Skip to main content Accessibility help
×
Hostname: page-component-5c6d5d7d68-wbk2r Total loading time: 0 Render date: 2024-08-14T15:35:48.289Z Has data issue: false hasContentIssue false

References

Published online by Cambridge University Press:  04 September 2009

Kathreen Ruckstuhl
Affiliation:
University of Cambridge
Peter Neuhaus
Affiliation:
University of Cambridge
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: 2006

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

Abernethy, K. A., White, L. J. T. and Wickings, E. J. (2002). Hordes of mandrills (Mandrillus sphinx): extreme group size and seasonal male presence. Journal of Zoology, London, 258, 131–7.CrossRefGoogle Scholar
Acuna, H. O. and Francis, J. M. (1995). Spring and summer prey of the Juan-Fernandez fur seal, Arctocephalus philippii. Canadian Journal of Zoology, 73, 1444–52.CrossRefGoogle Scholar
Adamopoulou, C. and Legakis, A. (2002). Diet of a lacertid lizard (Podarcis milensis) in an insular dune ecosystem. Israel Journal of Zoology, 48, 207–19.CrossRefGoogle Scholar
Afanasyev, V. (2004). A miniature daylight level and activity data recorder for tracking animals over long periods. Memoirs of the National Institute of Polar Research, Special Issue, 58, 227–233.Google Scholar
Agrimi, U. and Luiselli, L. (1992). Feeding strategies of the viper V. u. ursinii (Reptilia: Viperidae) in the Apennines. Herpetological Journal, 2, 37–42.Google Scholar
Alberts, S. A. and Altmann, J. (1995). Balancing costs and benefits: dispersal in male baboons. American Naturalist, 145, 279–306.CrossRefGoogle Scholar
Albon, S. D., Clutton-Brock, T. H. and Langvatn, R. (1992). Cohort variation in reproduction and survival: implications for population demography. In The Biology of Deer, ed. Brown, R. D.. Berlin: Springer Verlag.CrossRefGoogle Scholar
Alexander, C., Lynch, J. J. and Mottershead, B. E. (1979). Use of shelter and selection of lambing sites by shorn and unshorn ewes in paddocks with closely or widely spaced shelters. Applied Animal Ethology, 5, 51–69.CrossRefGoogle Scholar
Allen, B. A., Burghardt, G. M. and York, D. S. (1984). Species and sex differences in substrate preference and tongue-flick rate in three sympatric species of water snakes (Nerodia). Journal of Comparative Psychology, 98, 358–67.CrossRefGoogle Scholar
Altmann, J. (1980). Baboon Mothers and Infants. Cambridge, MA: Harvard University Press.Google Scholar
Altmann, J. (1990). Primate males go where the females are. Animal Behaviour, 39, 193–5.CrossRefGoogle Scholar
Altringham, J. D. (1996). Bats: Biology and Behaviour. Oxford: Oxford University Press.Google Scholar
Altringham, J. D. and Fenton, M. B. (2003). Sensory ecology and communication in the Chiroptera. In Bat Ecology, eds. Kunz, T. H. and Fenton, M. B.. Chicago: University of Chicago Press, pp. 90–127.Google Scholar
Anderson, D. J., Reeve, J., Gomez, J. E. M.et al. (1993). Sexual size dimorphism and food requirements of nestling birds. Canadian Journal of Zoology, 71, 2541–5.CrossRefGoogle Scholar
Anderson, D. J., Schwandt, A. J. and Douglas, H. D. (1998). Foraging ranges of waved albatrosses in the Eastern tropical Pacific. In Albatross Biology and Conservation, eds. Robertson, G. and Gales, R.. Chipping Norton: Surrey Beatty and Sons, pp. 180–5.Google Scholar
Andersson, M. (1994). Sexual Selection, Princeton: Princeton University Press.Google Scholar
Andrews, R. M. (1971). Structural habitat and time budget of a tropical Anolis lizard. Ecology, 52, 262–70.CrossRefGoogle Scholar
Andrews, R. M., Cruz, F. R. M. and Cruz, Santa M. V. (1997). Body temperatures of female Sceloporus grammicus: Thermal stress or impaired mobility? Copeia, 1997, 108–15.CrossRefGoogle Scholar
Angelici, F. M., Effah, C., Inyang, M. A. and Luiselli, L. (2000). A preliminary radiotracking study of movements, activity patterns and habitat use of free-ranging Gaboon vipers, Bitis gabonica. Terre et la Vie, 55, 45–55.Google Scholar
Anibaldi, C., Luiselli, L. and Angelici, F. M. (1998). Notes on the ecology of a suburban population of rainbow lizards in coastal Kenya. African Journal of Ecology, 36, 199–206.CrossRefGoogle Scholar
Antonelis, G. A., Lowry, M. S., Fiscus, C. H., Stewart, B. S. and Delong, R. L. (1994). Diet of the northern elephant seal. In Elephant Seals: Population Ecology, Behaviour, and Physiology, ed. Boeuf, B. J. and Laws, R. M.. Berkley: University of California Press, pp. 211–23.Google Scholar
Appleby, M. C. (1982). The consequences and causes of high social rank in red deer stags. Behaviour, 80, 259–73.CrossRefGoogle Scholar
Appleby, M. C. (1983). Competition in a red deer stag social group – rank, age and relatedness of opponents. Animal Behaviour, 31, 913–18.CrossRefGoogle Scholar
Arcese, P., Keller, L. F. and Cary, J. R. (1997). Why hire a behaviorist into a conservation or management team? In Behavioral Approaches to Conservation in the Wild, eds. Clemmons, J. R. and Buchholz, R.. Cambridge, United Kingdom: Cambridge University Press, pp. 48–71.Google Scholar
Archer, J. and Lloyd, B. (2002). Sex and Gender. Cambridge, UK: Cambridge University Press.CrossRefGoogle Scholar
Archibald, G. W. and Meine, C. D. (1996). Family Gruidae (Cranes). In Handbook of the Birds of the World, Vol. 3., eds. Hoyo, J. del, Elliot, A. and Sargatal, J.. Barcelona: Lynx Edicions.Google Scholar
Ardia, D. R. and Bildstein, K. L. (1997). Sex-related differences in habitat selection in wintering American kestrels, Falco sparverius. Animal Behaviour, 53, 1305–11.CrossRefGoogle ScholarPubMed
Ardia, D. R. and Bildstein, K. L. (2001). Sex-related differences in habitat use in wintering American kestrels. Auk, 118, 746–50.CrossRefGoogle Scholar
Arkhipkin, A. I. and Middletion, D. A. J. (2002). Sexual segregation in ontogenetic migrations by the squid Loligo gahi around the Falkland Islands. Bulletin of Marine Science, 71, 109–27.Google Scholar
Arlettaz, R. (1999). Habitat selection as a major resource partitioning mechanism between the two sympatric sibling bat species Myotis myotis and Myotis blythii. Journal of Animal Ecology, 68, 460–71.CrossRefGoogle Scholar
Arnbom, T. and Lundberg, S. (1995). Notes on Lepas australis (Cirripedia, Lepadidae) recorded on the skin of southern elephant seal (Mirounga leonina). Crustaceana, 68, 655–8.CrossRefGoogle Scholar
Arnbom, T., Papastavrou, V., Weilgart, L. S. and Whitehead, H. (1987). Sperm whales react to an attack by killer whale. Journal of Mammalogy, 68, 450–3.CrossRefGoogle Scholar
Arnold, G. W., Boundy, C. A. P., Morgan, P. D. and Bartle, G. (1975). The roles of sight and hearing in the lamb in the location and discrimination between ewes. Applied Animal Ethology, 5, 43–50.CrossRefGoogle Scholar
Arnold, G. W., Steven, D. E. and Grassia, A. (1990). Associations between individuals and classes in groups of different size in a population of western grey kangaroos, Macropus fuliginosus. Australian Wildlife Research, 17, 551–62.CrossRefGoogle Scholar
Arnold, G. W., Steven, D. E. and Weeldenberg, J. R. (1994). Comparative ecology of western grey kangaroos (Macropus fuliginosus) and Euros (M. robustus erubescens) in Durokoppin Nature Reserve, isolated in the central wheatbelt of Western Australia. Wildlife Research, 21, 307–22.CrossRefGoogle Scholar
Arnold, P. W. (1972). Predation on harbour porpoise, Phocoena phocoena, by a white shark, Carcharodon carcharias. Journal of Fisheries Research Board of Canada, 29, 1213–14.CrossRefGoogle Scholar
Arnold, S. J. (1993). Foraging theory and prey-size-predator-size relations in snakes. In Snakes: Ecology and Behavior, eds. Seigel, R. A. and Collins, J. T.. New York: McGraw-Hill, pp. 87–116.Google Scholar
Arnold, S. J. and Peterson, C. R. (2002). A model for optimal reaction norms: the case of the pregnant garter snake and her temperature-sensitive embryos. American Naturalist, 160, 306–16.Google ScholarPubMed
Arnould, J. P. Y. and Duck, C. D. (1997). The cost and benefits of territorial tenure, and factors affecting mating success in male Antarctic fur seals. Journal of Zoology, 241, 649–64.CrossRefGoogle Scholar
Arnould, J. P. Y., Green, J. A. and Rawlins, D. R. (2001). Fasting metabolism in Antarctic fur seal (Arctocephalus gazella) pups. Comparative Biochemistry and Physiology. Part A, Molecular and Integrative Physiology, 129, 829–41.CrossRefGoogle ScholarPubMed
Atsalis, K. (1999). The diet of the brown mouse lemur, Microcebus rufus, in Ranomafana National Park, Madagascar. International Journal of Primatology, 20, 193–229.CrossRefGoogle Scholar
Backus, R. H., Springer, S. and Arnold, E. L. (1956). A contribution to the natural history of the white-tip shark, Pterolamiops longimanus (Poey). Deep-Sea Research, 3, 178–88.CrossRefGoogle Scholar
Badyaev, A. V. (2002). Growing apart: an ontogenetic perspective on the evolution of sexual size dimorphism. Trends in Ecology and Evolution, 17, 369–78.CrossRefGoogle Scholar
Baird, R. W. (2000). The killer whale: Foraging specializations and group hunting. In Cetacean Societies: Field Studies of Dolphins and Whales, eds. Mann, J., Connor, R. C., Tyack, P. L. and Whitehead, H.. Chicago: Academic Press, pp. 127–53.Google Scholar
Baird, R. W. and Whitehead, H. (2000). Social organization of mammal-eating killer whales: group stability and dispersal patterns. Canadian Journal of Zoology, 78, 2096–105.CrossRefGoogle Scholar
Baker, J. R. and McCann, T. S. (1989). Pathology and bacteriology of adult male Antarctic fur seals, Arctocephalus gazella, dying at Bird Island, South Georgia. British Veterinary Journal, 145, 263–75.CrossRefGoogle ScholarPubMed
Baldassarre, G. A. and Bolen, E. G. (1994). Waterfowl Ecology and Management. New York: John Wiley and Sons.Google Scholar
Baldellou, M. and Aden, A. (1997). Time, gender, and seasonality in vervet activity: a chronobiological approach. Primates, 38, 33–43.CrossRefGoogle Scholar
Barboza, P. S. and Bowyer, R. T. (2000). Sexual segregation in dimorphic deer: a new gastrocentric hypothesis. Journal of Mammalogy, 81, 473–89.2.0.CO;2>CrossRefGoogle Scholar
Barclay, R. M. R. (1991). Population structure of temperate zone insectivorous bats in relation to foraging behaviour and energy demand. Journal of Animal Ecology, 60, 165–78.CrossRefGoogle Scholar
Barclay, R. M. R. and Harder, L. D. (2003). Life histories of bats: life in the slow lane. In Bat Ecology, eds. Kunz, T. H. and Fenton, M. B.. Chicago: University of Chicago Press, pp. 209–53.Google Scholar
Barrette, C. (1991). The size of Axis deer fluid groups in Wilpattu National Park, Sri Lanka. Mammalia, 2, 207–20.Google Scholar
Barros, N. B. and Odell, D. K. (1990). Food habits of bottlenose dolphins in Southeastern United States. In The Bottlenose Dolphin, eds. Leatherwood, S. and Reeves, R. R.. San Diego: Academic Press, pp. 309–28.Google Scholar
Barry, R. E., Botje, M. A. and Grantham, L. B. (1984). Vertical stratification of Peromyscus leucopus and P. maniculatus in southwestern Virginia. Journal of Mammalogy, 65, 145–8.CrossRefGoogle Scholar
Bartle, J. A. (1990). Sexual segregation of foraging zones in procellariiform birds: implications of accidental capture on commercial fishery longlines of grey petrels (Procellaria cinerea). Notornis, 37, 146–50.Google Scholar
Bass, A. J., D'Aubrey, J. and Kistnasamy, N. (1973). Sharks of the east coast of southern Africa. I. The genus Carcharhinus (Carcharhinidae). Investigation Report of the Oceanographic Research Institute, 33, 1–168.Google Scholar
Bateson, P. P. G. and Martin, P. (1999). Design for a Life: How Behaviour Develops. London: Jonathan Cape.Google Scholar
Bauwens, D. and Thoen, C. (1981). Escape tactics and vulnerability to predation associated with reproduction in the lizard Lacerta vivipara. Journal of Animal Ecology, 50, 733–43.CrossRefGoogle Scholar
Beamish, F. W. H. (1978). Swimming capacity of fish. In Fish Physiology, Vol. 7, ed. Randall, D. J.. New York: Academic Press, pp. 101–87.Google Scholar
Bearder, S. K. (1987). Lorises, bushbabies, and tarsiers: diverse societies in solitary foragers. In Primate Societies, eds. Smuts, B. B., Cheney, D. L., Seyfarth, R. M., Wrangham, R. W., and Struhsaker, T. T.. Chicago: University of Chicago Press, pp. 11–24.Google Scholar
Beck, C. A., Bowen, W. D. and Iverson, S. J. (2003a). Sex differences in the seasonal patterns of energy storage and expenditure in a phocid seal. Journal of Animal Ecology, 72, 280–91.CrossRefGoogle Scholar
Beck, C. A., Bowen, W. D., McMillan, J. I. and Iverson, S. J. (2003b). Sex differences in the diving behaviour of a size-dimorphic capital breeder: the grey seal. Animal Behaviour, 66, 777–89.CrossRefGoogle Scholar
Becker, C. D. and Ginsberg, J. R. (1990). Mother-infant behavior of wild Grevy's zebra: adaptation for survival in semi-desert East Africa. Animal Behaviour, 40, 1111–18.CrossRefGoogle Scholar
Becker, D. S. (1988). Relationship between sediment character and sex segregation in English sole, Parophrys vetulus. U.S. Fishery Bulletin, 86, 517–24.Google Scholar
Becker, P. H., González-Solís, J., Behrends, B. and Croxall, J. P. (2002). Feather mercury levels in seabirds at South Georgia: Influence of trophic position, sex and age. Marine Ecology Progress Series, 243, 261–9.CrossRefGoogle Scholar
Beer, J. R. and Richards, A. G. (1956). Hibernation of the big brown bat. Journal of Mammalogy, 37, 31–41.CrossRefGoogle Scholar
Beier, P. (1987). Sex differences in quality of white-tailed deer diets. Journal of Mammalogy, 68, 323–9.CrossRefGoogle Scholar
Beier, P. and McCullough, D. R. (1990). Factors influencing white-tailed deer activity patterns and habitat use. Wildlife Monographs, 109, 1–51.Google Scholar
Beissinger, S. R. (1997). Integrating behavior into conservation biology: potential and limitations. In Behavioral Approaches to Conservation in the Wild, eds. Clemmons, J. R. and Buchholz, R.. Cambridge, United Kingdom: Cambridge University Press, pp. 23–47.Google Scholar
Béland, P., Faucher, A. and Corbeil, P. (1990). Observations on the birth of a beluga whale (Delphinapterus leucas) in the St. Lawrence Estuary, Quebec, Canada. Canadian Journal of Zoology, 68, 1327–9.CrossRefGoogle Scholar
Belcher, C. A. (2003). Demographics of tiger quoll (Dasyurus maculatus maculatus) populations in south-eastern Australia. Australian Journal of Zoology, 51, 611–26.CrossRefGoogle Scholar
Bell, R. H. V. (1969). The use of the herbaceous layer by grazing ungulates in the Serengeti National Park, Tanzania. Ph.D. thesis, University of Manchester.
Bell, R. H. V. (1971). A grazing ecosystem in the Serengeti. Scientific American, 225, 86–93.CrossRefGoogle Scholar
Benham, P. F. J. (1982). Synchronization of behaviour in grazing cattle. Applied Animal Ethology, 8, 403–4.CrossRefGoogle Scholar
Bennett, P. M. and Owens, P. F. (2002). Evolutionary Ecology of Birds. Oxford: Oxford University Press.Google Scholar
Benshemesh, J. and Johnson, K. (2003). Biology and conservation of marsupial moles (Notoryctes). In Predators with Pouches: The Biology of Carnivorous Marsupials, eds. Jones, M. E., Dickman, C. and Archer, M.. Melbourne: CSIRO Publishing, pp. 464–74.Google Scholar
Bercovitch, F. B. (1983). Time budgets and consorts in olive baboons (Papio anubis). Folia Primatologica, 41, 180–90.CrossRefGoogle Scholar
Bercovitch, F. B. (1997). Reproductive strategies of rhesus macaques. Primates, 38, 247–63.CrossRefGoogle Scholar
Bergan, J. F. and Smith, L. M. (1989). Differential habitat use by diving ducks wintering in South Carolina. Journal of Wildlife Management, 53, 1117–26.CrossRefGoogle Scholar
Berger, J. (1986). Wild Horses of the Great Basin. Chicago: University of Chicago Press.Google Scholar
Berger, J. (1991). Pregnancy incentives, predation constraints, and habitat shifts: experimental and field evidence for wild bighorn sheep. Animal Behaviour, 41, 66–77.CrossRefGoogle Scholar
Berger, J. and Cunningham, C. (1988). Size-related effects on search times in North American grassland female ungulates. Ecology, 69, 177–83.CrossRefGoogle Scholar
Berger, J. and Cunningham, C. (1995). Predation, sensitivity, and sex: why female black rhinoceroses outlive males. Behavioral Ecology, 6, 57–64.CrossRefGoogle Scholar
Berger, J. and Gompper, M. E. (1999). Sex ratios in extant ungulates: products of contemporary predation or past life histories? Journal of Mammalogy, 80, 1084–113.CrossRefGoogle Scholar
Berger, J., Swenson, J. E. and Persson, I.-L. (2001). Recolonizing carnivores and naïve prey: conservation lessons from Pleistocene extinctions. Science, 291, 1036–9.CrossRefGoogle ScholarPubMed
Bergerud, A. T. (1974). Rutting behaviour of Newfoundland caribou. In The Behaviour of Ungulates and its Relationship to Management, eds. V. Giest and F. Walther, pp. 395–435. International Union for Conservation of Nature and Natural Resources: New Series, 24, 1–940.
Bergerud, A. T., Butler, H. E. and Miller, D. R. (1984). Antipredator tactics of calving caribou: dispersion in mountains. Canadian Journal of Zoology, 62, 1566–75.CrossRefGoogle Scholar
Bernard, H. J. and Hohn, A. A. (1989). Differences in feeding habits between pregnant and lactating spotted dolphins (Stenella attenuata). Journal of Mammalogy, 70, 211–15.CrossRefGoogle Scholar
Bernstein, N. P. and Maxson, S. J. (1984). Sexually distinct activity patterns of blue-eyed shags in Antarctica. Condor, 86, 151–6.CrossRefGoogle Scholar
Berrow, S. D. and Croxall, J. P. (2001). Provisioning rate and attendance patterns of wandering albatrosses at Bird Island, South Georgia. Condor, 103, 230–9.CrossRefGoogle Scholar
Berrow, S. D., Humpidge, R. and Croxall, J. P. (2000). Influence of adult breeding experience on growth and provisioning of wandering albatross Diomedea exulans chicks at South Georgia. Ibis, 142, 199–207.CrossRefGoogle Scholar
Berry, J. F. and Shine, R. (1980). Sexual size dimorphism and sexual selection in turtles (Order Chelonia). Oecologia, 44, 185–91.CrossRefGoogle Scholar
Berta, A. (2002). Pinnipedia, overview. In Encyclopedia of Marine Mammals, eds. Perrin, W. F., Wursig, B. and Thewissen, J. G. M.. San Diego: Academic Press, pp. 931–9.Google Scholar
Berteaux, D. (1993). Female-biased mortality in a sexually dimorphic ungulate: feral cattle of Amsterdam Island. Journal of Mammalogy, 74, 732–7.CrossRefGoogle Scholar
Best, P. B. (1979). Social organization in sperm whales, Physeter macrocephalus. In Behavior of Marine Mammal: Current Perspectives in Research, vol. 3, Cetaceans, eds. Winn, H. E. and Olla, B. L.. New York: Plenum Press, pp. 227–89.Google Scholar
Best, R. C. and da Silva, V. M. F. (1993). Inia geoffrensis. Mammalian Species, 423, 1–8.Google Scholar
Best, T. L. and Gennaro, A. L. (1984). Feeding ecology of the lizard, Uta stansburiana, in southwestern Mexico. Journal of Herpetology, 18, 291–301.CrossRefGoogle Scholar
Best, T. L. and Pfaffenberger, G. S. (1987). Age and sexual variation in the diet of collared lizards (Crotaphytus collaris). Southwestern Naturalist, 32, 415–26.CrossRefGoogle Scholar
Beuchat, C. A. (1986). Reproductive influences on the thermoregulatory behavior of a live-bearing lizard. Copeia, 1986, 971–9.CrossRefGoogle Scholar
Bibby, C. J. and Thomas, D. K. (1984). Sexual dimorphism in size, moult and movements of Cetti's warbler Cettia cetti. Bird Study, 31, 28–34.CrossRefGoogle Scholar
Bigg, M. A. (1973). Census of California sea lions on southern Vancouver Island, British Columbia. Journal of Mammal Research, 54, 285–7.CrossRefGoogle Scholar
Bigg, M. A. (1990). Migration of northern for seals (Callorhinus ursinus) off western North America. Canadian Technical Report of Fisheries and Aquatic Science, 1764, 1–64.Google Scholar
Bigg, M. A., Olesiuk, P. F., Ellis, G. M., Ford, J. K. B. and Balcomb, K. C. III (1990). Social organization and genealogy of resident killer whales (Orcinus orca) in the coastal waters of British Columbia and Washington State. Report of the International Whaling Commission, Special issue, 12, 383–405.Google Scholar
BirdLife International (2000). Threatened Birds of the World. Barcelona and Cambridge, UK: Lynx edicions and BirdLife International.
Blackburn, D. G. (1985). Evolutionary origins of viviparity in the Reptilia. II. Serpentes, Amphisbaenia, and Icthyosauria. Amphibia-Reptilia, 6, 259–91.CrossRefGoogle Scholar
Blatchford, P., Baines, , and Pellegrini, A. D. (2003). The social context of school playground games: sex and ethnic differences and changes over time after entry into junior school. British Journal of Developmental Psychology, 21, 459–71.CrossRefGoogle Scholar
Blaxter, J. H. S. and Holliday, F. G. T. (1969). The behaviour and physiology of herring and other clupeids. Advances in Marine Biology, 1, 261–393.CrossRefGoogle Scholar
Bleich, V. C., Bowyer, R. T. and Wehausen, J. D. (1997). Sexual segregation in mountain sheep: resources or predation? Wildlife Monographs, 134, 1–50.Google Scholar
Bleich, V. C., Wehausen, J. D. and Holl, S. A. (1990). Desert-dwelling mountain sheep: conservation implications of a naturally fragmented distribution. Conservation Biology, 4, 383–90.CrossRefGoogle Scholar
Bleich, V. C., Wehausen, J. D., Ramey, R. R. and Rechel, J. L. (1996). Metapopulation theory and mountain sheep: implications for conservation. In Metapopulations and Wildlife Conservation, ed. McCullough, D. R.. Covelo, California: Island Press, pp. 353–73.Google Scholar
Bleske, A. L. and Buss, D. M. (2000). Can men and women just be friends? Personal Relationships, 7, 131–51.CrossRefGoogle Scholar
Bleske-Rechek, A. L. and Buss, D. M. (2001). Opposite-sex friendship: sex differences and similarities in initiation, selection, and interpretation. Personality and Social Psychology Bulletin, 27, 1310–23.CrossRefGoogle Scholar
Blondel, J., Perret, P., Anstett, M.-C. and Thébaud, C. (2002). Evolution of sexual size dimorphism in birds: test of hypotheses using blue tits in contrasted Mediterranean habitats. Journal of Evolutionary Biology, 15, 440–50.CrossRefGoogle Scholar
Blouin-Demers, G. and Weatherhead, P. J. (2001). Thermal ecology of black rat snakes (Elaphe obsoleta) in a thermally challenging environment. Ecology, 82, 3025–43.CrossRefGoogle Scholar
Boase, H. (1926). Proportion of male and female duck on Tay estuary. British Birds, 20, 169–72.Google Scholar
Boesch, C. and Boesch-Achermann, H. (2000). The Chimpanzees of the Tai Forest. Oxford: Oxford University Press.Google Scholar
Boinski, S. (1987). Birth synchrony in squirrel monkeys: a strategy to reduce neonatal predation. Behavioral Ecology and Sociobiology, 21, 393–400.CrossRefGoogle Scholar
Boinski, S. (1988). Sex differences in the foraging behavior of squirrel monkeys. Behavioral Ecology and Sociobiology, 23, 177–86.CrossRefGoogle Scholar
Boinski, S. (1994). Affiliation patterns among Costa Rican squirrel monkeys. Behaviour, 130, 191–209.CrossRefGoogle Scholar
Boinski, S. (1999). The social organizations of squirrel monkeys: implications for ecological models of social evolution. Evolutionary Anthropology, 8, 101–12.3.0.CO;2-O>CrossRefGoogle Scholar
Boinski, S., Sughrue, K., Selvaggi, L.et al. (2002). An expanded test of the ecological model of primate social evolution: competitive regimes and female bonding in three species of squirrel monkeys (Saimiri oerstedi, Saimiri boliviensis, and Saimiri sciureus). Behaviour, 139, 227–61.CrossRefGoogle Scholar
Bon, R. (1991). Social and spatial segregation of males and females in polygamous ungulates: proximate factors. In Ongulés/Ungulates 1991, eds. Spitz, F., Janeau, G., Gonzalez, G., and Aulagnier, S.. Paris/Toulouse: SFEPM-IRGM, pp. 195–8.Google Scholar
Bon, R. (1998). Perspective éthologique de l'étude de la ségrégation sexuelle chez les ongulés. Habilitation à Diriger les Recherches. Université Paul Sabatier, Toulouse.
