize visitor satisfaction in night-time tours. Tourism Management Perspectives, 4, 164175.Google Scholar
Wolin, L. R., & Massopust, L. C. (1970). Morphology of the primate retina. In Noback, C. R. & Montagna, W. (Eds.), The primate brain (pp. 127). New York, NY: Appleton Century Crofts.Google Scholar
Woodroffe, R., Thirgood, S., & Rabinowitz, A. (2005). The impact of human–wildlife conflict on natural systems. In Woodroffe, R., Thirgood, S., & Rabinowitz, A. (Eds.), People and wildlife: Conflict and coexistence (pp. 112). New York, NY: Cambridge University Press.Google Scholar
World Animal Protection. (2016). Checking out of cruelty: How to end wildlife tourism’s holiday horrors. Retrieved 10 September 2016, from www.worldanimalprotection.ca/news/holiday-horror-show-animals-suffering-tourists-entertainment.Google Scholar
Wortman, J. L. (1903). Studies of Eocene Mammalia in the Marsh Collection: Peabody Museum. American Journal of Science (1880–1910), 16(95), 345.Google Scholar
Worton, B. J. (1989). Kernel methods for estimating the utilization distribution in home‐range studies. Ecology, 70(1), 164168.Google Scholar
Worton, B. J. (1995). Using Monte Carlo simulation to evaluate kernel-based home range estimators. Journal of Wildlife Management, 59, 794800.Google Scholar
Wright, S. (1932). The roles of mutation, inbreeding, cross-breeding and selection in evolution. Proceedings of the 6th Int Congress on Genetics, 1, 356366.Google Scholar
Wright, S. (1982). The shifting balance theory and macroevolution. Annual Reviews in Genetics, 16: 119.Google Scholar
Wu, J. Y. (1990). Morphological observation of dentition of slow lorises. Chinese Journal of Zoology, 25(1), 2527.Google Scholar
Wu, J. Y., & Wang, Z. K. (1991). Morphological characters of skull and comparative study of slow loris. Chinese Journal of Zoology, 26(4), 2933.Google Scholar
Wu, J., Jiao, H., Simmons, N. B., Lu, Q., & Zhao, H. (2018). Testing the sensory trade-off hypothesis in New World bats. Proceedings of the Royal Society B: Biological Sciences, 285(1885), 20181523.Google Scholar
Wu, M. C. (1983). Taxonomy of primates in Guangxi and its population estimation. Acta Theriologica Sin, 3(1), 16.Google Scholar
Wunderlich, R. E., Tongen, A., Gardiner, J., Miller, C. E., & Schmitt, D. (2014). Dynamics of locomotor transitions from arboreal to terrestrial substrates in Verreaux’s sifaka (Propithecus verreauxi). Integrative and Comparative Biology, 54(6), 11481158.Google Scholar
Wyler, L. S., & Sheik, P. A. (2013). International Illegal Trade in Wildlife: Threats and U.S. Policy. Washington, DC: Congressional Research Service.Google Scholar
Xiao, C. H., Wang, R., Wang, Z. K., Liu, P. F., Chu, Y. X., Qian, L. C., …, & Zhu, W. L. (2009). Energy metabolism of Pygmy loris (Nycticebus pygmaeus) in cage of Kunming Zoo. Acta Theriologica Sinica, 29(4), 443446.Google Scholar
Xiao, C., Wang, Z., Zhu, W., Chu, Y., Liu, C., Jia, T., …, & Cai, J. (2010). Energy metabolism and thermoregulation in pygmy lorises (Nycticebus pygmaeus) from Yunnan Daweishan Nature Reserve. Acta Ecologica Sinica, 30(3), 129134.Google Scholar
Xu, B., Huang, Z. X., Wang, X. Y., Gao, R. C., Tang, X. H., Mu, Y. L., …, & Zhu, L. D. (2010). Phylogenetic analysis of the fecal flora of the wild pygmy loris. American Journal of Primatology, 72(8), 699706.Google Scholar
Xu, B., Xu, W., Yang, F., Li, J., Yang, Y., Tang, X., …, & Huang, Z. (2013). Metagenomic analysis of the pygmy loris fecal microbiome reveals unique functional capacity related to metabolism of aromatic compounds. PLoS One, 8(2), e56565.Google Scholar
Yamashita, N. (2017). Toothcomb. In Fuentes, A. (Ed.), The international encyclopedia of primatology (pp. 13961397). Oxford: Wiley-Blackwell.Google Scholar
Yang, Z., & Yoder, A. D. (2003). Comparison of likelihood and Bayesian methods for estimating divergence times using multiple gene loci and calibration points, with application to a radiation of cute-looking mouse lemur species. Systematic Biology, 52(5), 705716.Google Scholar
Yin, A., Dubey, C. S., Kelty, T. K., Webb, A. A. G., Harrison, T. M., Chou, C. Y., & Célérier, J. (2010). Geologic correlation of the Himalayan orogen and Indian craton: Part 2. Structural geology, geochronology, and tectonic evolution of the Eastern Himalaya. Geological Survey of America Bulletin, 122(3–4), 360395.Google Scholar
Yoder, A. D. (1994). Relative position of the Cheirogaleidae in strepsirhine phylogeny: A comparison of morphological and molecular methods and results. American Journal of Physical Anthropology, 94(1), 2546.Google Scholar
Yoder, A. D. (1997). Back to the future: A synthesis of strepsirrhine systematics. Evolutionary Anthropology: Issues, News, and Reviews, 6(1), 1122.Google Scholar
Yoder, A. D., & Yang, Z. (2004). Divergence dates for Malagasy lemurs estimated from multiple gene loci: Geological and evolutionary context. Molecular Ecology, 13(4), 757773.Google Scholar