Bon, R. and Campan, R. (1989). Social tendencies of the Corsican mouflon Ovis ammon musimon in the Caroux-Espinouse (South of France). Behavioural Processes, 19, 57–78.CrossRefGoogle Scholar
Bon, R. and Campan, R. (1996). Unexplained sexual segregation in polygamous ungulates: a defense of an ontogenetic approach. Behavioural Processes, 38, 131–54.CrossRefGoogle ScholarPubMed
Bon, R., Dubois, M. and Maublanc, M. L. (1993). Does age influence between-rams companionship in mouflon (Ovis gmelini)? Revue d'Ecologie (Terre vie), 47, 57–64.Google Scholar
Bon, R., Joachim, J. and Maublanc, M. L. (1995). Do lambs affect feeding habitat use of lactating female mouflons in spring within an area free of predators? Journal of Zoology, London, 235, 43–51.CrossRefGoogle Scholar
, Bon R., Rideau, C., Villaret, J.-C. and Joachim, J. (2001). Segregation is not only a matter of sex in Alpine ibex (Capra ibex ibex). Animal Behaviour, 62, 495–504.Google Scholar
Boness, D. J., Clapham, P. J. and Mesnick, S. L. (2002). Life history and reproductive strategies. In Marine Mammal Biology: An Evolutionary Approach, ed. Hoelzel, A. R.. Oxford: Blackwell Science, pp. 278–324.Google Scholar
Bonner, W. N. (1968). The fur seal of South Georgia. British Antarctic Survey Scientific Report, 56, 81.Google Scholar
Bonnet, X., Guy, N. and Shine, R. (1999a). The dangers of leaving home: dispersal and mortality in snakes. Biological Conservation, 89, 39–50.CrossRefGoogle Scholar
Bonnet, X., Naulleau, G., Shine, R. and Lourdais, O. (1999b). What is the appropriate timescale for measuring costs of reproduction in a ‘capital breeder’ such as the aspic viper? Evolutionary Ecology, 13, 485–97.CrossRefGoogle Scholar
Bonnet, X., Shine, R., Naulleau, G. and Thiburce, C. (2001). Plastic vipers: genetic and environmental influences on the size and shape of Gaboon vipers, Bitis gabonica. Journal of Zoology, London, 255, 341–51.CrossRefGoogle Scholar
Booth, J. and Peters, J. A. (1972). Behavioural studies on the green turtle (Chelonia mydas) in the sea. Animal Behaviour, 20, 808–12.CrossRefGoogle Scholar
Bourne, W. R. P. and Warham, J. (1966). Geographical variation in the giant petrels of the genus Macronectes. Ardea, 54, 45–67.Google Scholar
Bowen, W. D., Oftedal, O. T. and Boness, D. J. (1985). Birth to weaning in 4 days: remarkable growth in the hooded seal, Cystophora cristata. Canadian Journal of Zoology, 63, 2841–6.CrossRefGoogle Scholar
Bowen, W. D., Read, A. J. and Estes, J. A. (2002). Feeding ecology. In Marine Mammal Biology: An Evolutionary Approach, ed. Hoelzel, A. R.. Oxford: Blackwell Science, pp. 217–46.Google Scholar
Bowers, M. A. (1994). Dynamics of age- and habitat-structured populations. Oikos, 69, 327–33.CrossRefGoogle Scholar
Bowman, J., Jaeger, J. A. G. and Fahrig, L. (2002). Dispersal distance of mammals is proportional to home range size. Ecology, 83, 2049–55.CrossRefGoogle Scholar
Bowyer, R. T. (1984). Sexual segregation in southern mule deer. Journal of Mammalogy, 65, 410–17.CrossRefGoogle Scholar
, Bowyer R. T., Kie, J. G. and Ballengerghe, V. (1996). Sexual segregation in black-tailed deer: effects of scale. Journal of Wildlife Management, 60, 10–17.Google Scholar
Bowyer, R. T., Pierce, B. M., Duffy, L. K. and Haggstrom, D. A. (2001). Sexual segregation in moose: effects of habitat manipulation. Alces, 37, 109–22.Google Scholar
Boyd, I. L. (1993a). Pup production and distribution of breeding Antarctic fur seals Arctocephalus gazella at South Georgia. Antarctic Science, 5, 17–24.CrossRefGoogle Scholar
Boyd, I. L. (1993b). Trends in marine mammal science. In Marine Mammals: Advances in Behavioural and Population Biology, ed. Boyd, I. L.. Oxford: Clarendon Press, pp. 1–12.Google Scholar
Boyd, I. L. and Croxall, J. P. (1996). Dive durations in pinnipeds and seabirds. Canadian Journal of Zoology, 74, 1696–1705.CrossRefGoogle Scholar
Boyd, I. L., Arnbom, T. A. and Fedak, M. A. (1994). Biomass and energy consumption of the South Georgia population of southern elephant seals. In Elephant Seals: Population Ecology, Behaviour, and Physiology, eds. Boeuf, B. J. and Laws, R. M.. Berkley: University of California Press.Google Scholar
Boyd, I. L., Lockyer, C. and Marsh, H. D. (1996). Reproduction in marine mammals. In Biology of Marine Mammals, eds. Reynolds, J. E. and Rommel, S. A.. Washington: Smithsonian Institution Press.Google Scholar
Boyd, I. L., McCafferty, D. J., Reid, K., Taylor, R. and Walker, T. R. (1998). Dispersal of male and female Antarctic fur seals Arctocephalus gazella. Canadian Journal of Fisheries and Aquatic Sciences, 55, 845–52.CrossRefGoogle Scholar
Boyd, I. L., Staniland, I. J. and Martin, A. R. (2002). Distribution of foraging by female Antarctic fur seals. Marine Ecology, Progress Series, 242, 285–94.CrossRefGoogle Scholar
Bradbury, J. W. (1977). Lek mating behaviour in the hammer-headed bat. Zeitschrift für Tierpsychologie, 45, 225–55.CrossRefGoogle Scholar
Bradbury, J. W. and Vehrencamp, S. L. (1976a). Social organisation and foraging in emballonurid bats. I. Field studies. Behavioral Ecology and Sociobiology, 1, 337–81.CrossRefGoogle Scholar
Bradbury, J. W. and Vehrencamp, S. L. (1976b). Social organisation and foraging in emballonurid bats. II. A model for the determination of group size. Behavioral Ecology and Sociobiology, 1, 383–404.CrossRefGoogle Scholar
Bradbury, J. W. and Vehrencamp, S. L. (1977). Social organisation and foraging in emballonurid bats. III. Mating systems. Behavioral Ecology and Sociobiology, 2, 1–17.CrossRefGoogle Scholar
Bradshaw, S. D. and Bradshaw, F. J. (2002). Short-term movements and habitat use of the marsupial honey possum (Tarsipes rostratus). Journal of Zoology, 258, 343–8.CrossRefGoogle Scholar
Brana, F. (1993). Shifts in body temperature and escape behaviour of female Podarcis muralis during pregnancy. Oikos, 66, 216–22.CrossRefGoogle Scholar
Brana, F. (1996). Sexual dimorphism in lacertid lizards: male head increase vs. female abdomen increase? Oikos, 75, 511–23.CrossRefGoogle Scholar
Breitwisch, R. (1989). Mortality patterns, sex ratios and parental investment in monogamous birds. In Current Ornithology 6, ed. Power, D.. New York and London: Plenum Press.CrossRefGoogle Scholar
Bried, J. and Jouventin, P. (2002). Site and mate choice in seabirds: an evolutionary approach. In Biology of Marine Birds, eds. Schreiber, E. A. and Burger, J.. Boca Raton, Florida: CRC Press, pp. 263–305.Google Scholar
Brodie, E. D. I. (1989). Behavioural modification as a means of reducing the cost of reproduction. American Naturalist, 134, 225–38.CrossRefGoogle Scholar
Brodie, P. F. (1971). A reconsideration of aspects of growth, reproduction, and behavior of the white whale (Delphinapterus leucas), with reference to the Cumberland Sound, Baffin Island, population. Journal of the Fisheries Research Board of Canada, 28, 1309–18.CrossRefGoogle Scholar
Brooke, A. P. (1990). Tent selection, roosting ecology and social organization of the tent-making bat, Ectophylla alba, in Costa Rica. Journal of Zoology, London, 221, 11–19.CrossRefGoogle Scholar
Brooks, G. R. and Mitchell, J. C. (1989). Predator-prey size relations in three species of lizard from Sonora, Mexico. Southwestern Naturalist, 34, 541–6.CrossRefGoogle Scholar
Broome, L. (2001). Density, home range, seasonal movements and habitat use of the mountain pygmy-possum Burramys parvus (Marsupialia: Burramyidae) at Mount Blue Cow, Kosciuszko National Park. Austral Ecology, 26, 275–92.CrossRefGoogle Scholar
Brothers, N. (1990). Albatross mortality and associated bait loss in the Japanese longline fishery in the Southern Ocean. Biological Conservation, 55, 255–68.CrossRefGoogle Scholar
Brothers, N., Gales, R. and Reid, T. (1999). The influence of environmental variables and mitigation measures on seabird catch rates in the Japanese tuna longline fishery within the Australian Fishing Zone, 1991–1995. Biological Conservation, 88, 85–101.CrossRefGoogle Scholar
Brown, D. R., Strong, C. M. and Stouffer, P. C. (2002). Demographic effects of habitat selection by hermit thrushes wintering in a pine plantation landscape. Journal of Wildlife Management, 66, 407–16.CrossRefGoogle Scholar
Brown, G. P. and Brooks, R. J. (1993). Sexual and seasonal differences in activity in a northern population of snapping turtles, Chelydra serpentina. Herpetologica, 49, 311–18.Google Scholar
Brown, G. P. and Shine, R. (2004). Effects of reproduction on the antipredator tactics of snakes (Tropidonophis mairii, Colubridae). Behavioral Ecology and Sociobiology, 56, 257–62.CrossRefGoogle Scholar
Brown, G. P. and Weatherhead, P. J. (2000). Thermal ecology and sexual size dimorphism in northern water snakes, Nerodia sipedon. Ecological Monographs, 70, 311–30.CrossRefGoogle Scholar
Brownell Jr, R. L. and Cipriano, F. (1989). Dusky dolphin – Lagenorhynchus obscurus (Gray, 1828). In Handbook of Marine Mammals: The Second Book of Dolphins and the Porpoises, vol. 6, eds. Ridgway, S. H. and Harrison, R.. London: Academic Press, pp. 85–104.Google Scholar
Bubenik, A. B. (1984). Ernaehrung, Verhalten und Umwelt des Schalenwildes. München: BLV Verlagsgesellschaft.Google Scholar
Bull, J. J. (1983). Evolution of Sex Determining Mechanisms. Menlo Park, California: Benjamin/Cummings Publ. Co.Google Scholar
Bull, J. J. and Charnov, E. L. (1989). Enigmatic reptilian sex ratios. Evolution, 43, 1561–6.CrossRefGoogle ScholarPubMed
Bullis, H. R. (1967). Depth segregations and distribution of sex-maturity groups in the marbled catshark, Galeus arae. In Sharks, Skates and Rays, eds. Gilbert, P. W., Mathewson, R. F. and Rall, D. P.. Baltimore, MD: Johns Hopkins University Press, pp. 141–8.Google Scholar
Bunnell, F. L. and Gillingham, M. P. (1985). Foraging behavior: dynamics of dining out. In Bioenergetics of Wild Herbivores, eds. Hudson, R. J. and White, R. G.. Boca Raton: CRC Press, Inc., pp. 53–75.Google Scholar
Bunnell, F. L. and Harestad, A. S. (1983). Dispersal and dispersion of black-tailed deer: models and observations. Journal of Mammalogy, 64, 201–9.CrossRefGoogle Scholar
Burger, A. E. (1991). Maximum diving depths and underwater foraging in alcids and penguins. In Studies of High Latitude Seabirds I. Behavioural, Energetic and Oceanographic Aspects of Seabird Feeding Ecology, eds. W. A. Montevecchi A. J. Gaston. Canadian Wildlife Service Occasional Papers, pp. 9–15.
Burgman, M. A., Ferson, S. and Akcakaya, H. R. (1993). Risk Assessment in Conservation Biology. London: Chapman and Hall.Google Scholar
Burke, R. L. and Mercurio, R. J. (2002). Food habits of a New York population of Italian wall lizards, Podarcis sicula (Reptilia, Lacertidae). American Midland Naturalist, 147, 368–75.CrossRefGoogle Scholar
Burland, T. M., Barratt, E. M., Beaumont, M. A. and Racey, P. A. (1999). Population genetic structure and gene flow in a gleaning bat, Plecotus auritus. Proceedings of the Royal Society, London B, 266, 975–80.CrossRefGoogle Scholar
Burland, T. M., Barratt, E. M., Nichols, R. A. and Racey, P. A. (2001). Mating patterns, relatedness and the basis of natal philopatry in the brown long-eared bat, Plecotus auritus. Molecular Ecology, 10, 1309–21.CrossRefGoogle ScholarPubMed
Burns, J. M., Trumble, S. J., Castellini, M. A. and Testa, J. W. (1998). The diet of Weddell seals in McMurdo Sound, Antarctica as determined from scat collections and stable isotope analysis. Polar Biology, 19, 272–82.CrossRefGoogle Scholar
Bury, R. B. (1972). Habits and home range of the Pacific pond turtle, Clemmys marmorata, in a stream community. Ph.D. Thesis, Berkeley, CA: University of California.Google Scholar
Bury, R. B. (1986). Feeding ecology of the turtle, Clemmys marmorata. Journal of Herpetology, 20, 515–21.CrossRefGoogle Scholar
Busack, S. D. and Jaksic, F. M. (1982). Autecological observations of Acanthodactylus erythrurus (Sauria: Lacertidae) in southern Spain. Amphibia-Reptilia, 3, 237–55.CrossRefGoogle Scholar
Busse, C. D. (1984). Spatial structure of chacma baboon groups. International Journal of Primatology, 5, 247–61.CrossRefGoogle Scholar
Butler, P. J. and Taylor, E. W. (1975). The effect of progressive hypoxia on respiration in the dogfish (Scyliorhinus canicula) at different seasonal temperatures. Journal of Experimental Biology, 63, 117–30.Google ScholarPubMed
Buttemer, W. A. and Dawson, W. R. (1993). Temporal pattern of foraging and microhabitat use by Galapagos marine iguanas, Amblyrhynchus cristatus. Oecologia, 96, 56–64.CrossRefGoogle ScholarPubMed
Byers, J. A. (1997a). American Pronghorn: Social Adaptations and the Ghosts of Predators Past. Chicago, University of Chicago Press.Google Scholar
Byers, J. A. (1997b). Mortality risk to young pronghorns from handling. Journal of Mammalogy, 78, 894–9.CrossRefGoogle Scholar
Byers, J. A. and Walker, C. (1995). Refining the motor training hypothesis for the evolution of play. American Naturalist, 146, 25–40.CrossRefGoogle Scholar
Caister, L. E., Shields, W. M. and Gosser, A. (2003). Female tannin avoidance: a possible explanation for habitat and dietary segregation of giraffes (Giraffa camelopardalis peralta) in Niger. African Journal of Ecology, 41, 201–10.CrossRefGoogle Scholar
Cameron, E., Linklater, W. L., Stafford, K. J. and Minot, E. O. (2003). Social grouping and maternal behaviour in feral horses (Equus caballus): the influence of males on maternal protectiveness. Behavioral Ecology and Sociobiology, 53, 92–101.Google Scholar
Cameron, E. Z. and du Toit, J. T. (2005). Social influences on vigilance behaviour in giraffes (Giraffa camelopardalis). Animal Behaviour 69, 1337–44.CrossRefGoogle Scholar
Camilleri, C. and Shine, R. (1990). Sexual dimorphism and dietary divergence: differences in trophic morphology between male and female snakes. Copeia, 1990, 649–58.CrossRefGoogle Scholar
Campagna, C., Werner, R., Karesh, W.et al. (2001). Movements and location at sea of South American sea lions (Otaria flavescens). Journal of Zoology, 2, 205–20.CrossRefGoogle Scholar
Campbell, A. (1999). Staying alive: evolution, culture, and women's intrasexual aggression. Behavioral and Brain Sciences, 22, 203–52.CrossRefGoogle ScholarPubMed
Campbell, A., Shirley, L., Heywood, C. and Crook, C. (2000). Infants' visual preference for sex-congruent babies, children, toys, and activities: a longitudinal study. British Journal of Developmental Psychology, 18, 479–98.CrossRefGoogle Scholar
Campbell, D. W. and Eaton, W. O. (1999). Sex differences in the activity level of infants. Infant and Child Development, 8, 1–17.3.0.CO;2-O>CrossRefGoogle Scholar
Campredon, P. (1983). Sexe et age ratios chez le canard siffleur Anas penelope en periode hivernale en Europe de l'ouest. Revue d'Ecologie (Terre et Vie), 37, 117–28.Google Scholar
Carbone, C. and Owen, M. (1995). Differential migration of the sexes of pochard Aythya ferina: results from a European survey. Wildfowl, 46, 99–108.Google Scholar
Caro, T. (1998). Behavioral Ecology and Conservation Biology. New York: Oxford University Press.Google Scholar
Carpenter, C. C. (1956). Body temperatures of three species of Thamnophis. Ecology, 37, 732–5.CrossRefGoogle Scholar
Carrier, J. C., Pratt, H. L. and Martin, L. K. (1994). Group reproductive behaviours in free-living nurse sharks, Ginglymostoma cirratum. Copeia, 1994, 646–56.CrossRefGoogle Scholar
Carson, J., Burks, V. and Parke, R. (1993). Parent-child physical play: determination and consequences. In Parent-child Play, ed. MacDonald, K.. Albany: State University of New York Press, pp. 197–220.Google Scholar
Carter, S. L., Haas, C. A. and Mitchell, J. C. (1999). Home range and habitat selection of bog turtles in southwestern Virginia. Journal of Wildlife Management, 63, 853–60.CrossRefGoogle Scholar
Catry, P., Campos, A., Almada, V. and Cresswell, W. (2004). Winter segregation of migrant European robins Erithacus rubecula in relation to sex, age and size. Journal of Avian Biology, 35, 204–9.CrossRefGoogle Scholar
Catry, P., Phillips, R. A. and Furness, R. W. (1999). Evolution of reversed sexual size dimorphism in skuas and jaegers. Auk, 116, 158–68.CrossRefGoogle Scholar
Caughley, G. and Sinclair, A. R. E. (1994). Wildlife Ecology and Management. Cambridge, MA: Blackwell Science.Google Scholar
Cederlund, B.-M. (1987). Parturition and early development of moose (Alces alces L.) calves. Swedish Wildllife Research Supplement, 1, 399–422.Google Scholar
Cederlund, G. and Sand, H. K. G. (1992). Dispersal of subadult moose (Alces alces) in nonmigratory population. Canadian Journal of Zoology, 70, 1309–14.CrossRefGoogle Scholar
Cederlund, G., Sandegren, F. and Larsson, K. (1987). Summer movements of female moose and dispersal of their offspring. Journal of Wildlife Management, 51, 342–52.CrossRefGoogle Scholar
Censky, E. J. (1995). The evolution of sexual size dimorphism in the teiid lizard Ameiva plei. Behaviour, 132, 529–57.CrossRefGoogle Scholar
Cerling, T. E., Harris, J. M. and Passey, B. H. (2003). Diets of East African Bovidae based on stable isotope analysis. Journal of Mammalogy, 84, 456–70.2.0.CO;2>CrossRefGoogle Scholar
Cerling, T. E., Harris, J. M. and Passey, B. H.Chambers (The) Encyclopedic English Dictionary. (1994). Edinburgh: Chambers.Google Scholar
Chapman, C. A. (1990). Association patterns of spider monkeys: the influence of ecology and sex on social organization. Behavioral Ecology and Sociobiology, 26, 409–14.CrossRefGoogle Scholar
Chapman, C. A. and Wrangham, R. W. (1993). Range use of the forest chimpanzees of Kibale: implications for the understanding of chimpanzee social organization. American Journal of Primatology, 31, 263–73.CrossRefGoogle Scholar
Chapman, C. A., Wrangham, R. W. and Chapman, L. J. (1995). Ecological constraints on group size: an analysis of spider monkey and chimpanzee subgroups. Behavioral Ecology and Sociobiology, 32, 199–209.Google Scholar
Charland, M. B. and Gregory, P. T. (1990). The influence of female reproductive status on thermoregulation in a viviparous snake, Crotalus viridis. Copeia, 1990, 1089–98.CrossRefGoogle Scholar
Chase, J. D., Dixon, K. R., Gates, J. E., Jacobs, D. and Taylor, G. J. (1989). Habitat characteristics, population size, and home range of the bog turtle, Clemmys muhlenbergii, in Maryland. Journal of Herpetology, 23, 356–62.CrossRefGoogle Scholar
Cheney, D. L. (1978). The play partners of immature baboons. Animal Behaviour, 26, 1038–50.CrossRefGoogle Scholar
Chism, J. and Rowell, T. (1986). Mating and residence patterns of male patas monkeys (Erythrocebus patas). Ethology, 103, 109–26.CrossRefGoogle Scholar
Chivers, D. J. (1974). The Siamang in Malaysia. Basel: S. Karger.Google Scholar
Chivers, D. P., Brown, G. E. and Smith, R. J. F. (1995). Familiarity and shoal cohesion in fathead minnows (Pimephales promelas) – implications for antipredator behaviour. Canadian Journal of Zoology – Revue Canadienne De Zoologie, 73, 955–60.CrossRefGoogle Scholar
Choudhury, S. and Black, J. M. (1991). Testing the behavioural dominance and dispersal hypothesis in pochard. Ornis Scandinavica, 22, 155–9.CrossRefGoogle Scholar
Christal, J. and Whitehead, H. (1997). Aggregations of mature male sperm whales on the Galápagos Islands breeding ground. Marine Mammal Science, 13, 59–69.CrossRefGoogle Scholar
Christal, J., Whitehead, H. and Lettevall, E. (1998). Sperm whale social units: variation and change. Canadian Journal of Zoology, 76, 1431–40.CrossRefGoogle Scholar
Ciofi, C. and Chelazzi, G. (1994). Analysis of homing pattern in the colubrid snake Coluber viridiflavus. Journal of Herpetology, 28, 477–84.CrossRefGoogle Scholar
Clapham, P. J. (1996). The social and reproductive biology of humpback whales: an ecological perspective. Mammal Review, 26, 27–49.CrossRefGoogle Scholar
Clarke, J., Manly, B., Kerry, K. R.et al. (1998). Sex differences in Adélie penguin foraging strategies. Polar Biology, 20, 248–58.CrossRefGoogle Scholar
Clauss, M., Frey, R., Kiefer, B.et al. (2003). The maximum attainable body size of herbivorous mammals: morphophysiological constraints on foregut, and adaptations of hindgut fermenters. Oecologia, 136, 14–27.CrossRefGoogle ScholarPubMed
Clemmons, J. R. and Buchholz, R. (1997). Behavioural Approaches to Conservation in the Wild. Cambridge: Cambridge University Press.Google Scholar
Clinton, W. L. (1994). Sexual selection and growth in male northern elephant seals. In Elephant Seals: Population Ecology, Behaviour and Physiology, eds. Boeuf, B. J. and Laws, R. M.. Berkley: University of California Press, pp. 154–68.Google Scholar
Clutton-Brock, J., Dennis-Bryan, K., Armitage, P. L. and Jewell, P. A. (1990). Osteology of the Soay sheep. Bulletin of the British Museum of Natural History, 56, 1–56.Google Scholar
Clutton-Brock, T. H. (1977). Some aspects of intra-specific variation in feeding and ranging behaviour in primates. In Primate Ecology, ed. Clutton-Brock, T. H.. Cambridge: Cambridge University Press, pp. 539–56.Google Scholar
Clutton-Brock, T. H. (1983). Selection in relation to sex. In Evolution: From Molecules to Men, ed. Bendall, D. S.: Cambridge University Press, pp. 457–81.Google Scholar
Clutton-Brock, T. H. (1988). Reproductive success. In Reproductive Success, ed. Clutton-Brock, T. H.. Chicago: University of Chicago Press, pp. 472–85.Google Scholar
Clutton-Brock, T. H. (1989). Mammalian mating systems. Proceedings of the Royal Society of London, Series B, 236, 339–72.CrossRefGoogle ScholarPubMed
Clutton-Brock, T. H., Albon, S. D. and , Guinness F. E. (1985). Parental investment and sex differences in juvenile mortality in birds and mammals. Nature, 313, 131–3.CrossRefGoogle Scholar
Clutton-Brock, T. H., Albon, S. D. and , Guinness F. E. (1987b). Interactions between population density and maternal characteristics affecting fecundity and juvenile survival in red deer (Cervus elaphus). Journal of Animal Ecology, 56, 857–71.CrossRefGoogle Scholar
Clutton-Brock, T. H., Albon, S. D. and Guinness F. E. (1988). Reproductive success in male and female red deer. In Reproductive Success, ed. Clutton-Brock, T. H.. Chicago: University of Chicago Press, pp. 325–43.Google Scholar
Clutton-Brock, T. H., Guinness, F. E. and Albon, S. D. (1982). Red Deer: behavior and ecology of two sexes. Chicago: University of Chicago Press.Google Scholar
Clutton-Brock, T. H., Iason, G. R. and Guinness, F. E. (1987a). Sexual segregation and density-related changes in habitat use in male and female red deer (Cervus elaphus). Journal of Zoology, London, 211, 275–89.CrossRefGoogle Scholar
Clutton-Brock, T. H., Major, M. and Guinness, F. E. (1985). Population regulation in male and female red deer. Journal of Animal Ecology, 54, 831–46.CrossRefGoogle Scholar
Cochran, P. A. and McConville, D. R. (1983). Feeding by Trionyx spiniferus in backwaters of the upper Mississippi River. Journal of Herpetology, 17, 82–6.CrossRefGoogle Scholar
Cockburn, A. and Lazenby-Cohen, K. A. (1992). Use of nest trees by Antechinus stuartii a semelparous lekking marsupial. Journal of Zoology, London, 226, 657–80.CrossRefGoogle Scholar
Cockcroft, V. G. and Ross, G. J. B. (1990). Food and feeding of the Indian Ocean bottlenose dolphin off Southern Natal, South Africa. In The Bottlenose Dolphin, eds. Leatherwood, S. and Reeves, R. R.. San Diego: Academic Press, pp. 295–308.Google Scholar
Cogger, H. G. (2000). Reptiles and Amphibians of Australia. Sydney: Reed New Holland.Google Scholar
Collar, N. J. (1996). Family Otididae (Bustards). In Handbook of the Birds of the World, vol. 3., eds. Hoyo, J. del, Elliot, A. and Sargatal, J.. Barcelona: Lynx Edicions.Google Scholar
Compagno, L. J. V. (1984). FAO Species Catalogue. Volume 4 Sharks of the World, Parts 1 and 2. Rome: Food and Agriculture Organization of the United Nations.Google Scholar
Compagno, L. J. V. (1999). Checklist of living elasmobranchs. In Sharks, Skates and Rays: The Biology of Elasmobranch Fishes, ed. Hamlett, W. C.. Baltimore, MD: Johns Hopkins University Press, pp. 471–98.Google Scholar
Connor, R. C., Mann, J., Tyack, P. L. and Whitehead, H. (1998). Social evolution in toothed whales. Trends in Ecology and Evolution, 13, 228–32.CrossRefGoogle ScholarPubMed
Connor, R. C., Read, A. J. and Wrangham, R. (2000). Male reproductive strategies and social bonds. In Cetacean Societies: Field Studies of Dolphins and Whales, eds. Mann, J., Connor, R. C., Tyack, P. L. and Whitehead, H.. Chicago: University of Chicago Press, pp. 247–70.Google Scholar
Connor, R. C., Richards, A. F., Smolker, R. A. and Mann, J. (1996). Patterns of female attractiveness in Indian Ocean bottlenose dolphins. Behaviour, 133, 37–69.CrossRefGoogle Scholar
Connor, R. C., Smolker, R. A. and Richards, A. F. (1992). Two levels of alliance formation among male bottlenose dolphins (Tursiops sp.). Proceedings of the Academy of Science of United States, 89, 987–90.CrossRefGoogle Scholar
Conradt, L. (1997). Causes of sex differences in habitat use in red deer (Cervus elaphus, L.). Ph.D. thesis, University of Cambridge.
Conradt, L. (1998a). Could asynchrony in activity between the sexes cause intersexual social segregation in ruminants? Proceedings of the Royal Society of London, Series B, Biological Sciences, 265, 1359–63.CrossRefGoogle Scholar
Conradt, L. (1998b). Measuring the degree of sexual segregation in group-living animals. Journal of Animal Ecology, 67, 217–26.CrossRefGoogle Scholar
Conradt, L. (1999). Social segregation is not a consequence of habitat segregation in red deer and feral Soay sheep. Animal Behaviour, 57, 1151–7.CrossRefGoogle ScholarPubMed
Conradt, L. (2000). Use of a seaweed habitat by red deer (Cervus elaphus L.). Journal of Zoology, 250, 541–9.CrossRefGoogle Scholar
Conradt, L. and Roper, T. J. (2000). Activity synchrony and social cohesion: a fisson-fusion model. Proceedings of the Royal Society of London, Series B, 267, 2213–18.CrossRefGoogle ScholarPubMed
Conradt, L. and Roper, T. J. (2003). Group decision-making in animals. Nature, 421, 155–8.CrossRefGoogle ScholarPubMed
Conradt, L., Clutton-Brock, T. H. and Guinness, F. E. (1999a). The relationship between habitat choice and lifetime reproductive success in female red deer. Oecologia, 120, 218–24.CrossRefGoogle Scholar
Conradt, L., Clutton-Brock, T. H. and Guinness, F. E. (2000). Sex differences in weather sensitivity can cause habitat segregation: red deer as an example. Animal Behaviour, 59, 1049–60.CrossRefGoogle Scholar
Conradt, L., Clutton-Brock, T. H. and Thomson, D. (1999b). Habitat segregation in ungulates: are males forced into suboptimal foraging habitats through indirect competition by females? Oecologia, 119, 367–77.CrossRefGoogle Scholar
Conradt, L., Gordon, I. J., Clutton-Brock, T. H., Thomson, D. and Guinness, F. E. (2001). Could the indirect competition hypothesis explain inter-sexual site segregation in red deer (Cervus elaphus L.)? Journal of Zoology, London, 254, 185–93.CrossRefGoogle Scholar
Conroy, J. W. H. (1972). Ecological aspects of the biology of the giant petrel Macronectes giganteus (Gmelin) in the maritime Antarctic. Scientific Report of the British Antarctic Survey, 75, 1–74.Google Scholar
Constable, A. J. and Nicol, S. (2002). Defining smaller-scale management units to further develop the ecosystem approach in managing large-scale pelagic krill fisheries in Antarctica. CCAMLR Science, 9, 117–31.Google Scholar
Conway, C. J., Powell, G. V. N. and Nichols, J. D. (1995). Overwinter survival of neotropical migratory birds in early-successional and mature tropical forests. Conservation Biology, 9, 855–64.CrossRefGoogle Scholar
Cooper, J., Brooke, M. L., Burger, A. E., Crawford, R. J. M., Hunter, S. and Williams, T. (2001). Aspects of the breeding biology of the Northern Giant Petrel (Macronectes halli) and the Southern Giant Petrel (M. giganteus) at sub-Antarctic Marion Island. International Journal of Ornithology, 4, 53–68.Google Scholar
Cooper, J., Henley, S. and Klages, N. (1992). The diet of the wandering albatross Diomedea exulans at sub-Antarctic Marion Island. Polar Biology, 12, 477–84.CrossRefGoogle Scholar
Cooper, J., Wilson, R. P. and Adams, N. J. (1993). Timing of foraging by the wandering albatross Diomedea exulans. Proceedings National Institute Polar Research Symposium on Polar Biology, 6, 55–61.Google Scholar
Cooper, W. E. (2003). Sexual dimorphism in distance from cover but not escape behavior by the keeled earless lizard Holbrookia propinqua. Journal of Herpetology, 37, 374–8.CrossRefGoogle Scholar
Cords, M. (1984). Mating patterns and social structure in redtail monkeys (Cercopithecus ascanius). Zietschrift für Tierpsycholgie, 64, 213–39.Google Scholar
Corfield, T. (1973). Elephant mortality in Tsavo National Park, Kenya. East African Wildlife Journal, 11, 339–68.CrossRefGoogle Scholar
Cork, S. J. (1991). Meeting the energy-requirements for lactation in a macropodid marsupial – current nutrition versus stored body reserves. Journal of Zoology, 225, 567–76.CrossRefGoogle Scholar
Corkeron, P. J. and Connor, R. C. (1999). Why do baleen whales migrate? Marine Mammal Science, 15, 1228–45.CrossRefGoogle Scholar
Côté, S. D., Schaefer, J. A. and Messier, F. (1997). Time budgets and synchrony of activities in muskoxen: the influence of sex, age, and season. Canadian Journal of Zoology, 75, 1628–35.CrossRefGoogle Scholar
Coulson, G. (1993). The influence of population density and habitat on grouping in the western grey kangaroo, Macropus fuliginosus. Wildlife Research, 20, 151–62.CrossRefGoogle Scholar
Coulson, J. C. (2002). Colonial breeding in seabirds. In Biology of Marine Birds, eds. Schreiber, E. A. and Burger, J.. Boca Raton, Florida: CRC Press.Google Scholar
Couzin, I. D. and Krause, J. (2003). Self-organization and collective behaviour in vertebrates. Advances in the Study of Behaviour, 32, 1–75.CrossRefGoogle Scholar
Craig, M. J. (1992). Radio-telemetry and tagging study of movement patterns, activity cycles, and habitat utilization in Cagle's map turtle, Graptemys caglei. M. Sc. thesis, Canyon, TX: West Texas State University.
Cransac, N., Gerard, J.-F., Maublanc, M. L. and Pépin, D. (1998). An example of segregation between age and sex classes only weakly related to habitat use in mouflon sheep (Ovis gmelini). Journal of Zoology, London, 244, 371–8.CrossRefGoogle Scholar
Crawley, M. C. (1973). A live-trapping study of Australian brush-tailed possums, Trichosurus vulpecula (Kerr), in the Orongorongo Valley, Wellington, New Zealand. Australian Journal of Zoology, 21, 75–90.CrossRefGoogle Scholar
Creer, S., Chou, W. H., Malhotra, A. and Thorpe, R. S. (2002). Offshore insular variation in the diet of the Taiwanese bamboo viper Trimeresurus stejnegeri (Schmidt). Zoological Science, 19, 907–13.CrossRefGoogle Scholar
Cristol, D. A., Baker, M. B. and Carbone, C. (1999). Differential migration revisited. Latitudinal segregation by age and sex class. In Current Ornithology, vol. 15, eds. , V. Nolan Jr, Ketterson, E. D. and Thompson, C. F.. New York: Kluwer Academic/Plenum Publishers.Google Scholar
Croft, D. B. (1981). Social behaviour of the euro, Macropus robustus (Gould), in the Australian Arid Zone. Australian Wildlife Research, 8, 13–49.CrossRefGoogle Scholar
Croft, D. B. (1987). Socio-ecology of the antilopine wallaroo, Macropus antilopinus, in the Northern Territory, with observations on sympatric M. robustus woodwardii and M. agilis. Australian Wildlife Research, 14, 243–55.CrossRefGoogle Scholar
Croft, D. B. (1989). Social organisation of the Macropodoidea. In Kangaroos, Wallabies and Rat-kangaroos, eds. Grigg, G., Jarman, P. and Hume, I.. New South Wales, Australia: Surrey Beatty and Sons, pp. 505–25.Google Scholar
Croft, D. B. (1991a). Home range of the euro, Macropus robustus erubescens. Journal of Arid Environments, 20, 99–111.Google Scholar
Croft, D. B. (1991b). Home range of the red kangaroo, Macropus rufus. Journal of Arid Environments, 20, 83–98.Google Scholar
Croft, D. P., Arrowsmith, B. J., Bielby, J.et al. (2003). Mechanisms underlying shoal composition in the Trinidadian guppy (Poecilia reticulata). Oikos, 100, 429–38.CrossRefGoogle Scholar
Croft, D. P., Botham, M. S. and Krause, J. (2004). Is sexual segregation in the guppy, Poecilia reticulata, consistent with the predation risk hypothesis? Environmental Biology of Fishes, 71(2), 127–33.CrossRefGoogle Scholar
Crow, J. F. and Kimura, M. (1970). An Introduction to Population Genetics Theory. New York: Harper and Row.Google Scholar
Croxall, J. P. (1991). Constraints on reproduction in albatrosses. Proceedings of the 20th International Ornithological Congress, 281–302.Google Scholar
Croxall, J. P. (1995). Sexual size dimorphism in seabirds. Oikos, 73, 399–403.CrossRefGoogle Scholar
Croxall, J. P. (1998). Research and conservation: a future for albatrosses? In Albatross Biology and Conservation, eds. Robertson, G. and Gales, R.. Chipping Norton, Australia: Surrey Beatty and Sons, pp. 267–88.Google Scholar
Croxall, J. P. and Gales, R. (1998). An assessment of the conservation status of albatrosses. In Albatross Biology and Conservation, eds. Robertson, G. and Gales, R.. Chipping Norton, Australia: Surrey Beatty and Sons, pp. 46–65.Google Scholar
Croxall, J. P. and Prince, P. A. (1990). Recoveries of wandering albatrosses Diomedea exulans ringed at South Georgia (1958–1986). Ringing and Migration, 11, 43–51.CrossRefGoogle Scholar
Croxall, J. P. and Prince, P. A. (1996). Potential interactions between wandering albatrosses and longline fisheries for Patagonian toothfish at South Georgia. CCAMLR Science, 3, 101–10.Google Scholar
Croxall, J. P., Black, A. D. and Wood, A. G. (1999). Age, sex and status of wandering albatrosses Diomedea exulans L. in Falkland Islands waters. Antarctic Science, 11, 150–6.CrossRefGoogle Scholar
Croxall, J. P., Prince, P. A. and Reid, K. (1997). Dietary segregation of krill-eating South Georgia seabirds. Journal of Zoology, London, 242, 531–56.CrossRefGoogle Scholar
Croxall, J. P., Prince, P. A., Rothery, P. and Wood, A. D. (1998). Population changes in albatrosses at South Georgia. In Albatross Biology and Conservation, eds. Robertson, G. and Gales, R.. Chipping Norton: Surrey Beatty and Sons, pp. 69–83.Google Scholar
Croxall, J. P., Rothery, P., Pickering, S. P. C. and Prince, P. A. (1990). Reproductive performance, recruitment and survival of wandering albatrosses Diomedea exulans at Bird Island, South Georgia. Journal of Animal Ecology, 59, 775–96.CrossRefGoogle Scholar
Croxall, J. P., Silk, J. R. D., Phillips, R. A., Afanasyev, V. and Briggs, D. R. (2005). Global circumnavigations: tracking year-round ranges of nonbreeding albatrosses. Science, 307, 249–50.CrossRefGoogle ScholarPubMed
Cryan, P. M., Bogan, M. A. and Altenbach, J. S. (2000). Effect of elevation on distribution of female bats in the Black Hills, South Dakota. Journal of Mammalogy, 81, 719–25.2.3.CO;2>CrossRefGoogle Scholar
Cuadrado, M. (1997). Why are migrant Robins Erithacus rubecula territorial in winter?: the importance of the anti-predation behaviour. Ethology, Ecology and Evolution, 9, 77–88.CrossRefGoogle Scholar
Curlewis, J. D. (1989). The breeding season of Bennett's wallaby (Macropus rufogriseus rufogriseus) in Tasmania. Journal of Zoology, London, 218, 337–9.CrossRefGoogle Scholar
Cutter, J. (2002). Dietary Segregation in the Western Grey Kangaroo, Macropus fuliginosus, at Hattah-Kulkyne National Park. Unpublished honours thesis, University of Melbourne.Google Scholar
D'Onghia, G., Matarrese, A., Tursi, A. and Sion, L. (1995). Observations on the depth distribution pattern of the small-spotted catshark in the north Aegean Sea. Journal of Fish Biology, 47, 421–6.CrossRefGoogle Scholar
Dahlheim, M. E. and Heyning, J. E. (1999). Killer whale – Orcinus orca (Linnaeus, 1758). In Handbook of Marine Mammals: The Second Book of Dolphins and the Porpoises, vol. 6, eds. Ridgway, S. H. and Harrison, R.. London: Academic Press, pp. 281–322.Google Scholar
Dalrymple, G. H. (1977). Intraspecific variation in the cranial feeding mechanism of the turtles of the genus Trionyx (Reptilia, Testudines, Trionychidae). Journal of Herpetology, 11, 255–85.CrossRefGoogle Scholar
Daltry, J. C., Wuster, W. and Thorpe, R. S. (1998). Intraspecific variation in the feeding ecology of the crotaline snake Calloselasma rhodostoma in Southeast Asia. Journal of Herpetology, 32, 198–205.CrossRefGoogle Scholar
Dalziell, J. and Poorter, M. (1993). Seabird mortality in longline fisheries around South Georgia. Polar Record, 29, 143–5.CrossRefGoogle Scholar
Daneri, G. A. and Coria, N. R. (1992). The diet of Antarctic fur seals, Arctocephalus gazella, during the summer-autumn period at Mossman Peninsula, Laurie Island (South Orkneys). Polar Biology, 11, 565–6.CrossRefGoogle Scholar
Darwin, C. (1871). The Decent of Man and Selection in Relation to Sex, London: Murray.Google Scholar
Daunt, F., Monaghan, P., Wanless, S., Harris, M. P. and Griffiths, R. (2001). Sons and daughters: age-specific differences in parental rearing capacities. Functional Ecology, 15, 211–16.CrossRefGoogle Scholar
Daut, E. F. and Andrews, R. M. (1993). The effect of pregnancy on thermoregulatory behavior of the viviparous lizard Chalcides ocellatus. Journal of Herpetology, 27, 6–13.CrossRefGoogle Scholar
Davis, L. S. and Speirs, E. A. H. (1990). Mate choice in penguins. In Penguin Biology, eds. Davis, S. and Darby, J. T.. San Diego, California: Academic Press, pp. 377–97.Google Scholar
Davis, R. B., Herreid, C. F. and Short, H. L. (1962). Mexican free-tailed bats in Texas. Ecological Monographs, 32, 311–46.CrossRefGoogle Scholar
Dawbin, W. H. (1982). The tuatara Sphenodon punctatus (Reptilia: Rhynchocephalia): a review. In New Zealand Herpetology, ed. Newman, D. G.. Wellington, New Zealand: New Zealand Wildlife Service, pp. 149–81.Google Scholar
Day, T., O'Connor, C. and Matthews, L. (2000). Possum social behaviour. In The Brushtail Possum: Biology, Impact and Management of an Introduced Marsupial, ed. Montague, T. L.. Lincoln, NZ: Manaaki Whenua Press, pp. 35–46.Google Scholar
Deere, J. A. (2001). The use of black rhinoceros (Diceros bicornis) dung sampling to investigate sexual differences in diet quality and midden site selection in Pilanesberg National Park. Unpublished dissertation, University of Pretoria.
del Hoyo, J. (1994). Family Cracidae (Chachalacas, Guans and Curassows). In Handbook of the Birds of the World, vol. 2., eds. Hoyo, J. del, Elliot, A. and Sargatal, J.. Barcelona: Lynx Edicions.Google Scholar
Delgado, R. A. and Schaik, C. P. (2000). The behavioral ecology and conservation of the orangutan (Pongo pygmaeus): a tale of two islands. Evolutionary Anthropology, 9, 201–18.3.0.CO;2-Y>CrossRefGoogle Scholar
Demment, M. W. and Soest, P. J. (1985). A nutritional explanation for body-size patterns of ruminant and nonruminant herbivores. American Naturalist, 125, 641–72.CrossRefGoogle Scholar
Demski, L. S. and Northcutt, R. G. (1996). The brain and cranial nerves of the white shark: an evolutionary perspective. In Great White Sharks: The Biology of Carcharodon carcharias, eds. Klimley, A. P. and Ainley, D. G.. San Diego, CA: Academic Press, pp. 121–30.Google Scholar
Deneubourg, J. L. and Goss, S. (1989). Collective patterns and decision making. Ethology, Ecology and Evolution, 1, 295–311.CrossRefGoogle Scholar
Deneubourg, J. L., Goss, S., Franks, N. et al. (1991). The dynamics of collective sorting robot-like ants and ant-like robots. In Simulation of Adaptive Behavior: From Animals to Animats, eds. Meyer, J. A. and Wilson, S. W.. Cambridge, MA: MIT Press/Bradford Books, pp. 356–65.Google Scholar
Dennis, A. J. and Marsh, H. (1997). Seasonal reproduction in musky rat-kangaroos, Hypsiprymnodon moschatus: a response to changes in resource availability. Wildlife Research, 24, 561–78.CrossRefGoogle Scholar
Desrochers, A. (1989). Sex, dominance, and microhabitat use in wintering black-capped chickadees: a field experiment. Ecology, 70, 636–45.CrossRefGoogle Scholar
Deutsch, C. J., Crocker, D. E., Costa, D. P. and Le Boeuf, B. J. (1994). Sex- and age-related variation in reproductive effort of northern elephant seals. In Elephant Seals: Population Ecology, Behaviour and Physiology, eds. Boeuf, B. J. and Laws, R. M.. Berkley: University of California Press, pp. 169–210.Google Scholar
Deutsch, C. J., Haley, M. P. and Leboeuf, B. J. (1990). Reproductive effort of male northern elephant seals – estimates from mass-loss. Canadian Journal of Zoology, 68, 2580–93.CrossRefGoogle Scholar
Di Bitetti, M. S. and Janson, C. H. (2000). When will the stork arrive? Patterns of birth seasonality in neotropic primates. American Journal of Primatology, 50, 109–30.3.0.CO;2-W>CrossRefGoogle Scholar
Dickman, C. R. (1995). Agile Antechinus. In The Mammals of Australia, ed. Strahan, R.. Sydney: Reed New Holland, pp. 99–101.Google Scholar
Diego-Rasilla, F. J. and Perez-Mellado, V. (2003). Home range and habitat selection by Podarcis hispanica (Squamata, Lacertidae) in western Spain. Folia Zoologica, 52, 87–98.Google Scholar
DiFiore, A. and Rodman, P. S. (2001). Time allocation patterns of lowland woolly monkeys (Lagothrix lagotricha poepigii) in a neotropical terra firma forest. International Journal of Primatology, 22, 449–80.Google Scholar
Dobson, A. and Poole, J. (1998). Conspecific aggregation and conservation biology. In Behavioral Ecology and Conservation Biology, ed. Caro, T.. New York: Oxford University Press, pp. 193–208.Google Scholar
, Dodd C. K. Jr, Enge, K. M. and Stuart, J. N. (1988). Aspects of the biology of the flattened musk turtle, Sternotherus depressus, in northern Alabama. Bulletin of the Florida State Museum, Biological Sciences Series, 34, 1–64.Google Scholar
Dodd, J. M. (1983). Reproduction in cartilaginous fishes (Chondrichthyes). In Fish Physiology, Vol. 9 – Reproduction: Endocrine Tissues and Hormones, eds. Hoar, W. S., Randall, D. J. and Donaldson, D. M.. New York: Academic Press, pp. 31–87.Google Scholar
Doidge, D. W. (1990). Integumentary heat loss and blubber distribution in the beluga, Delphinapterus leucas, with comparisons to narwhal, Monodon monoceros. Canadian Bulletin of Fisheries and Aquatic Sciences, 224, 129–40.Google Scholar
Doidge, D. W., McCann, T. S. and Croxall, J. P. (1986). Attendance behavior of Antarctic fur seals. In Fur Seals: Maternal Strategies on Land and at Sea, eds. Gentry, R. L. and Kooyman, G. L.. Princeton: Princeton University Press, pp. 102–14.CrossRefGoogle Scholar
Doody, J. S., Young, J. E. and Georges, A. (2002). Sex differences in activity and movements in the pig-nosed turtle, Carettochelys insculpta, in the wet-dry tropics of Australia. Copeia, 2002, 93–103.CrossRefGoogle Scholar
Dressen, W. (1993). On the behaviour and social organisation of agile wallabies, Macropus agilis (Gould, 1842) in two habitats of northern Australia. Zeitschrift für Säugetierkunde, 58, 201–11.Google Scholar
Du, W. G., Yan, S. J. and Ji, X. (2000). Selected body temperature, thermal tolerance and thermal dependence of food assimilation and locomotor performance in adult blue-tailed skinks, Eumeces elegans. Journal of Thermal Biology, 25, 197–202.CrossRefGoogle Scholar
du Toit, J. T. (1988). Patterns of resource use within the browsing ruminant guild in the central Kruger National Park. Unpublished Ph.D. thesis, Johannesburg: University of the Witwatersrand.
du Toit, J. T. (1990). Feeding height stratification among African browsing ruminants. African Journal of Ecology, 28, 55–61.CrossRefGoogle Scholar
du Toit, J. T. (1995). Sexual segregation in kudu: sex differences in competitive ability, predation risk or nutritional needs? South African Journal of Wildlife Research, 25, 127–32.Google Scholar
du Toit, J. T. (2003). Large herbivores and savanna heterogeneity. In The Kruger Experience: Ecology and Management of Savanna Heterogeneity, eds. du Toit, J. T., Rogers, K. H. and Biggs, H. C.. Washington, DC: Island Press, pp. 292–309.Google Scholar
Dubois, M., Bon, R., Cransac, N. and Maublanc, M. L. (1994). Dispersion patterns among ewes of Corsican mouflon: importance and proximate influences. Applied Animal Behaviour Science, 42, 29–40.CrossRefGoogle Scholar
Dubois, M., Quenette, P. Y., Bideau, E. and Magnac, M. P. (1993). Seasonal range use by European mouflon rams in medium altitude mountains. Acta Theriology, 38, 185–98.CrossRefGoogle Scholar
Duellman, W. E. and Trueb, L. (1986). Biology of Amphibians. New York: McGraw-Hill.Google Scholar
Dufault, S. and Whitehead, H. (1995a). An encounter with recently wounded sperm whales (Physeter macrocephalus). Marine Mammal Science, 11, 560–3.CrossRefGoogle Scholar
Dufault, S. and Whitehead, H. (1995b). The geographic stock structure of female and immature sperm whales in the South Pacific. Report of the International Whaling Commission, 45, 401–5.Google Scholar
Dulvy, N. K. and Reynolds, J. D. (1997). Evolutionary transitions among egg-laying, live-bearing and maternal inputs in sharks and rays. Proceedings of the Royal Society of London B, 264, 1309–15.CrossRefGoogle Scholar
Dunbar, R. I. M. (1977). Feeding ecology of gelada baboons: a preliminary report. In Primate Ecology, ed. Clutton-Brock, T. H.. Cambridge: Cambridge University Press, pp. 251–73.Google Scholar
Dunbar, R. I. M. (1984). Reproductive Decisions: An Economic Analysis of Gelada Baboon Social Organization. Princeton: Princeton University Press.Google Scholar
Dunbar, R. I. M. and Dunbar, P. (1988). Maternal time budgets of gelada baboons. Animal Behaviour, 76, 970–80.CrossRefGoogle Scholar
Dunn, D. G., Barco, S. G., Pabst, D.-A. and McLellan, W. A. (2002). Evidence of infanticide in bottlenose dolphins of the Western North Atlantic. Journal of Wildlife Diseases, 38, 505–10.CrossRefGoogle ScholarPubMed
Durell, S. E. A. Le V. dit (2000). Individuals feeding specialisation in shorebirds: population consequences and conservation implications. Biological Reviews of the Cambridge Philosophical Society, 75, 503–18.CrossRefGoogle Scholar
Durell, S. E. A. Le V. dit and Goss-Custard, J. D. (1996). Oystercatcher Haematopus ostralegus sex ratios on the wintering grounds: the case of the Exe estuary. Ardea, 84A, 373–81.Google Scholar
Durell, S. E. A. Le V. dit, Goss-Custard, J. D. and Caldow, R. W. G. (1993). Sex-related differences in diet and feeding method in the oystercatcher Haematopus ostralegus. Journal of Animal Ecology, 62, 205–15.CrossRefGoogle Scholar
Durell, S. E. A. Le V. dit, Goss-Custard, J. D., Caldow, R. W. G., Malcolm, H. M. and Osborn, D. (2001a). Sex, diet and feeding method-related differences in body condition in the oystercatcher Haematopus ostralegus. Ibis, 143, 107–19.CrossRefGoogle Scholar
Durell, S. E. A. Le V. dit, Goss-Custard, J. D. and Clarke, R. T. (2001b). Modelling the population consequences of age- and sex-related differences in winter mortality in the oystercatcher, Haematopus ostralegus. Oikos, 95, 69–77.CrossRefGoogle Scholar
Durell, S. E. A. Le V. dit, Ormerod, S. J. and Dare, P. J. (1996). Differences in population structure between two oystercatcher Haematopus ostralegus roosts on the Burry Inlet, South Wales. Ardea, 84A, 383–8.Google Scholar
Durner, G. M. and Gates, J. E. (1993). Spatial ecology of black rat snakes on Remington Farms, Maryland. Journal of Wildlife Management, 57, 812–26.CrossRefGoogle Scholar
Durtsche, R. D. (1992). Feeding time strategies of the fringe-toed lizard, Uma inornata, during breeding and non-breeding seasons. Oecologia, 89, 85–9.CrossRefGoogle ScholarPubMed
Duvall, D., King, M. B. and Gutzwiller, M. J. (1985). Behavioral ecology and ethology of the prairie rattlesnake. National Geographic Research, 1, 80–111.Google Scholar
Dwyer, D. P. (1966). The population pattern of Miniopterus schreibersii (Chiroptera) in north-eastern New South Wales. Australian Journal of Zoology, 14, 1073–137.CrossRefGoogle Scholar
Eaton, W. C. and Enns, L. R. (1986). Sex differences in human motor activity level. Psychology Bulletin, 100, 19–28.CrossRefGoogle ScholarPubMed
Eaton, W. C. and Yu, A. (1989). Are sex function of sex differences in maturational status? Child Development, 60, 1005–11.CrossRefGoogle ScholarPubMed
Eaton, W. C., Enns, L. R. and Presse, M. (1987). Scheme for observing activity. Journal of Psychoeducational Assessment, 3, 273–80.CrossRefGoogle Scholar
Eberle, M. and Kappeler, P. (2002). Mouse lemurs in space and time: a test of the socioecological model. Behavioral Ecology and Sociobiology, 51, 131–9.CrossRefGoogle Scholar
Ebert, D. A. (2002). Ontogenetic changes in the diet of the sevengill shark (Notorynchus cepedianus). Marine and Freshwater Research, 53, 517–23.CrossRefGoogle Scholar
Economakis, A. E. and Lobel, P. S. (1998). Aggregation behaviour of the grey reef shark, Carcharhinus amblyrhynchos, at Johnston Atoll, Central Pacific Ocean. Environmental Biology of Fishes, 51, 129–39.CrossRefGoogle Scholar
Edwards, J. (1983). Diet shifts in moose due to predator avoidance. Oecologia, 60, 185–9.CrossRefGoogle ScholarPubMed
Eifler, D. A. and Eifler, M. A. (1999a). Foraging behavior and spacing patterns of the lizard Oligosoma grande. Journal of Herpetology, 33, 632–9.CrossRefGoogle Scholar
Eifler, D. A. and Eifler, M. A. (1999b). The influence of prey distribution on the foraging strategy of the lizard Oligosoma grande (Reptilia: Scincidae). Behavioral Ecology and Sociobiology, 45, 397–402.CrossRefGoogle Scholar
Eisenberg, J. F. (1981). The Mammalian Radiations: An Analysis of Trends in Evolution, Adaptation, and Behavior. Chicago: The University of Chicago Press.Google Scholar
Ellis, J. R., Pawson, M. G. and Shackley, S. E. (1996). The comparative feeding ecology of six species of shark and four species of ray (Elasmobranchii) in the north-east Atlantic. Journal of the Marine Biological Association of the United Kingdom, 76, 89–106.CrossRefGoogle Scholar
Emlen, S. T. and Oring, L. W. (1977). Ecology, sexual selection, and the evolution of mating systems. Science, 197, 215–23.CrossRefGoogle ScholarPubMed
Ena, V., Lucio, A. and Purroy, F. J. (1985). The great bustard in Leon, Spain. Bustard Studies, 2, 35–52.Google Scholar
Endler, J. A. (1980). Natural selection on colour patterns in Poecillia reticulata. Evolution, 34, 76–91.CrossRefGoogle Scholar
Entwistle, A. C., Racey, P. A. and Speakman, J. R. (1996). Habitat exploitation by a gleaning bat, Plecotus auritus. Philosophical Transactions of the Royal Society London, B, 351, 921–31.CrossRefGoogle Scholar
Entwistle, A. C., Racey, P. A. and Speakman, J. R. (2000). Social and population structure of a gleaning bat, Plecotus auritus. Journal of Zoology, London, 252, 11–17.CrossRefGoogle Scholar
Ernst, C. H., Lovich, J. E. and Barbour, R. W. (1994). Turtles of the United States and Canada. Washington, D.C.: Smithsonian Institution Press.Google Scholar
Estep, D. Q., Crowell-Davis, S. L., Earl-Costello, S.-A. and Beatey, S. A. (1993). Changes in the social behaviour of drafthorse (Equus caballus) mares coincident with foaling. Applied Animal Behaviour Science, 35, 199–213.CrossRefGoogle Scholar
Estes, R. D. (1991a). The Behavior Guide to African Mammals: including hoofed mammals, carnivores, primates. Berkley: University of California Press.Google Scholar
Estes, R. D. (1991b). The significance of horns and other male secondary sexual characters in female bovids. Applied Animal Behaviour Science, 29, 403–51.CrossRefGoogle Scholar
Fabes, R. A. (1994). Physiological, emotional, and behavioral correlates of gender segregation. In Childhood Gender Segregation: Causes and Consequences, ed. Leaper, C.. San Francisco: Jossey-Bass, pp. 33–50.Google Scholar
Fagot, B. I. (1994). Peer relations and the development of competence in boys and girls. In Childhood Gender Segregation: Causes and Consequences, ed. Leaper, C.. San Francisco: Jossey-Bass, pp. 53–66.Google Scholar
Fairbanks, L. A. (1993). Juvenile vervet monkeys: establishing relationships and practicing skills for the future. In Juvenile Primates, eds. Pereira, M. E. and Fairbanks, L. A.. Oxford: Oxford University Press, pp. 211–27.Google Scholar
Fashing, P. (2001). Activity and ranging patterns of guerezas in the Kakamega Forest: inter-group variation and implications for intra-group feeding competition. International Journal of Primatology, 22, 549–78.CrossRefGoogle Scholar
Feare, C. J., Gill, E. L., McKay, H. V. and Bishop, J. D. (1995). Is the distribution of starlings Sturnus vulgaris within roosts determined by competition? Ibis, 137, 379–82.CrossRefGoogle Scholar
Fein, G. (1981). Pretend play in childhood: an integrative review. Child Development, 52, 1095–118.CrossRefGoogle Scholar
Feldheim, K. A., Gruber, S. H. and Ashley, M. V. (2002). The breeding biology of lemon sharks at a tropical nursery lagoon. Proceedings of the Royal Society of London B, 269, 1655–61.CrossRefGoogle Scholar
Fenton, M. B. (1990). The foraging behaviour and ecology of animal-eating bats. Canadian Journal of Zoology, 68, 411–22.CrossRefGoogle Scholar
Fenton, M. B., Rautenbach, I. L., Smith, S. E., Swanepoel, C. M., Grosell, J. and Jaarsveld, J. (1994). Raptors and bats: threats and opportunities. Animal Behaviour, 48, 9–18.CrossRefGoogle Scholar
Festa-Bianchet, M. (1986). Site fidelity and seasonal range use by bighorn rams. Canadian Journal of Zoology, 64, 2126–32.CrossRefGoogle Scholar
Festa-Bianchet, M. (1988). Seasonal range selection in bighorn sheep: conflicts between forage quality, forage quantity, and predator avoidance. Oecologia, 75, 580–6.CrossRefGoogle ScholarPubMed
Festa-Bianchet, M. (1991). The social system of bighorn sheep: grouping patterns, kinship and female dominance rank. Animal Behaviour, 42, 71–82.CrossRefGoogle Scholar
Festa-Bianchet, M. and Apollonio, M. (2003). Animal Behavior and Wildlife Conservation. Washington, D.C.: Island Press.Google Scholar
Fietz, C. (1999). Mating systems of Microcebus murinus. American Journal of Primatology, 48, 127–33.3.0.CO;2-4>CrossRefGoogle ScholarPubMed
Fisher, D. O. and Owens, I. P. F. (2000). Female home range size and the evolution of social organization in macropod marsupials. Journal of Animal Ecology, 69, 1083–98.CrossRefGoogle Scholar
Fisher, D. O., Blomberg, S. P. and Owens, I. P. F. (2002). Convergent maternal care strategies in ungulates and macropods. Evolution, 56, 167–76.CrossRefGoogle ScholarPubMed
Fitch, H. S. (1960). Autecology of the copperhead. University of Kansas Publications, Museum of Natural History, 13, 85–288.Google Scholar
Fitch, H. S. (1982). Resources of a snake community in prairie-woodland habitat of northeastern Kansas. In Herpetological Communities, ed. , N. J. Scott Jr.Washington, D.C.: U.S. Department of the Interior, Fish and Wildlife Service, pp. 83–98.Google Scholar
Fitch, H. S. (1999). A Kansas Snake Community: Composition and Changes Over 50 Years. Malabar, Florida: Krieger.Google Scholar
Fitch, H. S. and Shirer, H. W. (1971). A radiotelemetric study of spatial relationships in some common snakes. Copeia, 1971, 118–28.CrossRefGoogle Scholar
Fleming, M. R. and Frey, H. (1984). Aspects of the natural history of Feathertail Gliders Acrobates pygmaeus (Marsupialia: Burramyidae) in Victoria. In Possums and Gliders, eds. Smith, A. P. and Hume, I. D.. Chipping Norton, NSW: Surrey Beatty and Sons, pp. 403–8.Google Scholar
Fleming, T. H. and Eby, P. (2003). Ecology of bat migration. In Bat Ecology, eds. Kunz, T. H. and Fenton, M. B.. Chicago: University of Chicago Press, pp. 156–208.Google Scholar
Fleming, T. H. and Hooker, R. S. (1975). Anolis cupreus: the response of a lizard to tropical seasonality. Ecology, 56, 1243–61.CrossRefGoogle Scholar
Fletcher, T. and Selwood, L. (2000). Possum reproduction and development. In The Brushtail Possum: Biology, Impact and Management of an Introduced Marsupial, ed. Montague, T. L.. Lincoln, NZ: Manaaki Whenua Press, pp. 62–81.Google Scholar
Ford, D. (1999). Foraging ecology and demography of Sternotherus odoratus in a southwestern Missouri population. M.Sc. thesis, Springfield: Southwest Missouri State University.