Yoder, A. D., Rasoloarison, R. M., Goodman, S. M., Irwin, J. A., Atsalis, S., Ravosa, M. J., & Ganzhorn, J. U. (2000). Remarkable species diversity in Malagasy mouse lemurs (Primates, Microcebus). Proceedings of the National Academy of Sciences, 97(21), 1132511330.Google Scholar
Yoder, A. D., Irwin, J. A., & Payseur, B. A. (2001). Failure of the ILD to determine data combinability for slow loris phylogeny. Systematic Biology, 50, 408424.Google Scholar
Yoder, A. D., Chan, L. M., dos Reis, M., Larsen, P. A., Campbell, C. R., Rasoloarison, R., …, & Yang, Z. (2014). Molecular evolutionary characterization of a V1R subfamily unique to strepsirrhine primates. Genome Biology and Evolution, 6(1), 213227.Google Scholar
Youlatos, D. (1993). Monkey (Alouatta seniculus). Folia Primatol, 61, 144147.Google Scholar
Youlatos, D. (2008). Locomotion and positional behavior of spider monkeys. In Campbell, C. (Eds.), Spider monkeys: Behavior, ecology, and evolution of the genus Ateles (pp. 185219). Cambridge: Cambridge University Press.Google Scholar
Young, J. M., Massa, H. F., Hsu, L., & Trask, B. J. (2010). Extreme variability among mammalian V1R gene families. Genome Research, 20(1), 1018.Google Scholar
Yu, F., Ye, Z., Pan, R., Peng, Y., & Wang, H. (1993). The multivariate analysis of the shoulder joint in slow loris. Zoological Research, 14(2), 102109.Google Scholar
Yu, L. G., Chen, M. J., Yang, S. J., Li, X. Y., & Shi, L. (2013). Camera trapping survey of Nyticebus pygmaeus, Nycticebus coucang and other sympatric mammals at Dawei Mountain, Yunnan. Sichuan Journal of Zoology, 32(6), 814818.Google Scholar
Yu, X., & Jia, W. (2015). Moving targets: Tracking online sales of illegal wildlife products in China. Cambridge: TRAFFIC.Google Scholar
Zhang, Y. P., Chen, Z. P., & Shi, L. M. (1993). Phylogeny of the slow lorises (genus Nycticebus): An approach using mitochondrial-DNA restriction enzyme analysis. International Journal of Primatology, 14, 167175.Google Scholar
Zhang, R. Z., Chen, L. W., Qu, W. Y., & Ke, L. S. (2002). Biogeography and natural conservation of primates in China: The past, present and future. Beijing: China Forestry Press.Google Scholar
Zhang, Y. Z., Quan, G. Q., Zhao, T. G., & Southwick, C. H. (1992). Distribution of primates (except Macaca) in China. Acta Theriologica Sinica, 12(2), 8195.Google Scholar
Zhao, H., Rossiter, S. J., Teeling, E. C., Li, C., Cotton, J. A., & Zhang, S. (2009). The evolution of color vision in nocturnal mammals. Proceedings of the National Academy of Sciences, 106(22), 89808985.Google Scholar
Zhdanova, I. V., Geiger, D. A., Schwagerl, A. L., Leclair, O. U., Killiany, R., Taylor, J. A., …, & Madras, B. K. (2002). Melatonin promotes sleep in three species of diurnal nonhuman primates. Physiology and Behavior,75, 523529.Google Scholar
Zhizhang, Y., Peng, F., Ye, Y., & Ruliang, P. (1992). Studies on the biomechanics of elbow joint in four species of primates. Acta Theriologica Sinica, 12(2), 94104.Google Scholar
Zhou, F., & Pan, G. P. (1995). Primates in the karst mountains of Guangxi. Journal of the Guangxi Academy of Sciences, 11(1), 811.Google Scholar
Zimmermann, E. (1989). Reproduction, physical growth and behavioral development in slow loris (Nycticebus coucang, Lorisidae). Human Evolution, 4(2–3), 171179.Google Scholar
Zimmermann, E. (1990). Differentiation of vocalizations in bushbabies (Galaginae, Prosimiae, Primates) and the significance for assessing phylogenetic relationships. Journal of Zoological Systematics and Evolutionary Research, 28(3), 217239.Google Scholar
Zootierliste. (2018). Zootierliste. Retrieved 2 August 2018, from www.zootierliste.de/en/?org=7.Google Scholar
ZSL. (2018). Animal records 2018. Retrieved 14 July 2018, from www.zsl.org/education/animal-records.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.

  • References
  • Edited by K. A. I. Nekaris, Oxford Brookes University, Anne M. Burrows, Duquesne University, Pittsburgh
  • Book: Evolution, Ecology and Conservation of Lorises and Pottos
  • Online publication: 29 February 2020
  • Chapter DOI: https://doi.org/10.1017/9781108676526.038
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.

  • References
  • Edited by K. A. I. Nekaris, Oxford Brookes University, Anne M. Burrows, Duquesne University, Pittsburgh
  • Book: Evolution, Ecology and Conservation of Lorises and Pottos
  • Online publication: 29 February 2020
  • Chapter DOI: https://doi.org/10.1017/9781108676526.038
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.

  • References
  • Edited by K. A. I. Nekaris, Oxford Brookes University, Anne M. Burrows, Duquesne University, Pittsburgh
  • Book: Evolution, Ecology and Conservation of Lorises and Pottos
  • Online publication: 29 February 2020
  • Chapter DOI: https://doi.org/10.1017/9781108676526.038
Available formats
×