Ford, E. (1921). A contribution to our knowledge of the life-histories of the dogfishes landed at Plymouth. Journal of the Marine Biological Association of the United Kingdom, 12, 468–505.CrossRefGoogle Scholar
Forero, M. G. and Hobson, K. A. (2003). Using stable isotopes of Nitrogen and Carbon to study seabird ecology: applications in the Mediterranean seabird community. Scientia Marina, 67, 23–32.CrossRefGoogle Scholar
Forero, M. G., Hobson, K. A., Bortolotti, G. R.et al. (2002). Food resource utilisation by the Magellanic penguin evaluated through stable-isotope analysis: segregation by sex and age and influence on offspring quality. Marine Ecology Progress Series, 234, 289–99.CrossRefGoogle Scholar
Fox, S. F., Conder, J. M. and Smith, A. E. 1998. Sexual dimorphism in the case of tail autotomy: Uta stansburiana with and without previous tail loss. Copeia, 1998, 376–82.CrossRefGoogle Scholar
Fragaszy, D. M. and Boinski, S. (1995). Patterns of individual diet choice and efficiency of foraging of wedge-capped capuchin monkeys (Cebus olivaceous). Journal of Comparative Psychology, 109, 339–48.CrossRefGoogle Scholar
Fraker, M. A., Gordon, C. D., McDonald, J. W., Ford, J. K. B. and Cambers, G. (1979). White whale (Delphinapterus leucas) distribution and abundance and the relationship to physical and chemical characteristics of the Mackenzie Estuary. Canadian Fisheries and Marine Service Technical Report, 863, v–56.Google Scholar
Francisci, F., Focardi, S. and Boitani, L. (1985). Male and female Alpine ibex: phenology of space use and herd size. In The Biology and Management of Mountain Ungulates, ed. Lovari, S.. London: Croom Helm, pp. 124–33.Google Scholar
Freake, M. J. (1998). Variation in homeward orientation performance in the sleepy lizard (Tiliqua rugosa): effects of sex and reproductive period. Behavioral Ecology and Sociobiology, 43, 339–44.CrossRefGoogle Scholar
Frid, A. (1999). Huemul (Hippocamelus bisulcus) sociality at a periglacial site: sexual segregation and habitat effects on group size. Canadian Journal of Zoology, 77, 1083–91.CrossRefGoogle Scholar
Friend, J. (1989). Myrmecobiidae: In Fauna of Australia, eds. Walton, D. and Richardson, B.. Canberra: Australian Government Publishing Service, pp. 583–90.Google Scholar
Friend, J. (1995). Numbat: In The Mammals of Australia, ed. Strahan, R.. Sydney: Reed New Holland, pp. 160–2.Google Scholar
Friend, J. (1996). Numbats on a junk food diet. Nature Australia, 29, 40–9.Google Scholar
Friend, J. and Whitford, D. (1993). Maintenance and breeding of the numbat (Myrmecobius fasciatus) in captivity. In Biology and Management of Australasian Carnivorous Marsupials, eds. Roberts, M., Carnio, J., Crawshaw, G. and Hutchins, M.. Toronto: Metropolitan Toronto Zoo and Monotreme and Marsupial Advisory Group of AAZPA, pp. 103–24.Google Scholar
Frischknecht, M. (1993). The breeding coloration of male 3-spined sticklebacks (Gasterosteus aculeatus) as an indicator of energy investment in vigor. Evolutionary Ecology, 7, 439–50.CrossRefGoogle Scholar
Frost, K. J., Russel, R. B. and Lowry, L. F. (1992). Killer whales, Orcinus orca, in the Southeastern Bearing Sea: recent sightings and predation on other marine mammals. Marine Mammal Science, 8, 110–19.CrossRefGoogle Scholar
Furness, R. W. (1993). Birds as monitors of pollutants. In Birds as Monitors of Environment Change, eds. Furness, R. W. and Greenwood, J. J. D.. London: Chapman and Hall, pp. 86–143.CrossRefGoogle Scholar
Gadsen, H. and Palacios-Orona, L. E. (1997). Seasonal dietary patterns of the Mexican fringe-toed lizard (Uma paraphygas). Journal of Herpetology, 31, 1–9.Google Scholar
Galan, P. (1999). Demography and population dynamics of the lacertid lizard Podarcis bocagei in north-west Spain. Journal of Zoology, London, 249, 203–18.CrossRefGoogle Scholar
Galdikas, B. (1988). Orangutan diet, range, and activity at Tanjung Putting, Central Borneo. International Journal of Primatology, 9, 1–35.CrossRefGoogle Scholar
Gales, N. (2002). New Zealand sea lion. In Encyclopedia of Marine Mammals, eds. Perrin, W. F., Wursig, B. and Thewissen, J. G. M.. San Diego: Academic Press.Google Scholar
Gales, R. (1998). Albatross populations: status and threats. In Albatross Biology and Conservation, eds. Robertson, G. and Gales, R.. Chipping Norton, Australia: Surrey Beatty and Sons, pp. 20–45.Google Scholar
Gales, R. (1993). Co-operative Mechanisms for the Conservation of Albatrosses. Hobart, Australia: Australian Nature Conservation Agency and Australian Antarctic Foundation.
Gales, R., Brothers, N. and Reid, T. (1998). Seabird mortality in the Japanese tuna longline fishery around Australia. Biological Conservation, 86, 37–56.CrossRefGoogle Scholar
Galois, P., Leveille, M., Bouthillier, L., Daigle, C. and Parren, S. (2002). Movement patterns, activity and home range of the eastern spiny softshell turtle (Apalone spinifera) in northern Lake Champlain, Quebec, Vermont. Journal of Herpetology, 36, 402–11.CrossRefGoogle Scholar
Gannon, V. P. J. and Secoy, D. M. (1985). Seasonal and daily activity patterns in a Canadian population of the prairie rattlesnake, Crotalus viridis viridis. Canadian Journal of Zoology, 63, 86–91.CrossRefGoogle Scholar
Garrick, L. D. (1974). Reproductive influences on behavioral thermoregulation in the lizard, Sceloporus cyanogenys. Physiology and Behaviour, 12, 85–91.CrossRefGoogle Scholar
Garstka, W. R., Camazine, B. and Crews, D. (1982). Interactions of behavior and physiology during the annual reproductive cycle of the red-sided garter snake (Thamnophis sirtalis parietalis). Herpetologica, 38, 104–23.Google Scholar
Garton, J. S. and Dimmick, R. W. (1969). Food habits of the copperhead in middle Tennessee. Journal of the Tennessee Academy of Science, 44, 113–17.Google Scholar
Gates, S. (2002). Review of methodology of quantitative reviews using meta-analysis in ecology. Journal of Animal Ecology, 71, 547–57.CrossRefGoogle Scholar
Gauthreaux, S. A. Jr (1978). The ecological significance of behavioural dominance. Perspectives in Ethology, 3, 17–54.CrossRefGoogle Scholar
Gauthreaux, S. A. Jr (1982). The ecology and evolution of avian migration systems. In Avian Biology Vol. 6, eds. Farner, D. S., King, J. R. and Parkes, K. C.. New York: Academic Press.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, 237–69.CrossRefGoogle Scholar
Gehrt, S. D. and Fritzell, E. K. (1996). Sex-biased response of raccoons (Procyon lotor) to live traps. American Midland Naturalist, 135, 23–32.CrossRefGoogle Scholar
Geisler, J. H. and Uhen, M. D. (2003). Morphological support for a close relationship between hippo and whales. Journal of Vertebrate Paleontology, 23, 991–6.CrossRefGoogle Scholar
Geist, V. (1968). On delayed social and physical maturation in mountain sheep. Canadian Journal of Zoology, 46, 899–904.CrossRefGoogle ScholarPubMed
Geist, V. (1971). Mountain Sheep: a Study in Behavior and Evolution. Chicago: University Chicago Press.Google Scholar
Geist, V. (1974). On the relationship of social evolution and ecology in ungulates. American Zoologist, 14, 205–20.CrossRefGoogle Scholar
Geist, V. and Bromley, P. T. (1978). Why deer shed antlers. Zeitschrift für Säugetierkunde – International Journal of Mammalian Biology, 43, 223–31.Google Scholar
Geist, V. and Petocz, R. G. (1977). Bighorn sheep in winter: do rams maximize reproductive fitness by spatial and habitat segregation from ewes? Canadian Journal of Zoology, 55, 1802–10.CrossRefGoogle Scholar
Gerard, J.-F. and Loisel, P. (1995). Spontaneous emergence of a relationship between habitat openness and mean group size and its possible evolutionary consequences in large herbivores. Journal of Theoretical Biology, 176, 511–22.CrossRefGoogle Scholar
Gerard, J.-F., Bideau, E., Maublanc, M.-L., Loisel, P. and Marchal, C. (2002). Herd size in large herbivores: encoded in the individual or emergent? Biological Bulletin, 202, 275–82.CrossRefGoogle ScholarPubMed
Gerard, J.-F., Pendu, Y., Maublanc, M.-L., Vincent, J.-P., Poulle, M.-L. and Cibien, C. (1995). Large group formation in European roe deer, an adaptive feature? Revue d'Ecologie (Terre Vie), 50, 391–401.Google Scholar
Gerpe, M. S., Moreno, J. E., Moreno, V. J. and Patat, M. L. (2000). Cadmium, zinc and copper accumulation in the squid Illex argentinus from the Southwest Atlantic Ocean. Marine Biology, 136, 1039–44.CrossRefGoogle Scholar
Gibbons, J. W. (1986). Movement patterns among turtle populations: applicability to management of the desert tortoise. Herpetologica, 42, 104–13.Google Scholar
Gibson, R. and Falls, J. B. (1979). Thermal biology of the common garter snake Thamnophis sirtalis (L.). 1. Temporal variation, environmental effects and sex differences. Oecologia, 43, 79–97.CrossRefGoogle Scholar
Gilardi, J. D. (1992). Sex-specific foraging distributions of brown boobies in the eastern tropical Pacific. Colonial Waterbirds, 15, 148–51.CrossRefGoogle Scholar
Gillis, R. (1991). Thermal biology of two populations of red-chinned lizards (Sceloporus undulatus erythrocheilus) living in different habitats in south-central Colorado. Journal of Herpetology, 25, 18–23.CrossRefGoogle Scholar
Gillooly, J. F., Brown, J. H., West, G. B., Savage, V. M. and Charnov, E. L. (2001). Effects of size and temperature on metabolic rate. Science, 293, 2248–51.CrossRefGoogle ScholarPubMed
Gilmore, R. G., Dodrill, J. W. and Linley, P. A. (1983). Reproduction and embryonic development of the sand tiger shark, Odontaspis taurus. U.S. Fishery Bulletin, 81, 201–25.Google Scholar
Ginnett, T. F. and Demment, M. W. (1997). Sex differences in giraffe foraging behavior at two spatial scales. Oecologia, 110, 291–300.CrossRefGoogle ScholarPubMed
Ginnett, T. F. and Demment, M. W. (1999). Sexual segregation by Masai giraffes at two spatial scales. African Journal of Ecology, 37, 93–106.CrossRefGoogle Scholar
Gjerde, I. (1991). Cues in winter habitat selection by capercaillie. I. Habitat characteristics. Ornis Scandinavica, 22, 197–204.CrossRefGoogle Scholar
Goldfoot, D. A., Wallen, K., Neff, K., McBrair, D. A. and Goy, M. C. (1984). Social influences on the display of sexually dimorphic behavior in rhesus monkeys: isosexual rearing. Archives of Sexual Behaviour, 13, 395–412.CrossRefGoogle Scholar
Goldingay, R. L. and Kavanagh, R. P. (1990). Socioecology of the yellow-bellied glider (Petaurus australis) at Waratah Creek, NSW. Australian Journal of Zoology, 38, 327–41.CrossRefGoogle Scholar
Goldsworthy, S. (1995). Differential expenditure of maternal resources in Antarctic fur seals, Arctocephalus gazella, at Heard Island, southern Indian Ocean. Behavioral Ecology, 6, 218–28.CrossRefGoogle Scholar
Gompper, M. E. (1996). Sociality and asociality in white-nosed coatis (Nasua narica): foraging costs and benefits. Behavioral Ecology, 7, 254–63.CrossRefGoogle Scholar
González-Solís, J. (2004a). Sexual size dimorphism in northern giant petrels: ecological correlates and scaling. Oikos, 105(2), 247–54.CrossRefGoogle Scholar
González-Solís, J. (2004b). The regulation of incubation shifts near hatching: a timed mechanism, embryonic signalling or food availability? Animal Behaviour, 67, 663–71.CrossRefGoogle Scholar
González-Solís, J., Croxall, J. P. and Briggs, D. R. (2002a). Activity patterns of giant petrels Macronectes spp. using different foraging strategies. Marine Biology, 140, 197–204.Google Scholar
González-Solís, J., Croxall, J. P. and Wood, A. G. (2000a). Sexual dimorphism and sexual segregation in foraging strategies of northern giant petrels, Macronectes halli, during incubation. Oikos, 90, 390–8.CrossRefGoogle Scholar
González-Solís, J., Croxall, J. P. and Wood, A. G. (2000b). Foraging partitioning between giant petrels Macronectes spp. and its relationship with breeding population changes at Bird Island, South Georgia. Marine Ecology Progress Series, 204, 279–288.CrossRefGoogle Scholar
González-Solís, J., Sanpera, C. and Ruiz, X. (2002b). Metals and selenium as bioindicators of geographic and trophic segregation in giant petrels Macronectes spp. Marine Ecology Progress Series, 244, 257–64.CrossRefGoogle Scholar
Goodall, J. (1986). The Chimpanzees of Gombe. Cambridge, MA: Harvard University Press.Google Scholar
Goodman-Lowe, G. D. (1998). Diet of the Hawaiian monk seal (Monachus schauinslandi) from the Northwestern Hawaiian islands during 1991 to 1994. Marine Biology, 132, 535–46.CrossRefGoogle Scholar
Gordon, D. M. and MacCulloch, R. D. (1980). An investigation of the ecology of the map turtle, Graptemys geographica (Le Seur), in the northern part of its range. Canadian Journal of Zoology, 58, 2210–19.CrossRefGoogle Scholar
Gordon, I. J. and Illius, A. W. (1988). Incisor arcade structure and diet selection in ruminants. Functional Ecology, 2, 15–22.CrossRefGoogle Scholar
Gosler, A. G. (1987). Pattern and process in the bill morphology of the great tit Parus major. Ibis, 129, 451–76.CrossRefGoogle Scholar
Gosler, A. G. (1990). The variable niche hypothesis revisited; an analysis of intra- and inter-specific differences in bill variation in Parus. In Population Biology of Passerine Birds, eds. Blondel, J., Gosler, A. G., Lebreton, J. and McCleery, R.. Berlin: Springer-Verlag, pp. 167–74.CrossRefGoogle Scholar
Gosler, A. G. (1991). On the use of greater covert moult and pectoral muscle as measures of condition in passerines with data for the great tit Parus major. Bird Study, 31, 1–9.CrossRefGoogle Scholar
Gosler, A. G. (1996). Environmental and social determinants of winter fat storage in the great tit Parus major. Journal of Animal Ecology, 65, 1–17.CrossRefGoogle Scholar
Gosler, A. G. and Carruthers, T. D. (1994). Bill size and niche breadth in the Irish coal tit Parus ater hibernicus. Journal of Avian Biology, 25, 171–7.CrossRefGoogle Scholar
Gosler, A. G. and Carruthers, T. D. (1999). Body reserves and social dominance in the great tit Parus major in relation to winter weather in Southwest Ireland. Journal of Avian Biology, 30, 447–59.CrossRefGoogle Scholar
Gosling, L. M. (1969). Parturition and related behaviour in Coke's Hartebeest, Alcelaphus buselaphus cokei Günther. Journal of Reproduction and Fertility, Supplement, 6, 265–86.Google Scholar
, Gosling L. M. and Sutherland, W. J. (2000). Behaviour and Conservation. Cambridge, United Kingdom: Cambridge University Press.Google Scholar
Gottman, J. M. (1983). How children become friends. Monographs of the Society for Research in Child Development, 48 (3) Serial No. 201.CrossRefGoogle Scholar
Gowans, S., Whitehead, H. and Hooker, S. K. (2001). Social organization in northern bottlenose whales, Hyperoodon ampullatus: not driven by deep-water foraging? Animal Behaviour, 62, 369–77.CrossRefGoogle Scholar
Graham, T. E. and Graham, A. A. (1992). Metabolism and behavior of wintering common map turtles, Graptemys geographica, in Vermont. Canadian Field-Naturalist, 104, 517–19.Google Scholar
Graham, T. E. and Graham, A. A.Grand Usuel Larousse (1997). Paris: Larousse-Bordas.Google Scholar
Grant, G. S. and Banak, S. A. (1995). Harem structure and reproductive behaviour of Pteropus tonganus in American Samoa. Department of Marine and Wildlife Research, American Samoan Government. Biological Reports, 69, 214–44.Google Scholar
Grassi, C. (2002). Sex differences in feeding, height, and space use in Hapalemur griseus. International Journal of Primatology, 23, 677–93.CrossRefGoogle Scholar
Graves, B. M. and Duvall, D. (1987). An experimental study of aggregation and thermoregulation in prairie rattlesnakes (Crotalus viridis viridis). Herpetologica, 43, 259–64.Google Scholar
Graves, B. M. and Duvall, D. (1995). Aggregation of squamate reptiles associated with gestation, oviposition, and parturition. Herpetological Monographs, 9, 102–19.CrossRefGoogle Scholar
Greenberg, R. (1986). Competition in migrant birds in the nonbreeding season. In Current Ornithology 3, ed. Johnston, R. F.. New York and London: Plenum Press, pp. 281–307.CrossRefGoogle Scholar
Greene, H. W. (1997). Snakes: The Evolution of Mystery in Nature. Berkeley, CA: University of California Press.Google Scholar
Gregory, P. T. (1974). Patterns of spring emergence of the red-sided garter snake (Thamnophis sirtalis parietalis) in the Interlake region of Manitoba. Canadian Journal of Zoology, 52, 1063–9.CrossRefGoogle Scholar
Gregory, P. T. and Isaac, L. A. (2004). Food habits of the grass snake in southeastern England: is Natrix natrix a generalist predator? Journal of Herpetology, 38, 88–95.CrossRefGoogle Scholar
Gregory, P. T., Crampton, L. H. and Skebo, K. M. (1999). Conflicts and interactions among reproduction, thermoregulation and feeding in viviparous reptiles: are gravid snakes anorexic? Journal of Zoology, London, 248, 231–41.CrossRefGoogle Scholar
Gregory, P. T., Macartney, J. M. and Larsen, K. W. (1987). Spatial patterns and movements. In Snakes: Ecology and Evolutionary Biology, eds. Seigel, R. A., Collins, J. T. and Novak, S. S.. New York: Macmillan, pp. 366–95.Google Scholar
Griffiths, R. (1992). Sex-biased mortality in the Lesser Black-backed Gull Larus fuscus during the nestling stage. Ibis, 134, 237–44.CrossRefGoogle Scholar
Griffiths, S. W. (2003). Learned recognition of conspecifics by fishes. In Fish are Smarter than You Think: Learning in Fishes, eds. C. Brown, K. N. Laland and J. Krause. Fish and Fisheries, Special edn., 4, 256–68.CrossRef
Griffiths, S. W. and Magurran, A. E. (1998). Sex and schooling behaviour in the Trinidadian guppy. Animal Behaviour, 56, 698–93.CrossRefGoogle ScholarPubMed
Grindal, S. D., Morissette, J. L. and Brigham, R. M. (1999). Concentration of bat activity in riparian habitats over an elevational gradient. Canadian Journal of Zoology, 77, 972–7.CrossRefGoogle Scholar
Gross, J. E. (1998). Sexual segregation in ungulates: a comment. Journal of Mammalogy, 79, 1404–9.CrossRefGoogle Scholar
Gross, J. E., Alkon, P. U. and Demment, M. W. (1996). Nutritional ecology of dimorphic herbivores: digestion and passage rates in Nubian ibex. Oecologia, 107, 170–8.CrossRefGoogle ScholarPubMed
Gross, J. E., Alkon, P. U. and Demment, M. W. (1995a). Grouping patterns and spatial segregation by Nubian ibex. Journal of Arid Environments, 30, 423–39.CrossRefGoogle Scholar
Gross, J. E., Demment, M. W., Alkon, P. U. and Kotzman, M. (1995b). Feeding and chewing behaviours of Nubian ibex: compensation for sex-related differences in body size. Functional Ecology, 9, 385–93.CrossRefGoogle Scholar
Grubb, P. (1974). Mating activity and the social significance of rams in a feral sheep community. In The Behaviour of Ungulates and its Relation to Management, vol. 2, eds. Geist, V. and Walther, F.. Morges, Switzerland: IUCN No 50, pp. 457–76.Google Scholar
Grubb, P. and Jewell, P. A. (1966). Social grouping and home range in feral Soay sheep. Symposium of Zoology, Society of London, 18, 179–210.Google Scholar
Grubb, T. C. and Woodrey, M. S. (1990). Sex, age, intraspecific dominance status, and the use of food by birds wintering in temperate-deciduous and cold-coniferous woodlands: a review. Studies in Avian Biology, 13, 270–9.Google Scholar
Gruber, S. H., Nelson, D. R. and Morrissey, J. F. (1988). Patterns of activity and space utilization of lemon sharks, Negaprion brevirostris, in a shallow Bahamian lagoon. Bulletin of Marine Science, 43, 61–76.Google Scholar
Gruys, R. C. (1993). Autumn and winter movements and sexual segregation of willow ptarmigan. Arctic, 46, 228–39.CrossRefGoogle Scholar
Gueron, S., Levin, S. A. and Rubenstein, D. I. (1996). The dynamics of herds: from individuals to aggregations. Journal of Theoretical Biology, 182, 85–98.CrossRefGoogle Scholar
Gustin, K. and McCracken, G. F. (1987). Scent recognition in the Mexican free-tailed bat, Tadarida brasiliensis mexicana. Animal Behaviour, 35, 13–19.CrossRefGoogle Scholar
Hamilton, I. A. and Barclay, R. M. R. (1994). Patterns of daily torpor and day-roost selection by male and female big brown bats (Eptesicus fuscus). Canadian Journal of Zoology, 72, 744–9.CrossRefGoogle Scholar
Hamilton, W. D. (1971). Geometry of the selfish herd. Journal of Theoretical Biology, 31, 295–311.CrossRefGoogle ScholarPubMed
Hammerson, G. A. (1978). Observations on the reproduction, courtship, and aggressive behavior of the striped racer, Masticophis lateralis euryxanthus (Reptilia, Serpentes, Colubridae). Journal of Herpetology, 12, 253–5.CrossRefGoogle Scholar
Hammond, K. A., Spotila, J. R. and Standora, E. A. (1988). Basking behavior of the turtle Pseudemys scripta: effects of digestive state, acclimation temperature, sex, and season. Physiological Zoology, 61, 69–77.CrossRefGoogle Scholar
Handasyde, K. A. and Martin, R. W. (1996). Field observations on the common striped possum (Dactylopsila trivirgata) in North Queensland. Wildlife Research, 23, 755–66.CrossRefGoogle Scholar
Hanley, T. A. (1982). The nutritional basis for food selection by ungulates. Journal of Range Management, 35, 146–51.CrossRefGoogle Scholar
Hanski, I. and Gilpin, M. (1991). Metapopulation dynamics: brief history and conceptual domain. Biological Journal of the Linnean Society, 42, 3–16.CrossRefGoogle Scholar
Haramis, G. M., Nichols, J. D., Pollock, K. H. and Hines, J. E. (1986). The relationship between body mass and survival of wintering canvasbacks. Auk, 103, 506–14.Google Scholar
Harestad, A. S. and Bunnell, F. L. (1979). Home range and body weight – a reevaluation. Ecology, 60, 389–402.CrossRefGoogle Scholar
Harper, P. C. (1987). Feeding behaviour and other notes on 20 species of Procellariiformes at sea. Notornis, 34, 169–92.Google Scholar
Harrel, J. B., Allen, C. M. and Hebert, S. J. (1996). Movements and habitat use of subadult alligator snapping turtles (Macroclemys temminckii) in Louisiana. American Midland Naturalist, 135, 60–7.CrossRefGoogle Scholar
Harris, J. E. (1952). A note on the breeding season, sex ratio and embryonic development of the dogfish Scyliorhinus canicula (L.). Journal of the Marine Biological Association of the United Kingdom, 31, 269–70.CrossRefGoogle Scholar
Hart, D. R. (1983). Dietary and habitat shift with size of red-eared turtles (Pseudemys scripta) in a southern Louisiana population. Herpetologica, 39, 285–90.Google Scholar
Hashimoto, C., Furuichi, T., Tashiro, Y. (2001). What factors influence the size of chimpanzee parties in the Kalinzu Forest, Uganda? Examination of fruit abundance and number of estrous females. International Journal of Primatology, 22, 947–59.CrossRefGoogle Scholar
Hasegawa, T. (1990). Sex differences in ranging patterns. In The Chimpanzees of the Mahale Mountains, ed. Nishida, T.. Tokyo: University of Tokyo Press, pp. 99–114.Google Scholar
Hass, C. C. and Jenni, D. A. (1993). Social play among juvenile bighorn sheep: structure, development, and relationship to adult behavior. Ethology, 93, 105–16.CrossRefGoogle Scholar
Hauksson, E. and Bogason, V. (1997). Comparative feeding of grey (Halichoerus grypus) and common seals (Phoca vitulina) in coastal waters of iceland, with a note on the diet of hooded (Cystophora cristata) and harp seals (Phoca groenlandica). Journal of Northwest Atlantic Fishery Science, 22, 125–35.CrossRefGoogle Scholar
Hayes, C. L., Rubin, E. S., Jorgensen, M. C., Botta, R. A. and Boyce, W. M. (2000). Mountain lion predation of bighorn sheep in the Peninsular Ranges, California. Journal of Wildlife Management, 64, 954–9.CrossRefGoogle Scholar
Heatwole, H. (1968). Relationship of escape behavior and camouflage in anoline lizards. Copeia, 1968, 109–13.CrossRefGoogle Scholar
Hebrard, J. J. and Madsen, T. (1984). Dry season intersexual habitat partitioning by flap-necked chameleons (Chamaeleo dilepis) in Kenya. Biotropica, 16, 69–72.CrossRefGoogle Scholar
Hedd, A., Gales, R. and Brothers, N. (2001). Foraging strategies of shy albatross Thalassarche cauta breeding at Albatross Island, Tasmania, Australia. Marine Ecology Progress Series, 224, 267–82.CrossRefGoogle Scholar
Hedrick, A. V. and Temeles, E. J. (1989). The evolution of sexual dimorphism in animals: hypotheses and tests. Trends in Ecology and Evolution, 4, 136–8.CrossRefGoogle ScholarPubMed
Heinsohn, G. E. (1966). Ecology and reproduction of the Tasmanian bandicoots (Perameles gunni and Isoodon obesulus). University of California Publications in Zoology, 80, 1–96.Google Scholar
Helfman, G. S., Collette, B. B. and Facey, D. E. (1997). The Diversity of Fishes. Malden, MA: Blackwell Science.Google Scholar
Henderson, R. W. (1993). Foraging and diet in West Indian Corallus enydris (Serpentes: Boidae). Journal of Herpetology, 27, 24–8.CrossRefGoogle Scholar
Henderson, R. W. and Binder, M. H. (1980). The ecology and behavior of vine snakes (Ahaetulla, Oxybelis, Thelotornis, Uromacer): a review. Milwaukee Public Museum Contributions in Biology and Geology, 37, 1–38.Google Scholar
Henry, S. R. (1984). Social organization of the greater glider (Petauroides volans) in Victoria. In Possums and Gliders, eds. Smith, A. P. and Hume, I. D.. Chipping Norton, NSW: Surrey Beatty and Sons, pp. 221–8.Google Scholar
Hepp, G. R. and Hair, J. D. (1984). Dominance in wintering waterfowl (Anatini): effects on distribution of sexes. Condor, 86, 251–7.CrossRefGoogle Scholar
Heppel, S. S., Walters, J. R. and Crowder, L. B. (1994). Evaluating management alternatives for red-cockaded woodpeckers: a management approach. Journal of Wildlife Management, 58, 479–87.CrossRefGoogle Scholar
Herbinger, I., Boesch, C. and Rothe, H. (2001). Territory characteristics among three neighboring chimpanzee communities in the Taï National Park, Côte d'Ivoire. International Journal of Primatology, 22, 143–67.CrossRefGoogle Scholar
Herremans, M. (1997). Habitat segregation of male and female red-backed shrikes Lanius collurio and lesser grey shrikes Lanius minor in the Kalahari basin, Botswana. Journal of Avian Biology, 28, 240–8.CrossRefGoogle Scholar
Hertz, P. E., Huey, R. B. and Stevenson, R. D. (1993). Evaluating temperature regulation by field-active ectotherms: the fallacy of the inappropriate question. American Naturalist, 142, 796–818.CrossRefGoogle ScholarPubMed
Heulin, B., Surget-Groba, Y., Guiller, A., Guillaume, C. P. and Deunff, J. (1999). Comparisons of mitochondrial DNA (mtDNA) sequences (16S rRNA gene) between oviparous and viviparous strains of Lacerta vivipara: a preliminary study. Molecular Ecology, 8, 1627–31.CrossRefGoogle ScholarPubMed
Hickey, M. B. C. and Fenton, M. B. (1990). Foraging by red bats (Lasiurus borealis) – do intraspecific chases mean territoriality?Canadian Journal of Zoology, 68, 2477–82.CrossRefGoogle Scholar
Hickling, C. F. (1930). A contribution towards the life-history of the spur-dog. Journal of the Marine Biological Association of the United Kingdom, 16, 529–76.CrossRefGoogle Scholar
Hillman, J. C. (1987). Group size and association patterns of the common eland (Tragelaphus oryx). Journal of Zoology, London, 213, 641–63.CrossRefGoogle Scholar
Hinch, G. N., Lynch, J. J., Elwin, R. L. and Green, G. C. (1990). Long-term associations between Merino ewes and their offspring. Applied Animal Behaviour Science, 27, 93–103.CrossRefGoogle Scholar
Hinde, R. A. (1974). Biological bases of human social behaviour. New York: McGraw-Hill.Google Scholar
Hirth, D. H. (1977). Social behavior of white-tailed deer in relation to habitat. Wildlife Monographs, 53, 1–55.Google Scholar
Hjelm, J. and Persson, L. (2001). Size-dependent attack rate and handling capacity: inter-cohort competition in zooplantivorous fish. Oikos, 95, 520–32.CrossRefGoogle Scholar
Hjelset, A. M., Andersen, M., Gjertz, I., Lydersen, C. and Gulliksen, B. (1999). Feeding habits of bearded seals (Erignathus barbatus) from the Svalbard area, Norway. Polar Biology, 21, 186–93.CrossRefGoogle Scholar
Hobson, E. S. (1968). Predatory behaviour of some shore fishes in the Gulf of California. United States Bureau of Sport Fisheries and Wildlife Research Report, 73, 1–91.Google Scholar
Hobson, K. A., Piatt, J. F. and Pitocchelli, J. (1994). Using stable isotopes to determine seabird trophic relationships. Journal of Animal Ecology, 63, 786–98.CrossRefGoogle Scholar
Hobson, K. A., Sease, J. L., Merrick, R. L. and Piatt, J. F. (1997). Investigating trophic relationships of pinnipeds in Alaska and Washington using stable isotope ratios of nitrogen and carbon. Marine Mammal Science, 13, 114–32.CrossRefGoogle Scholar
Hocking, G. J. (1981). The population ecology of the brush-tail possum, Trichosurus vulpecula (Kerr) in Tasmania. Unpublished Masters thesis, University of Tasmania.Google Scholar
Hodum, P. J. and Hobson, K. A. (2000). Trophic relationships among Antarctic fulmarine petrels: insights into dietary overlap and chick provisioning strategies inferred from stable-isotope (d15N and d13C) analyses. Marine Ecology Progress Series, 198, 273–81.CrossRefGoogle Scholar
Hoek, W. (1992). An unusual aggregation of harbor porpoises (Phocoena phocoena). Marine Mammal Science, 8, 152–4.CrossRefGoogle Scholar
Hoelzel, A. R., Boeuf, B. J., Reiter, J. and Campagna, C. (1999). Alpha-male paternity in elephant seals. Behavioral Ecology and Sociobiology, 46, 298–306.CrossRefGoogle Scholar
Hofmann, R. R. (1989). Evolutionary steps of ecophysiological adaptation and diversification of ruminants: a comparative view of their digestive system. Oecologia, 78, 443–57.CrossRefGoogle ScholarPubMed
Hoge, A. R. and Federsoni, P. A. J. (1977). Observations on a brood of Bothrops atrox (Linnaeus, 1758): (Serpentes: Viperidae: Crotalinae). Memorias do Instituto Butantan (Sao Paulo), 40/41, 19–36.Google Scholar
Hogg, J. T. (1984). Mating in bighorn sheep: multiple creative male strategies. Science, 225, 526–9.CrossRefGoogle ScholarPubMed
Hohn, A. A. and Brownell, R. L. (1990). Harbor porpoise in central Californian waters: life history and incidental catches. Paper SC/42/SM47 presented at 42nd meeting of the scientific committee, International Whaling Commission, Nordwijk, Holland.Google Scholar
Holekamp, K. E. and Sherman, P. W. (1989). Why do male ground squirrels disperse?American Scientist, 77, 232–9.Google Scholar
Holmes, R. T. (1986). Foraging patterns of forest birds: male-female differences. Wilson Bulletin, 98, 196–213.Google Scholar
Hölzenbein, S. and Marchinton, L. (1992). Spatial integration of maturing-male white-tailed deer into the adult population. Journal of Mammalogy, 73, 326–34.CrossRefGoogle Scholar
Honda, K., Yamamoto, Y. and Tatsukawa, R. (1987). Distribution of heavy metals in Antarctic marine ecosystems. Proceedings of the NIPR Symposium of Polar Biology, 1, 184–97.Google Scholar
Hooge, P. N. and Eichenlaub, B. (1997). Animal movement extension to Arcview ver. 1.1. Alaska Science Center – Biological Science Center, U.S. Geological Survey, Anchorage, AK, USA.Google Scholar
Horak, I. G., Boomker, J. and Flamand, J. R. B. (1995). Parasites of domestic and wild animals in South Africa. XXXIV. Arthropod parasites of nyalas in north-eastern KwaZulu-Natal. Onderstepoort Journal of Veterinary Research, 62, 171–9.Google ScholarPubMed
Houde, A. E. (1997). Sex, Color, and Mate Choice in Guppies. Princeton: Princeton University Press.Google Scholar
Houston, D. L. and Shine, R. (1993). Sexual dimorphism and niche divergence: feeding habits of the Arafura filesnake. Journal of Animal Ecology, 62, 737–49.CrossRefGoogle Scholar
Houston, D. L. and Shine, R. (1994). Population demography of Arafura filesnakes (Serpentes, Acrochordidae) in tropical Australia. Journal of Herpetology, 28, 273–80.CrossRefGoogle Scholar
How, R. A. (1972). The ecology and management of Trichosurus species (Marsupialia) in New South Wales. Unpublished Ph.D. thesis, University of New England.Google Scholar
How, R. A., Barnett, J. L., Bradley, A. J., Humphreys, W. F. and Martin, R. (1984). The population biology of Pseudocheirus peregrinus in a Leptospermum laevigatum thicket. In Possums and Gliders, eds. Smith, A. P. and Hume, I. D.. Chipping Norton, NSW: Surrey Beatty and Sons, pp. 261–8.Google Scholar
Howell, D. J. (1979). Flock-foraging in nectar-feeding bats: advantages to the bats and to the host plants. American Naturalist, 114, 23–49.CrossRefGoogle Scholar
Hoyer, R. E. and Stewart, G. R. (2000). Biology of the rubber boa (Charina bottae) with emphasis on C. b. umbratica. Part I: Capture, size, sexual dimorphism, and reproduction. Journal of Herpetology, 34, 348–54.CrossRefGoogle Scholar
Hrabar, H. and du Toit, J. T. (in press). Dynamics of an introduced population of black rhinoceros (Diceros bicornis): Pilanesberg National Park, South Africa. Animal Conservation.Google Scholar
Hudson, R. J. (1985). Body size, energetics, and adaptive radiation. In Bioenergetics of Wild Herbivores, eds. Hudson, R. J. and White, R. G.. Boca Raton: CRC Press, Inc., pp. 1–24.Google Scholar
Hudson, R. J. and White, R. G. (1985). Bioenergetics of Wild Herbivores. Boca Raton: CRC Press Inc.Google Scholar
Huey, R. and Slatkin, M. (1976). Costs and benefits of lizard thermoregulation. Quarterly Review of Biology, 51, 363–84.CrossRefGoogle ScholarPubMed
Hulscher, J. B. (1985). Growth and abrasion of the oystercatcher bill in relation to dietary switches. Netherlands Journal of Zoology, 35, 124–54.CrossRefGoogle Scholar
Humple, D. L., Nur, N., Geupel, G. R. and Lynes, M. P. (2001). Female-biased sex ratio in a wintering population of ruby-crowned kinglets. Wilson Bulletin, 113, 419–24.CrossRefGoogle Scholar
Hungate, R. E. (1975). The rumen microbial ecosystem. Annual Review of Ecology and Systematics, 6, 39–66.CrossRefGoogle Scholar
Hunter, S. (1983). The food and feeding of the giant petrels Macronectes halli and M. giganteus at South Georgia. Journal of Zoology, London, 200, 521–38.CrossRefGoogle Scholar
Hunter, S. (1984). Breeding biology and population dynamics of giant petrels Macronectes at South Georgia (Aves: Procellariiformes). Journal of Zoology, London, 203, 441–60.CrossRefGoogle Scholar
Hunter, S. (1985). The role of giant petrels in the Southern Ocean ecosystem. In Antarctic Nutrient Cycles and Food Webs, eds. Siegfried, W. R., Laws, R. M. and Condy, P. R.. Berlin: Springer-Verlag, pp. 534–42.CrossRefGoogle Scholar
Hunter, S. (1987). Species and sexual isolation mechanisms in sibling species of giant petrels Macronectes. Polar Biology, 7, 295–301.CrossRefGoogle Scholar
Hunter, S. and Brooke, M. L. (1992). The diet of giant petrels Macronectes spp. at Marion Island, Southern Indian Ocean. Colonial Waterbirds, 15, 56–65.CrossRefGoogle Scholar
Huntingford, F. A. and Turner, A. K. (1987). Animal Conflict. London: Chapman and Hall.CrossRefGoogle Scholar
Hyrenbach, K. D., Fernández, P. and Anderson, D. J. (2002). Oceanographic habitats of two sympatric North Pacific albatrosses during the breeding season. Marine Ecology Progress Series, 233, 283–301.CrossRefGoogle Scholar
Illius, A. W. and Gordon, I. J. (1987). The allometry of food intake in grazing ruminants. Journal of Animal Ecology, 56, 989–99.CrossRefGoogle Scholar
Illius, A. W. and Gordon, I. J. (1992). Modelling the nutritional ecology of ungulate herbivores: evolution of body size and competitive interactions. Oecologia, 89, 428–34.CrossRefGoogle ScholarPubMed
Irwin, L. N. (1965). Diel activity and social interaction of the lizard Uta stansburiana steinegeri. Copeia, 1965, 99–101.CrossRefGoogle Scholar
Isaac, J. and Johnson, C. (2004). Sexual dimorphism and synchrony of breeding: variation in polygyny potential among populations in the common brushtail possum, Trichosurus vulpecula. Behavioral Ecology, 14, 818–22.CrossRefGoogle Scholar
Isbell, L. A., Cheney, D. L. and Seyfarth, R. M. (2002). Why vervet monkeys (Cercopithecus aethiops) live in multimale groups. In The Guenons: Diversity and Adaptation in African Monkeys, eds. Glenn, M. E. and Cords, M.. New York: Kluwer, pp. 173–88.Google Scholar
Jack, K. M. and Pavelka, M. S. M. (1997). The behavior of peripheral males during the mating season in Macaca fuscata. Primates, 38, 369–77.CrossRefGoogle Scholar
Jackes, A. D. (1973). The use of wintering ground by red deer in Ross-shire, Scotland. M.Phil. thesis, University of Edinburgh.Google Scholar
Jacklin, C. N. and Maccoby, E. E. (1978). Social behaviour at thirty-three months in same-sex and mid-sex dyads. Child Development, 49, 557–69.CrossRefGoogle Scholar
Jakimchuk, R. D., Ferguson, S. H. and Sopuck, L. G. (1987). Differential habitat use and sexual segregation in the Central Arctic caribou herd. Canadian Journal of Zoology, 65, 534–41.CrossRefGoogle Scholar
Janson, C. H. (1990). Ecological consequences of individual spatial choice in foraging brown capuchin monkeys (Cebus apella). Animal Behaviour, 38, 922–34.CrossRefGoogle Scholar
Janson, C. H. (1992). Evolutionary ecology of primate social structure. In Evolutionary Ecology and Human Behavior, eds. Smith, E. A. and Winterhalder, B.. Hawthorne, NY: Aldine.Google Scholar
Janson, C. H. (1998). Testing the predation hypothesis for vertebrate sociality: prospects and pitfalls. Behaviour, 135, 389–410.CrossRefGoogle Scholar
Jantschke, F. (1973). On the breeding and rearing of bush dogs, Speothos venaticus, at the Frankfurt Zoo. International Zoo Yearbook, 13, 141–3.CrossRefGoogle Scholar
Jaremovic, R. V. and Croft, D. B. (1991a). Social organisation of eastern grey kangaroos in southeastern New South Wales. II. Associations within mixed groups. Mammalia, 55, 543–54.Google Scholar
Jaremovic, R. V. and Croft, D. B. (1991b). Social organization of the eastern grey kangaroo (Macropodidae, Marsupialia) in southeastern New South Wales. I. Groups and group home ranges. Mammalia, 55, 169–85.Google Scholar
Jarman, P. J. (1968). The effect of the creation of Lake Kariba upon the terrestrial ecology of the middle Zambezi Valley, with particular references to the large mammals. Ph. D. thesis, University of Manchester.Google Scholar
Jarman, P. J. (1974). The social organisation of antelope in relation to their ecology. Behaviour, 48, 215–67.CrossRefGoogle Scholar
Jarman, P. J. (1983). Mating system and sexual dimorphism in large, terrestrial, mammalian herbivores. Biological Review, 58, 485–520.CrossRefGoogle Scholar
Jarman, P. J. (1989). Sexual dimorphism in Macropodoidea. In Kangaroos, Wallabies and Rat-kangaroos, eds. Grigg, G., Jarman, P. and Hume, I.. New South Wales, Australia: Surrey Beatty and Sons, pp. 433–47.Google Scholar
Jarman, P. J. (1991). Social behaviour and organization in the Macropodoidea. Advances in the Study of Behavior, 20, 1–50.CrossRefGoogle Scholar
Jarman, P. J. and Coulson, G. (1989). Dynamics and adaptiveness of grouping in macropods. In Kangaroos, Wallabies and Rat-kangaroos, eds. Grigg, G., Jarman, P. and Hume, I.. New South Wales, Australia: Surrey Beatty and Sons, pp. 527–47.Google Scholar
Jarman, P. J. and Southwell, C. J. (1986). Grouping, associations, and reproductive strategies in eastern grey kangaroos. In Ecological Aspects of Social Evolution, eds. Rubenstein, D. I. and Wrangham, R. W.. Princeton: Princeton University Press, pp. 399–428.Google Scholar
Jefferson, T. A., Stacey, P. J. and Baird, R. W. (1991). A review of killer whale interactions with other marine mammals: predation to co-existence. Mammal Review, 21, 151–80.CrossRefGoogle Scholar
Jenni, L. (1993). Structure of a brambling Fringilla montifringilla roost according to sex, age and body-mass. Ibis, 135, 85–90.CrossRefGoogle Scholar
Jensen, K. H., Jakobsen, P. J. and Kleiven, O. T. (1998). Fish kairomone regulation of internal swarm structure in Daphnia pulex (Cladocera: Crustacea). Hydrobiologia, 368, 123–7.CrossRefGoogle Scholar
Jenssen, T. A. (1970). The ethoecology of Anolis nebulosusi (Sauria, Iguanidae). Journal of Herpetology, 4, 1–38.CrossRefGoogle Scholar
Jenssen, T. A. and Nunez, S. C. (1998). Spatial and breeding relationships of the lizard, Anolis carolinensis: evidence of intrasexual selection. Behaviour, 135, 981–1003.CrossRefGoogle Scholar
Jessopp, M. J., Forcada, J., Reid, K., Trathan, P. N., Murphy, E. J. (2004). Winter dispersal of Leopard seals (Hydrurga leptonyx): environmental factors influencing demographics and seasonal abundance. Journal of Zoology, 263, 251–8.CrossRefGoogle Scholar
Jewell, P. A. (1986). Survival in a feral population of primitive sheep on St. Kilda, Outer Hebrides, Scotland. National Geographic Research, 2, 402–6.Google Scholar
Jewell, P. A. (1997). Survival and behaviour of castrated Soay sheep (Ovis aries) in a feral island population on Hirta, St. Kilda, Scotland. Journal of Zoology, London, 243, 623–36.CrossRefGoogle Scholar
Jiang, Z., Liu, B., Zeng, Y., Han, G. and Hu, H. (2000). Attracted by the same sex, or repelled by the opposite sex? – Sexual segregation in Pere David's deer. Chinese Science Bulletin, 45, 485–91.CrossRefGoogle Scholar
Johannes, R. E., Squire, L. and Graham, T. (1994). Developing a protocol for monitoring spawning aggregations of Palauan serranids to facilitate the formulation and evaluation of strategies for their management. South Pacific Forum Fisheries Agency Report 94/28. Honiara, Solomon Islands.Google Scholar
Johnson, C. N. (1983). Variations in group size and composition in red and western grey kangaroos, Macropus rufus (Desmarest) and M. fuliginosus (Desmarest). Australian Wildlife Research, 10, 25–31.CrossRefGoogle Scholar
Johnson, C. N. (1989). Grouping and the structure of association in the red-necked wallaby. Journal of Mammalogy, 70, 18–26.CrossRefGoogle Scholar
Johnson, C. N. and Bayliss, P. G. (1981). Habitat selection by sex, age and reproductive class in the red kangaroo, Macropus rufus, in western New South Wales. Australian Wildlife Research, 8, 465–74.CrossRefGoogle Scholar
, Johnson F. A. and Moore, C. T. (1996). Harvesting multiple stocks of ducks. Journal of Wildlife Management, 60, 551–9.Google Scholar
Johnson, P. M. (1980). Field observations on group compositions in the agile wallaby, Macropus agilis (Gould) (Marsupialia: Macropodidae). Australian Wildlife Research, 7, 327–31.CrossRefGoogle Scholar
Johnson, P. M. and Strahan, R. (1982). A further description of the Musky Rat-kangaroo, Hypsiprymnodon moschatus (Ramsay, 1876) (Marsupialia, Potoroidae), with notes on its biology. Australian Zoologist, 21, 27–46.CrossRefGoogle Scholar
Johnstone, G. W. (1977). Comparative feeding ecology of the giant petrels Macronectes giganteus (Gmelin) and M. halli (Mathews). In Adaptations within Antarctic Ecosystems, ed. Llano, G.. Houston: Gulf Publishing Company, pp. 647–68.Google Scholar
Jones, G. and Rydell, J. (1994). Foraging strategy and predation risk as factors influencing emergence time in echolocating bats. Philosophical Transactions of the Royal Society London, B, 346, 445–55.CrossRefGoogle Scholar
Jones, M. (1995). Tasmanian devil. In The Mammals of Australia, ed. Strahan, R.. Sydney: Reed New Holland, pp. 82–4.Google Scholar
Jones, M. and Barmuta, L. (1998). Diet overlap and relative abundance of sympatric dasyurid carnivores: a hypothesis of competition. Journal of Animal Ecology, 67, 410–21.CrossRefGoogle Scholar
Jones, M. and Barmuta, L. (2000). Niche differentiation among sympatric Australian dasyurid carnivores. Journal of Mammalogy, 81, 434–47.2.0.CO;2>CrossRefGoogle Scholar
Jones, M., Grigg, G. and Beard, L. (1997). Body temperatures and activity patterns of Tasmanian devils (Sarcophilus harrisii) and eastern quolls (Dasyurus viverrinus) through a subalpine winter. Physiological Zoology, 70, 53–60.CrossRefGoogle ScholarPubMed
Jones, R. L. (1996). Home range and seasonal movements of the turtle Graptemys flavimaculata. Journal of Herpetology, 30, 376–85.CrossRefGoogle Scholar
Jonsson, K. I. (1997). Capital and income breeding as alternative tactics of resource use in reproduction. Oikos, 78, 57–66.CrossRefGoogle Scholar
Joung, S.-J., Chen, S.-T., Clark, E., Uchida, S. and Huang, W. Y. P. (1996). The whale shark, Rhincodon typus, is a livebearer: 300 embryos found in one ‘megamamma’ supreme. Environmental Biology of Fishes, 46, 219–23.CrossRefGoogle Scholar
Jouventin, P. and Weimerskirch, H. (1990a). Long-term changes in seabird and seal populations in the Southern Ocean. In Antarctic Ecosystems: Ecological Change and Conservation, eds. Kerry, K. R. and Hempel, G.. Berlin: Springer-Verlag, pp. 208–13.CrossRefGoogle Scholar
Jouventin, P. and Weimerskirch, H. (1990b). Satellite tracking of wandering albatrosses. Nature, 343, 746–8.CrossRefGoogle Scholar
Jouventin, P., Lequette, B. and Dobson, F. S. (1999). Age-related mate choice in the wandering albatross. Animal Behaviour, 57, 1099–106.CrossRefGoogle ScholarPubMed
Kajimura, H. (1985). Opportunistic feeding by the northern fur seal (Callorhinus ursinus). In Marine Mammals and Fisheries, eds. Beddington, J. R., Beverton, R. J. H. and Lavigne, D. M.. London: George Allen and Unwin Ltd, pp. 300–18.Google Scholar
Kastelein, R. A. (2002). Walrus. In Encyclopedia of Marine Mammals, eds. Perrin, W. F., Wursig, B. and Thewissen, J. G. M.. San Diego: Academic Press, pp. 1294–300.Google Scholar
Kato, A., Watanuki, Y., Mishiumi, I., Kuroki, M., Shaughnessy, P. and Naito, Y. (2000). Variation in foraging and parental behaviour of king cormorants. Auk, 117, 718–30.CrossRefGoogle Scholar
Katsikaros, K. and Shine, R. (1997). Sexual dimorphism in the tusked frog, Adelotus brevis (Anura: Myobatrachidae): the roles of natural and sexual selection. Biological Journal of the Linnean Society, 60, 39–51.Google Scholar
Kaufman, G. W., Siniff, D. B. and Reichle, R. (1975). Colony behaviour of Weddell seals, Leptonychotes weddellii, at Hutton Cliffs, Antarctica. Rapport et Procès-verbeax des Réunions du Conseil international pour l'Exploration de la mer, 11, 228–46.Google Scholar
Kaufman, J. H. (1992). Habitat use by wood turtles in central Pennsylvania. Journal of Herpetology, 26, 315–21.CrossRefGoogle Scholar
Kaufmann, J. H. (1974). Social ethology of the whiptail wallaby, Macropus parryi, in northeastern New South Wales. Animal Behaviour, 22, 281–369.CrossRefGoogle Scholar
Keenlyne, K. D. (1972). Sexual differences in feeding habits of Crotalus horridus horridus. Journal of Herpetology, 6, 234–7.CrossRefGoogle Scholar
Kendrick, K. M., Atkins, K., Hinton, M. R.et al. (1995). Facial and vocal discrimination in sheep. Animal Behaviour, 49, 1665–76.CrossRefGoogle Scholar
Kerle, J. A. (1983). The population biology of the Northern brushtail possum. Unpublished Ph.D. thesis, Macquarie University.Google Scholar
Kerle, J. A. (1984). Variation in the ecology of Trichosurus: its adaptive significance. In Possums and Gliders, eds. Smith, A. P. and Hume, I. D.. Chipping Norton, NSW: Surrey Beatty and Sons, pp. 115–28.Google Scholar
Kerle, J. A., McKay, G. M. and Sharman, G. B. (1991). A systematic analysis of the brushtail possum, Trichosurus vulpecula (Kerr, 1792) (Marsupialia: Phalangeridae). Australian Journal of Zoology, 39, 313–31.CrossRefGoogle Scholar
Kerth, G. and Reckardt, K. (2003). Information transfer about roosts in female Bechstein's bats: an experimental field study. Proceedings of the Royal Society London, B, 270, 511–15.CrossRefGoogle Scholar
Kerth, G., Mayer, F. and Konig, B. (2000). Mitochondrial DNA (mtDNA) reveals that female Bechstein's bats live in closed societies. Molecular Ecology, 9, 793–800.CrossRefGoogle ScholarPubMed
Kerth, G., Wagner, M. and Konig, B. (2001a). Roosting together, foraging apart: information transfer about food is unlikely to explain sociality in female Bechstein's bats (Myotis bechsteinii). Behavioural Ecology and Sociobiology, 50, 283–91.CrossRefGoogle Scholar
Kerth, G., Weissman, K. and König, B. (2001b). Day roost selection in female Bechstein's bats (Myotis bechsteinii): a field experiment to determine the influence of roost temperature. Oecologia, 126, 1–9.CrossRefGoogle Scholar
Ketterson, E. D. and Nolan, V. Jr (1983). The evolution of differential bird migration. In Current Ornithology 1, ed. Johnston, R. F.. New York: Plenum Press.CrossRefGoogle Scholar
Kie, J. G. and Bowyer, R. T. (1999) Sexual segregation in white-tailed deer: density-dependent changes in use of space, habitat selection, and dietary niche. Journal of Mammalogy, 80, 1004–20.CrossRefGoogle Scholar
King, J. E. (1983). Seals of the World, 2nd edn. Oxford: Oxford University Press.Google Scholar
King, R. B. (1986). Population ecology of the Lake Erie water snake, Nerodia sipedon insularum. Copeia, 1986, 757–72.CrossRefGoogle Scholar
King, R. B. (1989). Sexual dimorphism in snake tail length: sexual selection, natural selection, or morphological constraint?Biological Journal of the Linnean Society, 38, 133–54.CrossRefGoogle Scholar
King, R. B. (1993). Microgeographic, historical, and size-correlated variation in water snake diet composition. Journal of Herpetology, 27, 90–4.CrossRefGoogle Scholar
Kinzey, W. G. (1977). Diet and feeding behavior of Callicebus torquatus. In Primate Ecology, ed. Clutton-Brock, T. H.. Cambridge: Cambridge University Press, pp. 127–51.Google Scholar
Kirsch, J. A. W., Lapointe, F. J. and Springer, M. S. (1997). DNA-hybridisation studies of marsupials and their implications for metatherian classification. Australian Journal of Zoology, 45, 211–80.CrossRefGoogle Scholar
Kissner, K. J., Forbes, M. R. L. and Secoy, D. M. (1997). Rattling behavior of prairie rattlesnakes (Crotalus viridis viridis, Viperidae) in relation to sex, reproductive status, body size, and body temperature. Ethology, 103, 1042–50.CrossRefGoogle Scholar
Kissner, K. J., Weatherhead, P. J. and Francis, C. M. (2003). Sexual size dimorphism and timing of spring migration in birds. Journal of Evolutionary Biology, 16, 154–62.CrossRefGoogle ScholarPubMed
Kitchen, D. W. (1974). Social behavior and ecology of the pronghorn. Wildlife Monographs, 38, 1–96.Google Scholar
Kitching, J. A. and Ebling, F. J. (1967). Ecological studies in Lough Ine. Advances in Ecological Research, 4, 197–291.CrossRefGoogle Scholar
Kleiber, M. (1975). An Introduction to Animal Energetics. Huntington, NY: R. E. Kreiger Publishing Co.Google Scholar
Kleiman, D. G. (1980). The sociobiology of captive propagation. In Conservation Biology: An Evolutionary-Ecological Perspective, eds. Soule, M. and Wilcox, B. A.. Sunderland, Massachusetts: Sinauer Associates, Inc., pp. 243–61.Google Scholar
Kleinenberg, S. E., Yablokov, A. V., Bel'kovich, B. M. and Tarasevich, M. N. (1964). Beluga (Delphinapterus leucas): investigation of the species. Translated by Israel Progr. Sci. Translat., Jerusalem 1969. Moscow: Academy Nauk USSR, p. 376.Google Scholar
Klimley, A. P. (1985). Schooling in Sphyrna lewini, a species with low risk of predation: a non-egalitarian state. Zeitschrift für Tierpsychologie, 70, 297–319.CrossRefGoogle Scholar
Klimley, A. P. (1987). The determinants of sexual segregation in the scalloped hammerhead shark, Sphyrna lewini. Environmental Biology of Fishes, 18, 27–40.CrossRefGoogle Scholar
Kodric-Brown, A. (1990). Mechanisms of sexual selection – insights from fishes. Annales Zoologici Fennici, 27, 87–100.Google Scholar
Kohlmann, S. G., Müller, D. M. and Alkon, P. U. (1996). Antipredator constraints on lactating nubian ibexes. Journal of Mammalogy, 77, 1122–31.CrossRefGoogle Scholar
Komdeur, J. and Deerenberg, C. (1997). The importance of social behavior studies for conservation. In Behavioral Approaches to Conservation in the Wild, eds. Clemmons, J. R. and Buchholz, R.. Cambridge, United Kingdom: Cambridge University Press, pp. 262–76.Google Scholar
Komers, P. E., Messier, F. and , Gates C. C. (1993). Group structure in wood bison: nutritional and reproductive determinants. Canadian Journal of Zoology, 71, 1367–71.CrossRefGoogle Scholar
Koopman, K. F. (1993). Bats. In Mammal Species of the World: Taxonomic and Geographic Reference, eds. Wilson, D. E. and Reeder, D. M.. Washington, DC: Smithsonian Institution Press, pp. 137–241.Google Scholar
Koskimies, J. (1957). Flocking behaviour in capercaillie Tetrao urogallus and blackgame Lyrurus tetrix. Papers on Game Research Published by the Finish Game Foundation, 18, 1–31.Google Scholar
Kovacs, K. M., Lydersen, C., Hammill, M. and Lavigne, D. M. (1996). Reproductive effort of male hooded seals (Cystophora cristata): estimates from mass loss. Canadian Journal of Zoology, 74, 1521–30.CrossRefGoogle Scholar
Krause, J. (1994). The influence of food competition and predation risk on size-assortative shoaling in juvenile chub (Leuciscus cephalus). Ethology, 96, 105–16.CrossRefGoogle Scholar
Krause, J. and Godin, J.-G. J. (1994). Shoal choice in banded killifish (Fundulus diaphanus, Teleostei, Cyprinodontidae) – Effects of predation risk, fish size, species composition and size of shoals. Ethology, 98, 128–36.CrossRefGoogle Scholar
Krause, J. and Godin, J.-G. J. (1996a). Influence of parasitism on shoal choice in the banded killifish (Fundulus diaphanus, Teleostei, Cyprinodontidae). Ethology, 102, 40–9.CrossRefGoogle Scholar
Krause, J. and Godin, J.-G. J. (1996b). Phenotypic variability within and between fish shoals. Ecology, 77, 1586–91.CrossRefGoogle Scholar
Krause, J. and Ruxton, G. D. (2002). Living in Groups. Oxford: Oxford University Press.Google Scholar
Krause, J., Butlin, R., Peuhkuri, N. and Pritchard, V. L. (2000a). The social organisation of fish shoals: a test of the predictive power of laboratory experiments for the field. Biological Reviews, 75, 477–501.CrossRefGoogle Scholar
Krause, J., Godin, J.-G. J. and Brown, D. (1996). Size-assortativeness in multi-species fish shoals. Journal of Fish Biology, 49, 221–5.CrossRefGoogle Scholar
Krause, J., Hoare, D. J., Croft, D., Lawrence, J., Ward, A., Ruxton, G. D., Godin, J. G. J. and James, R. (2000b). Fish shoal composition: mechanisms and constraints. Proceedings of the Royal Society of London Series B, Biological Sciences, 267, 2011–17.CrossRefGoogle Scholar
Krause, J., Loader, S. P., McDermott, J. and Ruxton, G. D. (1998). Refuge use by fish as a function of body length-related metabolic expenditure and predation risks. Proceedings of the Royal Society of London Series B, Biological Sciences, 265, 2373–9.CrossRefGoogle Scholar
Krebs, C. J. (1989). Ecological Methodology. New York: Harper Collins Publishers.Google Scholar
Krohmer, R. W. and Aldridge, R. D. (1985). Female reproductive cycle of the lined snake (Tropidoclonium lineatum). Herpetologica, 41, 39–44.Google Scholar
Krützen, M., Sherwin, W. B., Berggren, P. and Gales, N. (2004). Population structure in an inshore cetacean revealed by microsatellite and mtDNA analysis: bottlenose dolphins (Tursiops sp.) in Shark Bay Western Australia. Marine Mammal Science, 20, 28–47.CrossRefGoogle Scholar
Kruuk, H. (1972) The Spotted Hyena. Chicago: University of Chicago Press.Google Scholar
Kunz, T. H. (1974). Feeding ecology of a temperate insectivorous bat (Myotis velifer). Ecology, 55, 693–711.CrossRefGoogle Scholar
Kunz, T. H. (1982). Roosting ecology of bats. In Ecology of Bats, ed. Kunz, T. H.. New York: Plenum Press, pp. 1–55.CrossRefGoogle Scholar
Kunz, T. H. and Lumsden, L. F. (2003). Ecology of cavity and foliage roosting bats. In Bat Ecology, eds. Kunz, T. H. and Fenton, M. B.. Chicago: University of Chicago Press, pp. 3–89.Google Scholar
Lack, D. (1968). Ecological Adaptations for Breeding in Birds. London: Methuen and Co.Google Scholar
, Lagarde F., Bonnet, X., Corbin, J., Henen, B. and Nagy, K. (2002). A short spring before a long jump: the ecological challenge to the steppe tortoise (Testudo horsfieldi). Canadian Journal of Zoology, 80, 493–502.Google Scholar
Lagarde, F., Bonnet, X., Henen, B.et al. (2003). Sex divergence in space utilisation in the steppe tortoise (Testudo horsfieldi). Canadian Journal of Zoology, 81, 380–7.CrossRefGoogle Scholar
LaGory, K. E., Bagshaw, C. III and Brisbin, I. L. Jr (1991). Niche differences between male and female white-tailed deer on Ossabaw Island, Georgia. Applied Animal Science, 29, 205–14.CrossRefGoogle Scholar
Lahanas, P. N. (1982). Aspects of the life history of the southern black-knobbed sawback, Graptemys nigrinoda delticola Folkerts and Mount. M.Sc. thesis, Auburn, AL: Auburn University.Google Scholar
Laidlaw, W. S., Hutchings, S. and Newell, G. R. (1996). Home range and movement patterns of Sminthopsis leucopus (Marsupialia: Dasyuridae) in coastal dry Heathland, Anglesea, Victoria. Australian Mammalogy, 19, 1–9.Google Scholar
Lailvaux, S. P., Alexander, G. J. and Whiting, M. J. (2003). Sex-based differences and similarities in locomotor performance, thermal preferences, and escape behaviour in the lizard Platysaurus intermedius wilhelmi. Physiological and Biochemical Zoology, 76, 511–21.CrossRefGoogle ScholarPubMed
Lamb, T. (1984). The influence of sex and breeding condition on microhabitat and diet in the pig frog Rana grylio. American Midland Naturalist, 111, 311–18.CrossRefGoogle Scholar
Lande, R. and Barrowclough, G. F. (1987). Effective population size, genetic variation, and their use in population management. In Viable Populations for Conservation, ed. Soule, M. E.. Cambridge, United Kingdom: Cambridge University Press, pp. 87–123.CrossRefGoogle Scholar
Landeau, L. and Terborgh, J. (1986). Oddity and the confusion effect in predation. Animal Behaviour, 34, 1372–80.CrossRefGoogle Scholar
Landys-Ciannelli, M. M., Piersma, T. and Jukema, J. (2003). Strategic size changes of internal organs and muscle tissue in the bar-tailed godwit during fat storage on a spring stopover site. Functional Ecology, 17, 151–9.CrossRefGoogle Scholar
Langman, V. A. (1977). Cow-calf relationships in giraffe (Giraffa camelopardalis giraffa). Zeitschrift für Tierpsychology, 43, 264–86.Google Scholar
Laska, A. L. (1970). The structural niche of Anolis scripta on Inagua. Breviora, 349, 1–6.Google Scholar
Latta, S. C. and Faaborg, J. (2002). Demographic and population responses of Cape May warbler wintering in multiple habitats. Ecology, 83, 2502–15.CrossRefGoogle Scholar
Launchbaugh, K. L., Provenza, F. D. and Pfister, J. A. (2001). Herbivore response to anti-quality factors in forages. Journal of Range Management, 54, 431–40.CrossRefGoogle Scholar
Lausen, C. L. and Barclay, R. M. R. (2002). Roosting behaviour and roost selection of female big brown bats (Eptesicus fuscus) roosting in rock crevices in south-eastern Alberta. Canadian Journal of Zoology, 80, 1069–76.CrossRefGoogle Scholar
Lausen, C. L. and Barclay, R. M. R. (2003). Thermoregulation and roost selection by reproductive female big brown bats (Eptesicus fuscus) roosting in rock crevices. Journal of Zoology, London, 260, 235–44.CrossRefGoogle Scholar
Lawrence, A. L. (1990). Mother-daughter and peer relationships of Scottish hill sheep. Animal Behaviour, 39, 481–6.CrossRefGoogle Scholar
Laws, R. M. (1993). Antarctic Seals: Research Methods and Techniques, 1st edn., Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Lazenby-Cohen, K. A. and Cockburn, A. (1988). Lek promiscuity in a semelparous mammal, Antechinus stuartii (Marsupialia: Dasyuridae)?Behavioral Ecology and Sociobiology, 22, 195–202.CrossRefGoogle Scholar
Lazenby-Cohen, K. A. and Cockburn, A. (1991). Social and foraging components of the home range in Antechinus stuartii (Dasyuridae: Marsupialia). Australian Journal of Ecology, 16, 301–8.CrossRefGoogle Scholar
Boeuf, B. J. (1971). The aggression of the breeding bulls. Natural History, 70, 83–94.Google Scholar
Le Boeuf, B. J. (1991). Pinniped mating systems on land, ice and in the water: emphasis on the Phocidae. In Behaviour of Pinnipeds, ed. Renouf, D.. London: Chapman and Hall, pp. 45–65.CrossRefGoogle Scholar
Le Boeuf, B. J. and Crocker, D. E. (1996). Diving behaviour of elephant seals: implications for predator avoidance. In Great White Sharks: The Biology of Carcharodon carcharias, eds. Klimley, A. P. and Ainley, D. G.. Berkley: University of California Press, pp. 193–206.Google Scholar
Le Boeuf, B. J. and Laws, R. M. (1994). Elephant seals: An introduction to the genus. In Elephant Seals: Population Ecology, Behaviour, and Physiology, eds. Boeuf, B. J. and Laws, R. M.. Berkley: University of California Press, pp. 1–26.Google Scholar
Le Boeuf, B. J. and Reiter, J. (1988). Lifetime reproductive success in northern elephant seals. In Reproductive Success: Studies of Individual Variation in Contrasting Breeding Systems, ed. Clutton-Brock, T. H.. Chicago: University of Chicago Press, pp. 344–62.Google Scholar
Le Boeuf, B. J., Crocker, D. E., Blackwell, S. B., Morris, P. A. and Thorson, P. H. (1993). Sex differences in foraging in northern elephant seals. In Marine Mammals: Advances in Behavioural and Population Biology, ed. Boyd, I. L.. London: Oxford University Press.Google Scholar
Boeuf, B. J., Crocker, D. E., Costa, D. P.et al. (2000). Foraging ecology of northern elephant seals. Ecological Monographs, 70, 353–82.CrossRefGoogle Scholar
Boeuf, B. J., Morris, P. A., Blackwell, S. B., Crocker, D. E. and Costa, D. P. (1996). Diving behavior of juvenile northern elephant seals. Canadian Journal of Zoology, 74, 1632–44.CrossRefGoogle Scholar
Pendu, Y., Guilhem, C., Briedermann, L., Maublanc, M.-L. and Gerard, J.-F. (2000). Interactions and associations between age and sex classes in mouflon sheep (Ovis gmelini) during winter. Behavioural Processes, 52, 97–107.CrossRefGoogle ScholarPubMed
Leaper, C. (1994). Exploring the consequences of gender segregation on social relationships. In Childhood Gender Segregation: Causes and Consequences, ed. Leaper, C.. San Francisco: Jossey-Bass, pp. 67–86.Google Scholar
Lee, A. and Cockburn, A. (1985). Evolutionary Ecology of Marsupials. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Lee, A., Woolley, P. and Braithwaite, R. (1982). Life history strategies of dasyurid marsupials. In Carnivorous Marsupials, ed. Archer, M.. Sydney: Royal Zoological Society of New South Wales, pp. 1–11.Google Scholar
Lee, D. S., Franz, R. and Sanderson, R. A. (1975). A note on the feeding habits of male Barbour's map turtles. Florida Field Naturalist, 3, 45–6.Google Scholar
Lee, Y. F. and McCracken, G. F. (2002). Foraging activity and food resource use of Brasilian free-tailed bats, Tadarida brasiliensis (Molossidae). Ecoscience, 9, 306–13.CrossRefGoogle Scholar
Legault, F. and Strayer, F. F. (1991). Genèse de la ségrégation sexuelle et différences comportementales chez des enfants d'âge préscolaire. Behaviour, 119, 285–301.CrossRefGoogle Scholar
Legler, J. M. L. (1985). Australian chelid turtles: reproductive patterns in wide-ranging taxa. In The Biology of Australasian Frogs and Reptiles, eds. Grigg, G. C., Shine, R. and Ehmann, H.. Sydney: Royal Zoological Society of New South Wales, pp. 117–23.Google Scholar
Lemos-Espinal, J. A., Smith, G. R. and Ballinger, R. E. (1997). Thermal ecology of the lizard, Sceloporus gadoviae, in an arid tropical scrub forest. Journal of Arid Environments, 35, 311–19.CrossRefGoogle Scholar
Lesage, V., Hammill, M. O. and Kovacs, K. M. (2001). Marine mammals and the community structure of the Estuary and Gulf of St Lawrence, Canada: evidence from stable isotope analysis. Marine Ecology Progress Series, 210, 203–21.CrossRefGoogle Scholar
Lettevall, E., Richter, C., Jaquet, N.et al. (2002). Social structure and residency in aggregations of male sperm whales. Canadian Journal of Zoology, 80, 1189–96.CrossRefGoogle Scholar
Leuzinger, Y. and Brossard, C. (1994). Répartition de M. daubentonii en fonction du sexes et de la période de l'année dans le Jura Bernois. Résultats préliminaires. Mitteilunngen der Naturforschenden Gesellschaft Schaffhausen, 39, 135–43.Google Scholar
Lewis, S., Benvenuti, S., Dall'Antonia, L.et al. (2002). Sex-specific foraging behaviour in a monomorphic seabird. Proceeding of the Royal Society of London, 269, 1687–93.CrossRefGoogle Scholar
Lewis, S. E. (1992). Behaviour of Peter's tent-making bat, Uroderma bilobatum, at maternity roosts in Costa Rica. Journal of Mammalogy, 73, 541–6.CrossRefGoogle Scholar
Lewis, S. E. (1996). Low roost-site fidelity in pallid bats: associated factors and effect on group stability. Behavioural Ecology and Sociobiology, 39, 335–44.CrossRefGoogle Scholar
Li, J., Luan, Y., Sun, I., Zhao, D. and Diao, Y. (1990). Studies on some problems of Agkistrodon shedaoensis population due to seasonal changes (in Chinese). In From Water Onto Land, ed. Zhao, E.. China Forestry Press, pp. 273–6.Google Scholar
Lillegraven, J. A., Kielan-Jaworowska, Z. and Clemens, W. A. (1979). Mesozoic Mammals: The First Two-Thirds of Mammalian History. Berkeley: University of California Press.Google Scholar
Lillywhite, H. B. and Henderson, R. W. (1993). Behavioral and functional ecology of arboreal snakes. In Snakes: Ecology and Behavior, eds. Seigel, R. A. and Collins, J. T.. New York: McGraw-Hill, pp. 1–48.Google Scholar
Lindeman, P. V. (2003). Sexual difference in habitat use of Texas map turtles (Emydidae: Graptemys versa) and its relationship to size dimorphism and diet. Canadian Journal of Zoology, 81, 1185–91.CrossRefGoogle Scholar
Ling, J. K. (2002). Australian sea lion. In Encyclopedia of Marine Mammals, eds. Perrin, W. F., Wursig, B. and Thewissen, J. G. M.. San Diego: Academic Press.Google Scholar
Lingle, S. (2000). Seasonal variation in coyote feeding behaviour and mortality of white-tailed deer and mule deer. Canadian Journal of Zoology, 78, 85–99.CrossRefGoogle Scholar
Lingle, S. (2001). Anti-predator strategies and grouping patterns in white-tailed deer and mule deer. Ethology, 107, 295–314.CrossRefGoogle Scholar
Lingle, S. (2002). Coyote predation and habitat segregation of white-tailed deer and mule deer. Ecology, 83, 2037–48.CrossRefGoogle Scholar
Lingle, S. and Pellis, S. M. (2002). Fight or flight? Antipredator behavior and the escalation of coyote encounters with deer. Oecologia, 131, 154–64.CrossRefGoogle ScholarPubMed
Lingle, S. and Wilson, W. F. (2001). Detection and avoidance of predators in white-tailed deer (Odocoileus virginianus) and mule deer (O. hemionus). Ethology, 107, 125–47.CrossRefGoogle Scholar
Lister, B. C. (1976). The nature of niche expansion in West Indian Anolis lizards I. Ecological consequences of reduced competition. Evolution, 30, 659–76.CrossRefGoogle ScholarPubMed
Lister, B. C. and Aguayo, A. G. (1992). Seasonality, predation, and the behaviour of a tropical mainland anole. Journal of Animal Ecology, 61, 717–33.CrossRefGoogle Scholar
Loison, A., Gaillard, J.-M., Pelabon, C. and Yoccoz, N. G. (1999). What factors shape sexual size dimorphism in ungulates?Evolutionary Ecology Research, 1, 611–33.Google Scholar
Long, J. A. and Pellegrini, A. D. (2003). Studying change in bullying and dominance with structural equation modeling. School Psychology Review, 32, 401–17.Google Scholar
Ornat, Lopez A. and Greenberg, R. (1990). Sexual segregation by habitat in migratory warblers in Quintana Roo, Mexico. Auk, 107, 539–43.Google Scholar
Lott, D. F. and Minta, S. C. (1983). Random individual association and social group instability in American bison (Bison bison). Zeitschrift für Tierpsychology, 61, 153–72.CrossRefGoogle Scholar
Lovern, M. B. (2000). Behavioral ontogeny in free-ranging juvenile male and female green anoles, Anolis carolinensis, in relation to sexual selection. Journal of Herpetology, 34, 274–81.CrossRefGoogle Scholar
Low, B. S. (2000). Why Sex Matters: A Darwinian Look at Human Behavior. Princeton: Princeton University Press.Google Scholar
Lowry, L. F., Burns, J. J. and Nelson, R. R. (1987). Polar bear, Ursus maritimus, predation on belugas, Delphinapterus leucas, in the Bering and Chukchi Seas. Canadian Field-Naturalist, 101, 141–6.Google Scholar
Lowry, L. F., Frost, K. J. and Seaman, G. A. (1985). Investigations of Beluga Whales in Coastal Waters of Western and Northern Alaska – III: Food Habits. Fairbanks (Alaska): Alaska Department of Fish and Game, p. 24.Google Scholar
Lue, K. Y. and Chen, T. H. (1999). Activity, movement patterns, and home range of the yellow-margined box turtle (Cuora flavomarginata) in northern Taiwan. Journal of Herpetology, 33, 590–600.CrossRefGoogle Scholar
Luiselli, L. and Agrimi, U. (1991). Composition and variation of the diet of Vipera aspis francisciredi in relation to age and reproductive stage. Amphibia-Reptilia, 12, 137–44.CrossRefGoogle Scholar
Luiselli, L. and Angelici, F. M. (1998). Sexual size dimorphism and natural history traits are correlated with intersexual dietary divergence in royal pythons (Python regius) from the rainforests of southeastern Nigeria. Italian Journal of Zoology, 65, 183–5.CrossRefGoogle Scholar
Luiselli, L., Akani, G. C. and Capizzi, D. (1999). Is there any interspecific competition between dwarf crocodiles (Osteolaemus tetraspis) and Nile monitors (Varanus niloticus ornatus) in the swamps of central Africa? A study from southeastern Nigeria. Journal of Zoology, London, 247, 127–31.CrossRefGoogle Scholar
Luiselli, L., Capula, M. and Shine, R. (1996). Reproductive output, costs of reproduction, and ecology of the smooth snake, Coronella austriaca, in the eastern Italian Alps. Oecologia, 106, 100–10.CrossRefGoogle ScholarPubMed
Luiselli, L., Capula, M. and Shine, R. (1997). Food habits, growth rates, and reproductive biology of grass snakes, Natrix natrix (Colubridae) in the Italian Alps. Journal of Zoology, London, 241, 371–80.CrossRefGoogle Scholar
Lunn, N. J. and Arnould, J. P. Y. (1997). Maternal investment in Antarctic fur seals: evidence for equality in the sexes. Behavioral Ecology and Sociobiology, 40, 351–62.CrossRefGoogle Scholar
Lunney, D. (1995). White-footed Dunnart. In The Mammals of Australia, ed. Strahan, R.. Sydney: Reed New Holland, pp. 143–5.Google Scholar
Lunney, D. and Leary, T. (1989). Movement patterns of the white-footed dunnart, Sminthopsis leucopus (Marsupialia: Dasyuridae), in a logged, burnt forest on the south coast of New South Wales. Australian Wildlife Research, 16, 207–16.CrossRefGoogle Scholar
Lunney, D., Matthews, A. and Grigg, J. (2001). The diet of Antechinus agilis and A. swainsonii in unlogged and regenerating sites in Mumbulla State Forest, south-eastern New South Wales. Wildlife Research, 28, 459–64.CrossRefGoogle Scholar
Lunney, D., O'Connell, M. and Sanders, J. (1989). Habitat of the white-footed dunnart, Sminthopsis leucopus (Gray) (Dasyuridae: Marupialia), in a logged, burnt forest near Bega, New South Wales. Australian Journal of Ecology, 14, 335–44.CrossRefGoogle Scholar
Lyderson, C., Hammill, M. O. and Kovacs, K. M. (1994). Activity of lactating ice-breeding grey seals, Halichoerus grypus, from the Gulf of St. Lawrence, Canada. Animal Behaviour, 48, 1417–25.CrossRefGoogle Scholar
Lyle, J. M. (1983). Food and feeding habits of the lesser spotted dogfish, Scyliorhinus canicula (L.) in Isle of Man waters. Journal of Fish Biology, 23, 725–38.CrossRefGoogle Scholar
Lynch, J. F., Morton, E. S. and Voort, M. E. (1985). Habitat segregation between the sexes of wintering hooded warblers, Wilsonia citrina. Auk, 102, 714–21.Google Scholar
MacArthur, R. A., , Geist V. and Johnston, R. H. 1982. Cardiac and behavioral responses of mountain sheep to human disturbance. Journal of Wildlife Management, 46, 351–8.CrossRefGoogle Scholar
Maccoby, E. E. (1998). The Two Sexes: Growing Up Apart, Coming Together. Cambridge, MA, Harvard University Press.Google Scholar
MacFarlane, A. M. and Coulson G. (in press). Synchrony and timing of breeding influences sexual segregation in western grey and red Kangaroos (Macropus Fuliginosus and M. rufus). Journal of Zoology, London.
Macdonald, D. (2001). The New Encyclopedia of Mammals. Oxford: Oxford University Press.Google Scholar
Madsen, T. (1983). Growth rates, maturation and sexual size dimorphism in a population of grass snakes, Natrix natrix, in southern Sweden. Oikos, 40, 277–82.CrossRefGoogle Scholar
Madsen, T. (1984). Movements, home range size and habitat use of radio-tracked grass snakes (Natrix natrix) in southern Sweden. Copeia, 1984, 707–13.CrossRefGoogle Scholar
Madsen, T. and Shine, R. (1993a). Costs of reproduction in a population of European adders. Oecologia, 94, 488–95.CrossRefGoogle Scholar
Madsen, T. and Shine, R. (1993b). Phenotypic plasticity in body sizes and sexual size dimorphism in European grass snakes. Evolution, 47, 321–5.CrossRefGoogle Scholar
Madsen, T. and Shine, R. (1996). Seasonal migration of predators and prey: pythons and rats in tropical Australia. Ecology, 77, 149–56.CrossRefGoogle Scholar
Madsen, T. and Shine, R. (2000). Energy versus risk: costs of reproduction in free-ranging pythons in tropical Australia. Austral Ecology, 25, 670–5.CrossRefGoogle Scholar
Madsen, T., Shine, R., Loman, J. and Håkansson, T. (1993). Determinants of mating success in male adders, Vipera berus. Animal Behaviour, 45, 491–9.CrossRefGoogle Scholar
Magnusson, W. E. (1993). Body temperatures of field-active Amazonian savanna lizards. Journal of Herpetology, 27, 53–8.CrossRefGoogle Scholar
Magurran, A. E. (1990). The adaptive significance of schooling as an antipredator defense in fish. Annales Zoologici Fennici, 27, 51–66.Google Scholar
Magurran, A. E. and Garcia, M. (2000). Sex differences in behaviour as an indirect consequence of mating system. Journal of Fish Biology, 57, 839–57.CrossRefGoogle Scholar
Magurran, A. E., Seghers, B. H., Shaw, P. W. and Carvalho, G. R. (1994). Schooling preferences for familiar fish in the Guppy, Poecilia reticulata. Journal of Fish Biology, 45, 401–6.CrossRefGoogle Scholar
Mahmoud, I. Y. (1969). Comparative ecology of the kinosternid turtles of Oklahoma. Southwestern Naturalist, 14, 31–66.CrossRefGoogle Scholar
Main, M. B. (1994). Advantages of habitat selection and sexual segregation in mule and white-tailed deer. Ph.D. thesis, Corvallis, OR, USA: Oregon State University.Google Scholar
Main, M. B. (1998). Sexual segregation in ungulates: a reply. Journal of Mammalogy, 79, 1410–15.CrossRefGoogle Scholar
Main, M. B. and Coblentz, B. E. (1990). Sexual segregation among ungulates: a critique. Wildlife Society Bulletin, 18, 204–10.Google Scholar
Main, M. B. and Coblentz, B. E. (1996). Sexual segregation in Rocky Mountain mule deer. Journal of Wildlife Management, 60, 97–507.CrossRefGoogle Scholar
Main, M. B., Weckerly, F. W. and Bleich, V. C. (1996). Sexual segregation in ungulates: new directions for research. Journal of Mammalogy, 77, 449–61.CrossRefGoogle Scholar
Mann, J. and Barnett, H. (1999). Lethal tiger shark (Galeocerdo cuvier) attack on bottlenose dolphin (Tursiops sp.) calf: defense and reactions by the mother. Marine Mammal Science, 15, 568–75.CrossRefGoogle Scholar
Mann, J., Connor, R. C., Barre, L. M. and Heithaus, M. R. (2000a). Female reproductive success in wild bottlenose dolphins (Tursiops sp.): Life history, habitat, provisioning, and group size effects. Behavioral Ecology, 11, 210–19.CrossRefGoogle Scholar
Mann, J., Connor, R. C., Tyack, P. L. and Whitehead, H. (2000b). Cetacean Societies: Field Studies of Dolphins and Whales. Chicago: University of Chicago Press.Google Scholar
Mansergh, I. and Broome, L. (1994). The Mountain Pygmy-Possum of the Australian Alps. Kensington: New South Wales University Press.Google Scholar
Mansergh, I. and Scotts, D. (1989). Habitat continuity and social organizations of the Mountain Pygmy-possum restored by tunnel. Journal of Wildlife Management, 53, 701–7.CrossRefGoogle Scholar
Mansergh, I. and Scotts, D. (1990). Aspects of the life history and breeding biology of the Mountain Pygmy-possum, Burramys parvus, (Marsupialia: Burramyidae) in alpine Victoria. Australian Mammalogy, 13, 179–91.Google Scholar
Mansergh, I., Baxter, B., Scotts, D., Brady, T. and Jolley, D. (1990). Diet of the mountain pygmy-possum, Burramys parvus (Marsupialia: Burramyidae) and other small mammals in the alpine environment at Mt. Higginbotham, Victoria. Australian Mammalogy, 13, 167–77.Google Scholar
Marchal, C., Gerard, J.-F., Boisaubert, B. and Bideau, E. (1998). Instability and diurnal variation in size of winter groupings of field roe deer. Revue d'Ecologie (Terre Vie), 53, 59–68.Google Scholar
Marchant, S. and Higgins, P. J. (1993). Handbook of Australian, New Zealand and Antarctic Birds. vol. 2, Melbourne, Oxford University Press.Google Scholar
Marquiss, M. (1980). Habitat and diet of male and female hen harriers in Scotland in winter. British Birds, 73, 555–60.Google Scholar
Marra, P. P. (2000). The role of behavioral dominance in structuring patterns of habitat occupancy in a migrant bird during the nonbreeding season. Behavioural Ecology, 11, 299–308.CrossRefGoogle Scholar
Marra, P. P. and Holberton, R. L. (1998). Corticosterone levels as indicators of habitat quality: effects of habitat segregation in a migratory bird during the non-breeding season. Oecologia, 116, 284–92.CrossRefGoogle Scholar
Marra, P. P. and Holmes, R. T. (1997). Avian removal experiments: do they test for habitat saturation or female availability?Ecology, 78, 947–52.CrossRefGoogle Scholar
Marra, P. P. and Holmes, R. T. (2001). Consequences of dominance-mediated habitat segregation in American redstarts during the nonbreeding season. Auk, 118, 92–104.CrossRefGoogle Scholar
Marra, P. P., Hobson, K. A. and Holmes, R. T. (1998). Linking winter and summer events in a migratory bird by using stable carbon isotopes. Science, 282, 1884–6.CrossRefGoogle Scholar
Marra, P. P., Sherry, T. W. and Holmes, R. T. (1993). Territorial exclusion by a long-distance migrant warbler in Jamaica: a removal experiment with American redstarts, Setophaga ruticilla. Auk, 110, 565–72.CrossRefGoogle Scholar
Marsden, S. J. and Sullivan, M. S. (2000). Intersexual differences in feeding ecology in a male-dominated wintering pochard Aythya ferina population. Ardea, 88, 1–7.Google Scholar
Marsh, C. (1981). Time budget of Tana River red colobus. Folia Primatologica, 35, 30–50.CrossRefGoogle ScholarPubMed
Martin, A. R. and da Silva, V. M. F. (2004). River dolphins and flooded forest: seasonal habitat use and sexual segregation of botos (Inia geoffrensis) in an extreme cetacean environment. Journal of Zoology, London, 263, 295–305.CrossRefGoogle Scholar
Martin, J. and Lopez, P. (1999). Nuptial coloration and mate guarding affect escape decisions of male lizards (Psammodromus algirus). Ethology, 105, 439–47.CrossRefGoogle Scholar
Martin, R. and Handasyde, K. (1990). Population dynamics of the koala (Phascolarctos cinereus) in southeastern Australia. In Biology of the Koala, eds. Lee, A. K., Handasyde, K. A. and Sanson, G. D.. Chipping Norton, NSW: Surrey Beatty and Sons, pp. 75–84.Google Scholar
Martin, R. D. (1972). A preliminary field study of the lesser mouse lemur (Microcebus murinus J. F. Miller 1777). Zietschrift für Tierpsycholgie, 9, 43–89.Google Scholar
Maruhashi, T. (1981). Activity patterns of a troop of Japanese macaques (Macaca fuscata yakuii) on Yakushima Island, Japan. Primates, 22, 1–14.CrossRefGoogle Scholar
Matthysen, E., Grubb, T. C. and Cimprich, D. (1991). Social control of sex-specific foraging behaviour in downy woodpeckers, Picoides pubescens. Animal Behaviour, 42, 515–17.CrossRefGoogle Scholar
Mautz, W. W. (1978). Sledding on a brushy hillside: the fat cycle in deer. Wildlife Society Bulletin, 6, 88–90.Google Scholar
Smith, Maynard J. and Brown, R. L. W. (1986). Competition and body size. Theoretical Population Biology, 30, 166–79.CrossRefGoogle ScholarPubMed
McBride, A. F. and Kritzler, H. (1951). Observations on pregnancy, parturition, and post-natal behavior in the bottlenose dolphin. Journal of Mammalogy, 32, 251–66.CrossRefGoogle Scholar
McCann, T. S. (1980). Territoriality and breeding behaviour of adult male Antarctic Fur seal, Arctocephalus gazella. Journal of Zoology London, 192, 295–310.CrossRefGoogle Scholar
McCracken, G. F. (1984). Communal nursing in Mexican free-tailed bat maternity colonies. Science, 223, 1090–1.CrossRefGoogle ScholarPubMed
McCracken, G. F. and Bradbury, J. W. (1981). Social organisation and kinship in the polygynous bat Phyllostomus hastatus. Behavioural Ecology and Sociobiology, 8, 11–34.CrossRefGoogle Scholar
McCracken, G. F. and Wilkinson, G. S. (2000). Bat mating systems. In Reproductive Biology of Bats, eds. Crichton, E. G. and Krutzsch, P. H., San Diego: Academic Press, pp. 321–62.Google Scholar
McCullough, D. R. (1979). The George Reserve Deer Herd. Ann Arbor, University of Michigan Press.Google Scholar
McCullough, D. R. and McCullough, Y. (2000). Kangaroos in Outback Australia. New York: Columbia University Press.Google Scholar
McCullough, D. R., Hirth, D. R. and Newhouse, S. J. (1989). Resource partitioning between sexes in white-tailed deer. Journal of Wildlife Management, 53, 277–83.CrossRefGoogle Scholar
McFarland-Symington, M. (1988). Demography, activity budgets, and ranging patterns of black spider monkeys (Ateles belzebuth chamek) in the Manu National Park. American Journal of Primatology, 15, 45–67.CrossRefGoogle Scholar
McFarland-Symington, M. (1990). Fission-fusion social organization in Ateles and Pan. International Journal of Primatology, 11, 47–61.CrossRefGoogle Scholar
McIlhenny, E. A. (1937). Life history of the boat-tailed grackle in Louisiana. Auk, 54, 274–95.CrossRefGoogle Scholar
McKibben, J. N. and Nelson, D. R. (1986). Patterns of movement and grouping of gray reef sharks, Carcharhinus amblyrhynchos, at Enewetak, Marshall Islands. Bulletin of Marine Science, 38, 89–110.Google Scholar
McKinnon, J. (1994). Feeding habits of the dusky dolphin, Lagenorhynchus obscurus, in the coastal waters of central Peru. Fishery Bulletin, 92, 569–78.Google Scholar
McLaughlin, R. H. and O'Gower, A. K. (1971). Life history and underwater studies of a heterodont shark. Ecological Monographs, 41, 271–89.CrossRefGoogle Scholar
McLoyd, V. (1980). Verbally expressed modes of transformation in the fantasy and play of black preschool children. Child Development, 51, 1133–9.CrossRefGoogle Scholar
McNaughton, S. J. and Georgiadis, N. J. (1986) Ecology of African grazing and browsing mammals. Annual Review of Ecology and Systematics, 17, 39–65.CrossRefGoogle Scholar
McRobert, S. P. and Bradner, J. (1998). The influence of body coloration on shoaling preferences in fish. Animal Behaviour, 56, 611–15.CrossRefGoogle Scholar
Meaney, M. J. (1988). The sexual differentiation of social play. Trends in Neurosciences, 11, 54–8.CrossRefGoogle ScholarPubMed
Meaney, M. J., Stewart, J. and Beatty, W. W. (1985). Sex differences in social play: the socialization of sex roles. Advances in the Study of Behaviour, 15, 1–58.CrossRefGoogle Scholar
Menchen, F. C. and Winfield, I. (2000). Job search and sex segregation: Does sex of social contact matter?Sex Roles, 42(9–10), 847–64.CrossRefGoogle Scholar
Merchant, M. E., Shukla, S. S. and Akers, H. A. (1991). Lead concentrations in wing bones of the mottled duck. Environmental Toxicology and Chemistry, 10, 1503–7.CrossRefGoogle Scholar
Mesnick, S. L. (2001). Genetic relateness in sperm whales: evidence and cultural implications. Behavior and Brain Science, 26, 346–7.CrossRefGoogle Scholar
Metten, H. (1939a). Reproduction of the dogfish. Nature, 143, 121–2.CrossRefGoogle Scholar
Metten, H. (1939b). Studies on the reproduction of the dogfish. Philosophical Transactions of the Royal Society of London B, 230, 217–38.CrossRefGoogle Scholar
Michaud, R. (1993). Distribution estivale du béluga du Saint-Laurent; synthèse 1986 à 1992. Rapport technique canadien des sciences halieutiques et aquatiques, 1906, vi–28.Google Scholar
Michaud, R. (1999). Social organization of the St. Lawrence beluga, Delphinapterus leucas. 13th Conference on the Biology of Marine Mammals, Maui, Hawaii, Society for marine mammalogy.Google Scholar
Michelena, P., Bouquet, P. M., Dissac, A.et al. (2004). An experimental test of hypotheses explaining social segregation in dimorphic ungulates. Animal Behaviour, 68, 1371–80.CrossRefGoogle Scholar
Miles, D. B., Snell, H. L. and Snell, H. M. (2001). Intrapopulation variation in endurance of Galapagos lava lizards (Microlophus albemarlensis): evidence for an interaction between natural and sexual selection. Evolutionary Ecology Research, 3, 795–804.Google Scholar
Milinski, M. (1993). Predation risk and feeding behaviour. In Behaviour of Teleost Fishes, ed. Pitcher, T. J.. London: Chapman and Hall, pp. 285–305.CrossRefGoogle Scholar
Millar, J. S. and Hickling, G. J. (1992). The fasting endurance hypothesis revisited. Functional Ecology, 6, 496–8.Google Scholar
Minchin, D. (1987). Fishes of the Lough Hyne marine reserve. Journal of Fish Biology, 31, 343–52.CrossRefGoogle Scholar
Miquelle, D. G., Peek, J. M. and Ballenberghe, V. (1992). Sexual segregation in Alaskan moose. Wildlife Monographs, 122, 1–57.Google Scholar
Miranda, J. P. and Andrade, G. V. (2003). Seasonality in diet, perch use, and reproduction of the gecko Gonatodes humeralis from Eastern Brazilian Amazon. Journal of Herpetology, 37, 433–8.CrossRefGoogle Scholar
Mitani, J. C. and Amsler, S. (2003). Social and spatial aspects of male subgrouping in a community of wild chimpanzees. Behaviour, 140, 869–84.CrossRefGoogle Scholar
Mitani, J. C. and Watts, D. P. (2002). Why do male chimpanzees hunt and share meat?Animal Behaviour, 61, 915–24.CrossRefGoogle Scholar
Mitani, J. C., Gros-Louis, J. and Richards, A. F. (1996). Sexual dimorphism, the operational sex ratio, and the intensity of male competition in poygynous primates. American Naturalist, 147, 966–80.CrossRefGoogle Scholar
Mitani, J. C., Watts, D. P. and Lwanga, J. (2002). Ecological and social correlates of chimpanzee party size and composition. In Behavioural Diversity in Chimpanzees and Bonobos, eds., Boesch, C., Hohmann, G. and Marchant, L. F.. Cambridge: Cambridge University Press, pp. 102–11.CrossRefGoogle Scholar
Mitchell, C. (1994). Migration alliances and coalitions among adult male South American squirrel monkeys. Behaviour, 130, 169–89.CrossRefGoogle Scholar
Mitchell, P. (1990). The home ranges and social activity of koalas – a quantitative analysis. In Biology of the Koala, eds. Lee, A. K., Handasyde, K. A. and Sanson, G. D.. Chipping Norton, NSW: Surrey Beatty and Sons, pp. 171–87.Google Scholar
Miyazaki, N. and Nishiwaki, M. (1978). School structure of the striped dolphins off the Pacific coast of Japan. Scientific Report of the Whale Research Institute, 30, 65–115.Google Scholar
Moll, D. (1990). Population sizes and foraging ecology in a tropical freshwater stream turtle community. Journal of Herpetology, 24, 48–53.CrossRefGoogle Scholar
Moll, E. O. and Legler, J. M. (1971). The life history of a neotropical slider turtle, Pseudemys scripta (Schoepff) in Panama. Bulletin of the Los Angeles County Museum of Natural History, Science, 11, 1–102.Google Scholar
Moore, C. L. (1985). Development of mammalian sexual behavior. In The Comparative Development of Adaptative Skills: Evolutionary Implications, ed. Gollin, E. S.. Hillsdale: Lawrence Erlbaum Associate, pp. 19–56.Google Scholar
Morgantini, L. E. and Hudson, R. J. (1981). Sex differential in use of the physical environment by bighorn sheep (Ovis canadensis). Canadian Field-Naturalist, 95, 69–74.Google Scholar
Mori, A. and Watanabe, K. (2003). Life history of male Japanese monkeys living on Koshima Islet. Primates, 44, 119–26.Google ScholarPubMed
Moritz, C. (1993). The origin and evolution of parthenogenesis in the Heteronotia binoei complex: synthesis. Genetica, 90, 269–80.CrossRefGoogle Scholar
Morreale, S. J., Gibbons, J. W. and Congdon, J. D. (1984). Significance of activity and movement in the yellow-bellied slider turtle (Pseudemys scripta). Canadian Journal of Zoology, 62, 1038–2.CrossRefGoogle Scholar
Morrison, M. L. and With, K. A. (1987). Interseasonal and intersexual resource partitioning in hairy and white-headed woodpeckers. Auk, 104, 225–33.Google Scholar
Morrissey, J. F. and Gruber, S. H. (1993). Habitat selection by juvenile lemon sharks, Negaprion brevirostris. Environmental Biology of Fishes, 38, 311–19.CrossRefGoogle Scholar
Morton, E. S. (1990). Habitat segregation by sex in the hooded warbler: experiments on proximate causation and discussion of its evolution. American Naturalist, 135, 319–33.CrossRefGoogle Scholar
Morton, E. S., Voort, M. E. and Greenberg, R. (1993). How a warbler chooses its habitat: field support for laboratory experiments. Animal Behaviour, 46, 47–53.CrossRefGoogle Scholar
Morton, S. (1978). An ecological study of Sminthopsis crassicaudata (Marsupialia: Dasyuridae) III. Reproduction and life history. Australian Wildlife Research, 5, 183–211.CrossRefGoogle Scholar
Morton, E. S., Lynch, J. F., Young, K. and Mehlhop, P. (1987). Do male hooded warblers exclude females from non-breeding territories in tropical forest?Auk, 104, 133–5.CrossRefGoogle Scholar
Moseby, K. E. and O'Donnell, E. (2003). Reintroduction of the greater bilby, Macrotis lagotis (Reid) (Marsupialia: Thylacomyidae), to northern South Australia: survival, ecology and notes on reintroduction protocols. Wildlife Research, 30, 15–27.CrossRefGoogle Scholar
Moskowitz, D. S., Suh, E. J. and Desaulniers, J. (1994). Situational influences on gender differences in agency and communion. Journal of Personality and Social Psychology, 66(4), 753–61.CrossRefGoogle ScholarPubMed
Moss, R. (1983). Gut size, body weight, and digestion of winter foods by grouse and ptarmigan. Condor, 85, 185–93.CrossRefGoogle Scholar
Mougin, J. L. (1968). Etude écologique de quatre espèces de pétrels antarctique. Oiseau Revue Francophone Ornithology, 38, 1–52.Google Scholar
Mougin, J. L. (1975). Ecologie comparée des Procellariidae Antarctiques et sub-Antarctiques. CNFRA, 36, 1–195.Google Scholar
Mueller, H. C. (1990). The evolution of reversed sexual dimorphism in size in monogamous species of birds. Biological Reviews of the Cambridge Philosophical Society, 65, 553–85.CrossRefGoogle Scholar
Müller, A. E. and Thalmann, U. (2000). Origin and evolution of primate social organisation: a reconstruction. Biological Reviews of the Cambridge Philosophical Society, 75, 405–35.CrossRefGoogle ScholarPubMed
Murphy, M. T., Pierce, A., Shoen, J.et al. (2001). Population structure and habitat use by overwintering neotropical migrants on a remote oceanic island. Biological Conservation, 102, 333–45.CrossRefGoogle Scholar
Murray, M. G. and Illius, A. W. (2000). Vegetation modification and resource competition in grazing ungulates. Oikos, 89, 501–8.CrossRefGoogle Scholar
Murray, T. E., Bartle, J. A., Kalish, S. R. and Taylor, P. R. (1993). Incidental capture of seabirds by Japanese southern bluefin tuna longline vessels in New Zealand waters, 1988–1992. Bird Conservation International, 3, 181–210.CrossRefGoogle Scholar
Mushinsky, H. R., Hebrard, J. J. and Vodopich, D. S. (1982). Ontogeny of water snake foraging ecology. Ecology, 63, 1624–9.CrossRefGoogle Scholar
Myers, J. P. (1981). A test of three hypotheses for latitudinal segregation of the sexes in wintering birds. Canadian Journal of Zoology, 59, 1527–34.CrossRefGoogle Scholar
Myrberg, A. A. and Gruber, S. H. (1974). The behaviour of the bonnethead, Sphyrna tiburo. Copeia, 1974, 358–74.CrossRefGoogle Scholar
Myres, B. C. and Eells, M. M. (1968). Thermal aggregation in Boa constrictor. Herpetologica, 24, 61–6.Google Scholar
Mysterud, A. (2000). The relationship between ecological segregation and sexual body size dimorphism in large herbivores. Oecologia, 124, 40–54.CrossRefGoogle ScholarPubMed
Nakagawa, N. (1989). Activity budget and diet of patas monkeys in Kala Maloue National Park, Cameroon. Primates, 30, 27–34.CrossRefGoogle Scholar
Naulleau, G. (1966). Etude complementaire de l'activitie de Vipera aspis dans la nature. Vie et Milieu, 17, 461–509.Google Scholar
Nel, D. C., Ryan, P. G., Nel, J. L.et al. (2002). Foraging interactions between wandering albatrosses Diomedea exulans breeding on Marion Island and long-line fisheries in the southern Indian Ocean. Ibis, 144 (on-line), E141–E154.CrossRefGoogle Scholar
Nelson, J. E. (1965). Behaviour of Australian Pteropodidae (Megachiroptera). Animal Behaviour, 13, 544–57.CrossRefGoogle Scholar
Nelson, M. E. and Mech, L. D. (1981). Deer social organization and wolf predation in northeastern Minnesota. Wildlife Monographs, 77, 1–53.Google Scholar
Nelson, M. E. and Mech, L. D. (1984). Home-range formation and dispersal of deer in Northeastern Minnesota. Journal of Mammalogy, 65, 567–75.CrossRefGoogle Scholar
Neuhaus, P. and Ruckstuhl, K. E. (2002a). The link between sexual dimorphism, activity budgets, and group cohesion: the case of the plains zebra (Equus burchelli). Canadian Journal of Zoology, 80, 1437–41.CrossRefGoogle Scholar
Neuhaus, P. and Ruckstuhl, K. E. (2002b). Foraging behaviour in Alpine ibex (Capra ibex): consequences of reproductive status, body size, age and sex. Ecology, Ethology and Evolution, 14, 373–81.CrossRefGoogle Scholar
Newsome, A. E. (1980). Differences in the diets of male and female red kangaroos in central Australia. African Journal of Ecology, 18, 27–31.CrossRefGoogle Scholar
Newton, I. (1986). The Sparrowhawk. Calton: T and AD Poyser.Google Scholar
Newton, I. (1998). Population Limitation in Birds. San Diego: Academic Press.Google Scholar
Newton-Fisher, N., Reynolds, V. and Plumtre, A. J. (2000). Food supply and chimpanzee (Pan troglodytes schweinfurthii) in the Budongo Forest Reserve, Uganda. International Journal of Primatology, 21, 613–28.CrossRefGoogle Scholar
Nicholls, D. G., Murray, M. D., Butcher, E. and Moors, P. (1997). Weather systems determine the non-breeding distribution of wandering albatrosses over southern oceans. Emu, 95, 240–4.Google Scholar
Nichols, J. D. and Aramis, G. M. (1980). Sex-specific differences in winter distribution patterns of canvasbacks. Condor, 82, 406–16.CrossRefGoogle Scholar
Nievergelt, B. (1967). Die zusammensetzung der gruppen beim alpensteinbock. Zeitschrift für Säugetierkunde, 32, 129–44.Google Scholar
Nievergelt, B. (1981). Ibexes in an African Environment. Berlin: Springer Verlag.CrossRefGoogle Scholar
Nikaido, M., Rooney, A. P. and Okada, N. (1999). Phylogenetic relationships among cetartiodactyls based on insertions of short and long interspersed elements: hippopotamuses are the closest extant relatives of whales. Proceedings of the National Academy of Science, 96, 10261–6.CrossRefGoogle Scholar
Nilsson, L. (1970). Food seeking activity of south Swedish diving ducks in the non-breeding season. Oikos, 21, 145–54.CrossRefGoogle Scholar
Nisbet, I. C. T. and Medway, L. (1972). Dispersion, population ecology and migration of eastern great reed warblers Acrocephalus orientalis wintering in Malaysia. Ibis, 114, 451–94.CrossRefGoogle Scholar
Nisbet, I. C. T., Montoya, J. P., Burger, J. and Hatch, J. J. (2002). Use of stable isotopes to investigate individual differences in diets and mercury exposures among common terns Sterna hirundo in breeding and wintering grounds. Marine Ecology Progress Series, 242, 267–74.CrossRefGoogle Scholar
Nishida, T. (1968). The social group of wild chimpanzees in the Mahale Mountains. Primates, 9, 167–224.CrossRefGoogle Scholar
Nishida, T. and Hosaka, K. (1996). Coalition strategies among adult male chimpanzees of the Mahale Mountains, Tanzania. In Great Ape Societies, eds. McGrew, W. A., Marchant, L. A., and Nishida, T.. Cambridge: Cambridge University Press, pp. 114–34.CrossRefGoogle Scholar
Nogueira, C., Sawaya, R. J. and Martins, M. (2003). Ecology of the pitviper, Bothrops moojeni, in the Brazilian Cerrado. Journal of Herpetology, 37, 653–9.CrossRefGoogle Scholar
Norbury, G. L., Coulson, G. M. and Walters, B. L. (1988). Aspects of the demography of the western grey kangaroo, Macropus fuliginosus melanops, in semiarid north-west Victoria. Australian Wildlife Research, 15, 257–66.CrossRefGoogle Scholar
Norris, K. S. (1994). Comparative view of cetacean social ecology, culture, and evolution. In The Hawaiian Spinner Dolphin, eds. Norris, K. S., Würsig, B., Well, R. S. and Würsig, M.. Berkeley: University of California Press, pp. 301–44.Google Scholar
Norris, K. S. and Dohl, T. P. (1980). The structure and functions of cetacean schools. In Cetacean Behavior: Mechanisms and Functions, ed. Herman, L. M.. New York: John Wiley and Sons, pp. 211–59.Google Scholar
Norris, K. S. and Schilt, C. R. (1988). Cooperative societies in three-dimensional space: on the origins of aggregations, flocks, and schools, with special reference to dolphins and fish. Ethology and Sociobiology, 9, 149–79.CrossRefGoogle Scholar
Northcutt, R. G. (1977). Elasmobranch central nervous system organization and its possible evolutionary significance. American Zoologist, 17, 411–29.CrossRefGoogle Scholar
Novacek, M. J. (1992). Mammalian phylogeny: shaking the tree. Nature, London, 356, 121–5.CrossRefGoogle ScholarPubMed
Nowak, R. M. (1991). Walker's Mammals of the World, 5th edn. Baltimore: John Hopkins University Press.Google Scholar
Nunn, C. (1999). The number of males in primate social groups: a comparative test of the socioecological model. Behavioral Ecology and Sociobiology, 46, 1–13.CrossRefGoogle Scholar
O'Donnell, C. F. J. and Sedgeley, J. A. (1999). Use of roosts by the long-tailed bat, Chalinolobus tuberculatus, in temperate rain forest in New Zealand. Journal of Mammalogy, 80, 913–23.CrossRefGoogle Scholar
Oakwood, M. (2000). Reproduction and demography of the northern quoll, Dasyurus hallucatus, in the lowland savanna of northern Australia. Australian Journal of Zoology, 48, 519–39.CrossRefGoogle Scholar
Oakwood, M. (2002). Spatial and social organization of a carnivorous marsupial Dasyurus hallucatus (Marsupialia: Dasyuridae). Journal of Zoology, 257, 237–48.CrossRefGoogle Scholar
Ohguchi, O. (1978). Experiments on the selection against colour oddity of water fleas by three-spined stickelbacks. Zeitschrift fur Tierpsychologie, 47, 254–67.CrossRefGoogle Scholar
Olesiuk, P. F., Bigg, M. A. and Ellis, G. M. (1990). Life history and population dynamics of resident killer whales (Orcinus orca) in the coastal waters of British Columbia and Washington state. Report of the International Whaling Commission, Special Issue, 12, 209–405.
Oli, M. K. (1996). Seasonal patterns in habitat use of blue sheep Pseudois nayaur (Artiodactyla, Bovidae) in Nepal. Mammalia, 60, 187–93.CrossRefGoogle Scholar
Oliver, W. R. B. (1955). New Zealand Birds. Wellington: A. H. and A. W. Reed.Google Scholar
Olsen, A. M. (1954). The biology, migration, and growth rate of the school shark, Galeorhinus australis (Macleay) (Carcharhinidae) in southeastern Australian waters. Australian Journal of Marine and Freshwater Research, 5, 353–410.Google Scholar
Ong, P. S. (1994). The Social Organization of the Common Ringtail Possum, Pseudocheirus peregrinus. Unpublished Ph.D. thesis, Monash University.Google Scholar
Ordway, L. L. and Krausman, P. R. (1986). Habitat use by mule deer. Journal of Wildlife Management, 50, 677–83.CrossRefGoogle Scholar
Ortega, J. and Arita, H. T. (2000). Defence of females by dominant males of Artibeus jamaicensis (Chiroptera: Phyllostomidae). Ethology, 106, 395–407.CrossRefGoogle Scholar
Ortega, J. and Arita, H. T. (2002). Subordinate males in harem groups of Jamaican fruit-eating bats (Artibeus jamaicensis): Satellites or sneaks?Ethology, 108, 1077–91.CrossRefGoogle Scholar
Oswald, C., Fonken, P., Atkinson, D. and Palladino, M. (1993). Lactational water-balance and recycling in white-footed mice, red-backed voles, and gerbils. Journal of Mammalogy, 74, 963–70.CrossRefGoogle Scholar
Owen, M. and Black, J. M. (1990). Waterfowl Ecology. Glasgow and London: Blackie.Google Scholar
Owen, M. and Black, J. M. (1991). The importance of migration mortality in non-passerine birds. In Bird Population Studies. Relevance to Conservation and Management, eds. Perrins, C. M., Lebreton, J.-D. and Hirons, G. J. M.. Oxford University Press.Google Scholar
Owen, M. and Dix, M. (1986). Sex ratios in some common British wintering ducks. Wildfowl, 37, 104–12.Google Scholar
Owens, I. P. F. and Bennet, P. M. (1994). Mortality costs of parental care and sexual dimorphism in birds. Proceedings of the Royal Society of London B, 257, 1–8.CrossRefGoogle Scholar
Owen-Smith, N. (1993). Comparative mortality rates of male and female kudus: the costs of sexual size dimorphism. Journal of Animal Ecology, 62, 428–40.CrossRefGoogle Scholar
Owen-Smith, R. N. (1988). Megaherbivores. The Influence of Very Large Body Size on Ecology. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Ozoga, J. J. (1968). Variations in microclimate in a conifer swamp deeryard in Northern Michigan. Journal of Wildlife Management, 32, 574–81.CrossRefGoogle Scholar
Ozoga, J. J. and , Verme L. J. (1970). Winter feeding patterns of penned white-tailed deer. Journal of Wildlife Management, 34, 431–9.CrossRefGoogle Scholar
Palombit, R. A. (1999). Infanticide and the evolution of pair bonds in nonhuman primates. Evolutionary Anthropology, 7, 117–28.3.0.CO;2-O>CrossRefGoogle Scholar
Palombit, R. A., Cheney, D. L., Fischer, J. et al. (2000). Male infanticide and defense of infants in chacma baboons. In Infanticide by Males and Its Implications, eds. Schaik, C. P. and Janson, C. H.. Cambridge: Cambridge University Press, pp. 123–52.
Papastavrou, V., Smith, S. C. and Whitehead, H. (1989). Diving behaviour of the sperm whale, Physeter macrocephalus, off the Galapagos Islands. Canadian Journal of Zoology, 67, 839–46.CrossRefGoogle Scholar
Parke, R. D. and Suomi, S. J. (1981). Adult male infant relationships: human and nonhuman primate evidence. In Behavioral Development, eds. Immelman, K., Barlow, G. W., Petronovitch, L., and Main, M.. New York: Cambridge University Press, pp. 700–25.Google Scholar
Parker, K. L., Gillingham, M. P., Hanley, T. A. and Robbins, C. T. (1999). Energy and protein balance of free-ranging black-tailed deer in a natural forest environment. Wildlife Monographs, 143, 1–48.Google Scholar
Parker, W. S. and Brown, W. S. (1980). Comparative ecology of two colubrid snakes, Masticophis t. taeniatus and Pituophis melanoleucus deserticola, in northern Utah. Milwaukee Public Museum Publications in Biology and Geology, 7, 1–104.Google Scholar
Parmelee, D. F., Parmelee, J. M. and Fuller, M. R. (1985). Ornithological investigations at Palmer Station: the first long-distance tracking of seabirds by satellites. Antarctic Journal of the United States, 162–3.Google Scholar
Parmelee, J. R. and Guyer, C. (1995). Sexual differences in foraging behavior of an anoline lizard, Norops humilis. Journal of Herpetology, 29, 619–21.CrossRefGoogle Scholar
Parr, A. E. (1931). Sex dimorphism and schooling behavior among fishes. American Naturalist, 65, 173–80.CrossRefGoogle Scholar
Parrish, J. D. and Sherry, T. W. (1994). Sexual habitat segregation by American redstarts wintering in Jamaica: importance of resource seasonality. Auk, 111, 38–49.CrossRefGoogle Scholar
Parsons, K. N., Jones, G., Davidson-Watts, I. and Greenaway, F. (2003). Swarming of bats at underground sites in Britain – implications for conservation. Biological Conservation, 111, 63–70.CrossRefGoogle Scholar
Pasinelli, G. (2000). Sexual dimorphism and foraging niche partitioning in the middle spotted woodpecker Dendrocopos medius. Ibis, 2000, 635–44.Google Scholar
Patterson, D. L. and Fraser, W. R. (2003). Satellite tracking southern giant petrels at Palmer station, Antarctica. Feature Articles, Microwave Telemetry, Inc., 8, 3–4.Google Scholar
Patterson, I. A. P., Reid, R. J., Wilson, B.et al. (1998). Evidence for infanticide in bottlenose dolphins: an explanation for violent interactions with harbour porpoises?Proceedings of the Royal Society of London, Series B, 265, 1167–70.CrossRefGoogle ScholarPubMed
Pearson, D., Shine, R. and How, R. (2002a). Sex-specific niche partitioning and sexual size dimorphism in Australian pythons (Morelia spilota imbricata). Biological Journal of the Linnean Society, 77, 113–25.CrossRefGoogle Scholar
Pearson, D., Shine, R. and Williams, A. (2002b). Geographic variation in sexual size dimorphism within a single snake species (Morelia spilota, Pythonidae). Oecologia, 131, 418–26.CrossRefGoogle Scholar
Pearson, D., Shine, R. and Williams, A. (2003). Thermal biology of large snakes in cool climates: a radio-telemetric study of carpet pythons (Morelia spilota imbricata) in south-western Australia. Journal of Thermal Biology, 28, 117–31.CrossRefGoogle Scholar
Pellegrini, A. D. (2003). Perceptions and possible functions of play and real fighting in early adolescence. Child Development, 74, 1459–70.CrossRefGoogle Scholar
Pellegrini, A. D. and Bartini, M. (2001). Dominance in early adolescent boys: affiliative and aggressive dimensions and possible functions. Merrill-Palmer Quarterly, 47, 142–63.CrossRefGoogle Scholar
Pellegrini, A. D. and Long, J. D. (2003). A sexual selection theory longitudinal analysis of sexual segregation and integration in early adolescence. Journal of Experimental Child Psychology, 85, 257–78.CrossRefGoogle ScholarPubMed
Pellegrini, A. D. and Perlmutter, J. C. (1989). Classroom contextual effects on children's play. Developmental Psychology, 25, 289–96.CrossRefGoogle Scholar
Pellegrini, A. D. and Smith, P. K. (1998). Physical activity play: the nature and function of a neglected aspect of play. Child Development, 69, 577–98.CrossRefGoogle Scholar
Pellegrini, A. D., Blatchford, P., Kato, K. and Baines, E. (2003). A short-time longitudinal study of children's playground games in primary school: implications for adjustment to school and social adjustment in the USA and the UK. Social Development, 13, 107–23.CrossRefGoogle Scholar
Pellegrini, A. D., Horvat, M. and Huberty, P. D. (1998). The relative cost of children's physical activity play. Animal Behaviour, 55, 1053–106.CrossRefGoogle ScholarPubMed
Pellew, R. A. (1983). The impacts of elephant, giraffe and fire upon the Acacia tortilis woodlands of the Serengeti. African Journal of Ecology, 21, 41–74.CrossRefGoogle Scholar
Pennycuick, C. J. (1987). Flight of seabirds. In Seabirds: Feeding Ecology and Role in Marine Ecosystems, ed. Croxall, J. P.. Cambridge: Cambridge University Press, pp. 43–62.Google Scholar
Pepper, J. M., Mitani, J. C. and Watts, D. P. (1999). General gregariousness and specific partner preference among wild chimpanzees. International Journal of Primatology, 20, 613–32.CrossRefGoogle Scholar
Pereira, M. E. (1988). Effects of age and sex on intra-group spacing in juvenile savanna baboons, Papio cynocephalus cynocephalus. Animal Behaviour, 36, 184–204.CrossRefGoogle Scholar
Pereira, M. E. (1989). Agonistic interactions of juvenile savanna baboons. II. Agonistic support and rank acquisition. Ethology, 80, 152–71.CrossRefGoogle Scholar
Pereira, M. E. and Fairbanks, L. A. (2002). Foreword 2002: family, friends, and the evolution of childood. In Juvenile Primates, eds. Pereira, M. E. and Fairbanks, L. A., 2nd edn. Chicago: University of Chicago Press, pp. vii–xxiv.Google Scholar
Pereira, M. E. and Leigh, S. R. (2003). Modes of primate development. In Primate Life Histories and Socioecology, eds. Kappeler, P. M. and Pereira, M. E.. Chicago: University of Chicago Press, pp. 149–76.Google Scholar
Pérez-Barbería, F. J. and Gordon, I. J. (1998a). The influence of sexual dimorphism in body size and mouth morphology on diet selection and sexual segregation in cervids. Acta Veterinaria Hungarica, 46, 357–67.Google Scholar
Pérez-Barbería, F. J. and Gordon, I. J. (1998b). Factors affecting food comminution during chewing in ruminants: a review. Biological Journal of the Linnean Society, 63, 233–56.CrossRefGoogle Scholar
Pérez-Barbería, F. J. and Gordon, I. J. (1999). Body size dimorphism and sexual segregation in polygynous ungulates: an experimental test with Soay sheep. Oecologia, 120, 258–67.Google ScholarPubMed
Pérez-Barbería, F. J., Gordon, I. J. and Pagel, M. (2002). The origins of sexual dimorphism in body size in ungulates. Evolution, 56, 1276–85.CrossRefGoogle ScholarPubMed
Perez-Mellado, V. and Riva, I. (1993). Sexual size dimorphism and ecology: the case of a tropical lizard, Tropidurus melanopleurus (Sauria, Tropiduridae). Copeia, 1993, 969–76.CrossRefGoogle Scholar
Pernetta, J. C. (1977). Observations on the habits and morphology of the snake Laticauda colubrina in Fiji. Canadian Journal of Zoology, 55, 1612–19.CrossRefGoogle Scholar
Perret, M. (1998). Energetic advantage of nest sharing in a solitary primate, the lesser mouse lemur (Microcebus murinus). Folia Primatologica, 59, 1–25.Google Scholar
Perrin, W. F. and Reilly, S. B. (1984). Reproductive parameters of dolphins and small whales of the family Delphinidae. Report of the International Whaling Commission, Special Issue, 6, 97–134.Google Scholar
Perrin, W. F., Wilson, C. E. and Archer II, F. I. (1994). Striped dolphin – Stenella coeruleoalba (Meyen, 1833). In Handbook of Marine Mammals: The First Book of Dolphins, vol. 5, eds. Ridgway, S. H. and Harrison, R.. London: Academic Press, pp. 129–60.Google Scholar
Perry, G. (1996). The evolution of sexual dimorphism in the lizard Anolis polylepis (Iguania): evidence from intraspecific variation in foraging behavior and diet. Canadian Journal of Zoology, 74, 1238–45.CrossRefGoogle Scholar
Perry, G. and Brandeis, M. (1992). Variation in stomach contents of the gecko Ptyodactylus hasselquistii guttatus in relation to sex, season, and locality. Amphibia-Reptilia, 13, 275–82.CrossRefGoogle Scholar
Perry, G. and Garland, T. (2002). Lizard home ranges revisited: effects of sex, body size, diet, habitat, and phylogeny. Ecology, 83, 1870–85.CrossRefGoogle Scholar
Peter, H.-U., Kaiser, M. and Gebauer, A. (1988). Investigations on birds and seals at King George Island. Geodätische und geophysikalische Veröffentlichungen, 1, 5–80.Google Scholar
Peters, W. D. and Grubb, T. C. (1983). An experimental analysis of sex-specific foraging in the downy woodpecker Picoides pubescens. Ecology, 64, 1437–43.CrossRefGoogle Scholar
Peterson, C. R., Gibson, A. R. and Dorcas, M. E. (1993). Snake thermal ecology: the causes and consequences of body-temperature variation. In Snakes: Ecology and Behavior, eds. Seigel, R. A. and Collins, J. T.. New York: McGraw-Hill, pp. 241–314.Google Scholar
Peterson, R. L. (1965). A review of the bats of the genus Ametrida, family Phyllostomidae. Life Science Contribution, Royal Ontario Museum, 65, 1–13.Google Scholar
Pettit, K. E., Bishop, C. A. and Brooks, R. J. (1995). Home range and movements of the common snapping turtle, Chelydra serpentina serpentina, in a coastal wetland of Hamilton Harbor, Lake Ontario, Canada. Canadian Field-Naturalist, 109, 192–200.Google Scholar
Peuhkuri, N. (1999). Size-assorted fish shoals and the majority's choice. Behavioural Ecology and Sociobiology, 46, 307–12.CrossRefGoogle Scholar
Pfeffer, P. (1967). Le mouflon de Corse (Ovis ammon musimon Schreber, 1782); position systématique, écologie, et éthologie comparées. Mammalia, 31, 1–262.Google Scholar
Phelps, T. W. (1978). Seasonal movement of the snakes Coronella austriaca, Vipera berus, and Natrix natrix in southern England. British Journal of Herpetology, 5, 775–81.Google Scholar
Phillips, J. A. (1995). Movement patterns and density of Varanus albigularis. Journal of Herpetology, 29, 407–16.CrossRefGoogle Scholar
Phillips, R. A., Silk, J. R. D., Croxall, J. P., Afanasyev, V. and Briggs, D. R. (2003). Accuracy of geolocation estimates for flying seabirds. Marine Ecology Progress Series, 266, 265–72.CrossRefGoogle Scholar
Phillips, R. A., Silk, J. R. D., Phalan, B., Catry, P. and Croxall, J. P. (2004). Seasonal sexual segregation in the two Thalassarche albatross species: competitive exclusion, reproductive role specialization or foraging niche divergence?Proceeding of the Royal Society of London, 271, 1283–91.CrossRefGoogle ScholarPubMed
Pianka, E. R. and Vitt, L. J. (2003). Lizards: Windows to the Evolution of Diversity. Berkeley, CA: University of California Press.Google Scholar
Pickering, S. P. C. (1989). Attendance patterns and behaviour in relation to experience and pair-bond formation in wandering albatrosses Diomedea exulans at South Georgia. Ibis, 131, 183–95.CrossRefGoogle Scholar
Pickering, S. P. C. and Berrow, S. D. (2002). Courtship behavior of the wandering albatross Diomedea exulans at Bird Island, South Georgia. Marine Ornithology, 29, 29–37.Google Scholar
Pierce, B. M., Bleich, V. C. and Bowyer, R. T. (2000). Prey selection by mountain lions and coyotes: effects of hunting style, body size, and reproductive status. Journal of Mammalogy, 81, 462–72.2.0.CO;2>CrossRefGoogle Scholar
Pierce, G. J. and Boyle, P. R. (1991). A review of methods for diet analysis in piscivorous marine mammals. Oceanography and Marine Biology, 29, 409–86.Google Scholar
Pierotti, R. and Annett, C. A. (1991). Diet choice in the herring gull: constraints imposed by reproductive and ecological factors. Ecology, 72, 319–28.CrossRefGoogle Scholar
Pippard, L. (1985). Status of the St. Lawrence River population of beluga, Delphinapterus leucas. Canadian Field-Naturalist, 99, 438–50.Google Scholar
Pitcher, E. G. and Schultz, L. H. (1983). Boys and Girls at Play: The Development of Sex Roles. South Hadley, MA: Bergin and Garvey.Google Scholar
Pitcher, T. J. and Parrish, J. K. (1993). Functions of shoaling behaviour in teleosts. In Behaviour of Teleost Fishes, ed. Pitcher, T. J.. Chapman and Hall, London, pp. 363–439.CrossRefGoogle Scholar
Pitcher, T. J., Magurran, A. E. and Edwards, J. I. (1985). Schooling mackerel and herring choose neighbours of similar size. Marine Biology, 86, 319–22.CrossRefGoogle Scholar
Pitman, R. L., Ballance, L. T., Mesnick, S. I. and Chivers, S. J. (2001). Killer whale predation on sperm whales: observations and implications. Marine Mammal Science, 17, 494–507CrossRefGoogle Scholar
Plummer, M. V. (1977). Activity, habitat, and population structure in the turtle Trionyx muticus. Copeia, 1977, 431–440.CrossRefGoogle Scholar
Plummer, M. V. (1981a). Communal nesting of Opheodrys aestivus in the laboratory. Copeia, 1981, 243–6.CrossRefGoogle Scholar
Plummer, M. V. (1981b). Habitat utilization, diet and movements of a temperate arboreal snake (Opheodrys aestivus). Journal of Herpetology, 15, 425–32.CrossRefGoogle Scholar
Plummer, M. V. and Farrar, D. B. (1981). Sexual dietary differences in a population of Trionyx muticus. Journal of Herpetology, 15, 175–9.CrossRefGoogle Scholar
Pluto, T. G. and Bellis, E. D. (1986). Habitat utilization by the turtle, Graptemys geographica, along a river. Journal of Herpetology, 20, 22–31.CrossRefGoogle Scholar
Pluto, T. G. and Bellis, E. D. (1988). Seasonal and annual movements of riverine map turtles, Graptemys geographica. Journal of Herpetology, 22, 152–8.CrossRefGoogle Scholar
Poindron, P., Lévy, F. and Krehbiel, D. (1988). Genital, olfactory, and endocrine interactions in the development of maternal behaviour in the parturient ewe. Psychoneuroendocrinology, 13, 99–125.CrossRefGoogle ScholarPubMed
Polis, G. A. (1984). Age structure component of niche width and intra-specific resource partitioning: can age groups function as ecological species?American Naturalist, 123, 541–64.CrossRefGoogle Scholar
Pope, T. (1990). The reproductive consequences of male cooperation in the red howler monkey: paternity exclusion in multi-male and single-male troops using genetic markers. Behavioral Ecology and Sociobiology, 27, 439–46.CrossRefGoogle Scholar
Pope, T. (2000a). The evolution of male philopatry in neotropical monkeys. In Primate Males, ed. Kappeler, P.. Cambridge: Cambridge University Press, pp. 219–35.Google Scholar
Pope, T. (2000b). Reproductive success increases with degree of kinship in cooperative coalitions of female red howler monkeys. Behavioral Ecology and Sociobiology, 48, 253–67.CrossRefGoogle Scholar
Porter, D. A. (1990). Feeding ecology of Graptemys caglei Haynes and McKown in the Guadalupe River, Dewitt County, Texas. M.Sc. thesis, Canyon, TX: West Texas State University.Google Scholar
Porter, R. H., Désiré, L., Bon, R., Orgeur, P. (2001). The role of familiarity in the development of social recognition by lambs. Behaviour, 134, 207–19.CrossRefGoogle Scholar
Porter, W. F. (1994). Family Meleagridae (Turkeys). In Handbook of the Birds of the World, vol. 2, eds. Hoyo, J. del, Elliot, A. and Sargatal, J.. Barcelona: Lynx Edicions.Google Scholar
Post, D., Hausfater, G. and McKusky, S. (1981). Feeding behavior of yellow baboons (Papio cynocephalus): relationship to age, gender, and dominance rank. Folia Primatologica, 34, 170–95.CrossRefGoogle Scholar
Post, D. M., Armbrust, T. S., Horne, E. A. and Goheen, J. R. (2001). Sexual segregation results in differences in content and quality of bison (Bos bison) diets. Journal of Mammalogy, 82, 407–13.2.0.CO;2>CrossRefGoogle Scholar
Post, E., Langvatn, R., Forchhammer, M. C. and Stenseth, N. C. (1999). Environmental variation shapes sexual dimorphism in red deer. Proceedings of the National Academy of Sciences, 96, 4467–71.CrossRefGoogle ScholarPubMed
Pough, F. H. (1973). Lizard energetics and diet. Ecology, 54, 837–44.CrossRefGoogle Scholar
Pough, F. H. (1980). The advantages of ectothermy for tetrapods. American Naturalist, 115, 92–112.CrossRefGoogle Scholar
Pough, F. H. and Groves, J. D. (1983). Specialization in the body form and food habits of snakes. American Zoologist, 23, 443–54.CrossRefGoogle Scholar
Pounds, J. A. and Jackson, J. F. (1983). Utilization of perch sites by sex and size classes of Sceloporus undulatus undulatus. Journal of Herpetology, 17, 287–9.CrossRefGoogle Scholar
Powell, G. L. and Russell, A. P. (1984). The diet of the eastern short-horned lizard (Phrynosoma douglassi brevirostre) in Alberta and its relationship to sexual size dimorphism. Canadian Journal of Zoology, 62, 428–40.CrossRefGoogle Scholar
Power, T. G. (2000). Play and Exploration in Children and Animals. Mahwah, NJ: Erlbaum.Google Scholar
Pratt, H. L. (1979). Reproduction in the blue shark, Prionace glauca. U.S. Fishery Bulletin, 77, 445–70.Google Scholar
Pratt, H. L. and Carrier, J. C. (2001). A review of elasmobranch reproductive behaviour with a case study on the nurse shark, Ginglymostoma cirratum. Environmental Biology of Fishes, 60, 157–88.CrossRefGoogle Scholar
Preest, M. R. (1994). Sexual size dimorphism and feeding energetics in Anolis carolinensis: why do females take smaller prey than males?Journal of Herpetology, 28, 292–8.CrossRefGoogle Scholar
Preston, C. R. (1990). Distribution of raptor foraging in relation to prey biomass and habitat structure. Condor, 92, 107–12.CrossRefGoogle Scholar
Prestt, I. (1971). An ecological study of the viper Vipera berus in southern Britain. Journal of Zoology, London, 164, 373–418.CrossRefGoogle Scholar
Prince, P. A. and Francis, M. D. (1984). Activity budgets of foraging grey-headed albatrosses. Condor, 86, 297–300.CrossRefGoogle Scholar
Prince, P. A. and Morgan, R. A. (1987). Diet and feeding ecology of Procellariiformes. In Seabirds: Feeding Ecology and Role in Marine Ecosystems, ed. Croxall, J. P.. Cambridge: Cambridge University Press, pp. 135–71.Google Scholar
Prince, P. A. and Walton, D. W. H. (1984). Automated measurement of feed size and feeding frequency in albatrosses. Journal of Applied Ecology, 21, 789–94.CrossRefGoogle Scholar
Prince, P. A., Croxall, J. P., Trathan, P. N. and Wood, A. G. (1998). The pelagic distribution of South Georgia albatrosses and their relationships with fisheries. In Albatross Biology and Conservation, eds. Robertson, G. and Gales, R.. Chipping Norton: Surrey Beatty and Sons, pp. 69–83.Google Scholar
Prince, P. A., Weimerskirch, H., Wood, A. G. and Croxall, J. P. (1999). Areas and scales of interactions between albatrosses and the marine environment: species, populations and sexes. In Proceedings of the 22 International Ornitholological Congress, Durban, eds. Adams, N. J. and Slotow, R. H.. South Africa: Birdlife, pp. 2001–20.Google Scholar
Prins, H. H. T. (1987). The buffalo of Manyara. Ph.D. thesis, Rijksuniversiteit te Groningen.
Prins, H. H. T. (1989). Condition changes and choice of social environment in African buffalo bulls. Behaviour, 108, 297–324.CrossRefGoogle Scholar
Prins, H. H. T. (1996). Ecology and Behaviour of the African Buffalo: Social Inequality and Decision Making. London: Chapman & Hall.CrossRefGoogle Scholar
Prins, H. H. T. and Iason, G. R. (1989). Dangerous lions and nonchalant buffalo. Behaviour, 108, 262–95.CrossRefGoogle Scholar
Promislow, D. E. L., Montgomerie, R. and Martin, T. E. (1992). Mortality costs of sexual dimorphism in birds. Proceedings of the Royal Society of London B, 250, 143–50.CrossRefGoogle Scholar
Provenza, F. D. and Balph, D. F. (1988). Development of dietary choice in livestock on rangelands and its implications for management. Journal of Animal Science, 66, 2356–68.CrossRefGoogle Scholar
Pusey, A. E. (1990). Behavioural changes at adolescence in chimpanzees. Behaviour, 115, 203–46.CrossRefGoogle Scholar
Pusey, A. E., Williams, J. and Goodall, J. (1997). The influence of dominance rank on the reproductive success of female chimpanzees. Science, 277, 828–31.CrossRefGoogle ScholarPubMed
Puttick, G. M. (1981). Sex-related differences in foraging behaviour of curlew sandpipers. Ornis Scandinavica, 12, 13–17.CrossRefGoogle Scholar
Qualls, R., Shine, R., Donnellan, S. and Hutchinson, M. (1996). The evolution of viviparity within the Australian scincid lizard Lerista bougainvillii. Journal of Zoology, London, 237, 13–26.CrossRefGoogle Scholar
Quinn, J. S. (1990). Sexual size dimorphism and parental care patterns in a monomorphic and a dimorphic larid. Auk, 107, 260–74.CrossRefGoogle Scholar
Quintana, F. and Dell'Arciprete, O. P. (2002). Foraging grounds of southern giant petrels (Macronectes giganteus) on the Patagonian shelf. Polar Biology, 25, 159–61.CrossRefGoogle Scholar
Radespiel, U. (2000). Sociality in the grey mouse lemur (Microcebus murinus). American Journal of Primatology, 51, 21–40.3.0.CO;2-C>CrossRefGoogle Scholar
Radespiel, U., Cepok, S., Zietemann, V. and Zimmermann, E. (1998). Sex specific usage patterns of sleeping sites in grey mouse lemurs (Microcebus murinus) in northwest Madagascar. American Journal of Primatology, 46, 77–84.3.0.CO;2-S>CrossRefGoogle Scholar
Radespiel, U., Ehresmann, P. and Zimmermann, E. (2001a). Contest versus scramble competition for mates: the composition and spatial structure of a population of grey mouse lemurs (Microcebus murinus) in Madagascar. Primates, 42, 207–20.CrossRefGoogle Scholar
Radespiel, U., Sarikaya, Z., Zimmermann, E. and Bruford, M. (2001b). Sociogenetic structure in a free-living nocturnal primate population: Sex specific differences in the grey mouse lemur (Microcebus murinus). Behavioral Ecology and Sociobiology, 50, 493–502.CrossRefGoogle Scholar
Radford, A. N. and Plessis, M. A. (2003). Bill dimorphism and foraging niche partitioning in the green woodhoopoe. Journal of Animal Ecology, 72, 258–69.CrossRefGoogle Scholar
Rahmani, A. R. (1989). The Great Indian Bustard. Bombay: Bombay Natural History Society.Google Scholar
Rajporohit, L. S., Sommer, V. and Mohnot, S. M. (1995). Wanderers between harems and bachelor bands. Male hanuman langurs (Presbytis entellus) at Jodhpur in Rajasthan. Behaviour, 132, 255–99.CrossRefGoogle Scholar
Ralls, K. and Mesnick, S. L. (2002). Sexual dimorphism. In Encyclopedia of Marine Mammals, eds. Perrin, W. F., Wursig, B. and Thewissen, J. G. M.. San Diego: Academic Press.Google Scholar
Raman, T. R. S. (1997). Factors influencing seasonal and monthly changes in the group size of chital or axis deer in southern India. Journal of Biosciences, 22, 203–18.CrossRefGoogle Scholar
Ramírez Ávila, G. M., Guisset, J. L. and Deneubourg, J. L. (2003). Synchronization in light-controlled oscillators. Physica, 182, 254–73.Google Scholar
Ramirez-Bautista, A. and Benabib, M. (2001). Perch height of the arboreal lizard Anolis nebulosus (Sauria: Polychrotidae) from a tropical dry forest of Mexico: effect of the reproductive season. Copeia, 2001, 187–93.CrossRefGoogle Scholar
Rands, S. A., Cowlishaw, G., Pettifor, R. A., Rowcliffe, J. M. and Johnstone, R. A. (2003). Spontaneous emergence of leaders and followers in foraging pairs. Nature, 423, 432–4.CrossRefGoogle ScholarPubMed
Ransome, R. D. (1990). The Natural History of Hibernating Bats. London: Christopher Helm.Google Scholar
Ranta, E. and Lindström, K. (1990). Assortative schooling in 3-spined sticklebacks. Annales Zoologici Fennici, 27, 67–75.Google Scholar
Ranta, E., Peuhkuri, N. and Laurila, A. (1994). A theoretical exploration of antipredatory and foraging factors promoting phenotype-assorted fish schools. Ecoscience, 1, 99–106.CrossRefGoogle Scholar
Rau, G. H., Ainley, D. G., Bengtson, J. L., Torre, J. J. and Hopkins, T. L. (1992). 15N/14N and 13C/12C in Weddell Sea birds, seals, and fish: implications for diet and trophic structure. Marine Ecology Progress Series, 84, 1–8.CrossRefGoogle Scholar
Read, A. J. (1990). Reproductive seasonality in harbour porpoises, Phocoena phocoena, from the Bay of Fundy. Canadian Journal of Zoology, 68, 284–8.CrossRefGoogle Scholar
Read, A. J. (1999). Harbour porpoise – Phocoena phocoena (Linnaeus, 1758). In Handbook of Marine Mammals: The Second Book of Dolphins and the Porpoises, vol. 6, eds. Ridgway, S. H. and Harrison, R.. London: Academic Press, pp. 323–56.Google Scholar
Read, A. J. and Hohn, A. A. (1995). Life in the fast lane: the life history of harbor porpoises from the Gulf of Maine. Marine Mammal Science, 11, 423–40.CrossRefGoogle Scholar
Read, A. J. and Tolley, K. A. (1997). Postnatal growth and allometry of harbour porpoises from the Bay of Fundy. Canadian Journal of Zoology, 75, 122–30.CrossRefGoogle Scholar
Read, A. J., Wells, R. S., Hohn, A. A. and Tolley, K. A. (1993). Patterns of growth in wild bottlenose dolphins, Tursiops truncatus. Journal of Zoology, London, 231, 107–23.CrossRefGoogle Scholar
Read, D. (1982). Observations of the movements of two arid zone planigales (Dasyuridae, Marsupialia). In Carnivorous Marsupials, ed. Archer, M.. Sydney: Royal Zoological Society of New South Wales, pp. 227–31.Google Scholar
Read, D. G. (1984a). Movements and home ranges of three sympatric dasyurids, Sminthopsis crassicaudata, Planigale gilesi and P. tenuirostris (Marsupialia), in semiarid western New South Wales. Australian Wildlife Research, 11, 223–34.CrossRefGoogle Scholar
Read, D. G. (1984b). Reproduction and breeding season of Planigale gilesi and P. tenuirostris (Marsupialia: Dasyuridae). Australian Mammalogy, 8, 161–74.Google Scholar
Read, D. G. (1989). Microhabitat separation and diel activity patterns of Planigale gilesi and P. tenuirostris (Marsupialia: Dasyuridae). Australian Mammalogy, 12, 45–54.Google Scholar
Read, D. G. (1995). Giles' Planigale. In The Mammals of Australia, ed. Strahan, R.. Sydney: Reed New Holland, pp. 107–9.Google Scholar
Reaney, L. T. and Whiting, M. J. (2002). Life on a limb: ecology of the tree agama (Acanthocercus a. atricollis) in southern Africa. Journal of Zoology, London, 257, 439–48.CrossRefGoogle Scholar
Recchia, C. A. and Read, A. J. (1989). Stomach content of harbour porpoises, Phoceona phoceona, (L.), from the Bay of Fundy. Canadian Journal of Zoology, 67, 2140–6.CrossRefGoogle Scholar
Recher, H. F. and Holmes, R. T. (2000). The foraging ecology of birds of eucalypt forest and woodland. I. Differences between males and females. Emu, 100, 205–15.CrossRefGoogle Scholar
Reddy, E. and Fenton, M. B. (2003). Exploiting vulnerable prey: moths and red bats (Lasiurus borealis; Vespertilionidae). Canadian Journal of Zoology, 18, 1553–60.CrossRefGoogle Scholar
Reebs, S. G. and Saulnier, N. (1997). The effect of hunger on shoal choice in golden shiners (Pisces: Cyprinidae, Notemigonus crysoleucas). Ethology, 103, 642–52.CrossRefGoogle Scholar
Reed, R. N. and Shine, R. (2002). Lying in wait for extinction: ecological correlates of conservation status among Australian elapid snakes. Conservation Biology, 16, 451–61.CrossRefGoogle Scholar
Reid, K. (1995). The diet of Antarctic fur seals Arctocephalus gazella Peters 1875 during winter at South Georgia. Antarctic Science, 7(3), 241–9.CrossRefGoogle Scholar
Reid, K. and Arnould, J. P. Y. (1996). The diet of Antarctic fur seals Arctocephalus gazella during the breeding season at South Georgia. Polar Biology, 16, 105–14.CrossRefGoogle Scholar
Reijinders, P., Brasseur, S., Toorn, J.et al. (1993). Seals, Fur Seals, Sea Lions, and Walrus. Status Survey and Conservation Action Plan. Gland, Switzerland: IUCN.Google Scholar
Reilly, J. J. and Fedak, M. A. (1991). Rates of water turnover and energy-expenditure of free-living male common seals (Phoca vitulina). Journal of Zoology, 223, 461–8.CrossRefGoogle Scholar
Reinert, H. K. (1984). Habitat variation within sympatric snake populations. Ecology, 65, 1673–82.CrossRefGoogle Scholar
Reinert, H. K. and Kodrich, W. R. (1982). Movements and habitat utilization by the massasauga, Sistrurus catenatus catenatus. Journal of Herpetology, 16, 162–71.CrossRefGoogle Scholar
Reinert, H. K. and Zappalorti, R. T. (1988). Timber rattlesnakes (Crotalus horridus) of the Pine Barrens (New Jersey, USA): their movement patterns and habitat preference. Copeia, 1988, 964–78.CrossRefGoogle Scholar
Remis, M. J. (1999). Tree structure and sex differences in arboreality among western lowland gorillas (Gorilla gorilla gorilla) at Bai Hokou, Central African Republic. Primates, 40, 383–96.CrossRefGoogle Scholar
Rendell, L. and Whitehead, H. (2001). Culture in whales and dolphins. Behavioral and Brain Sciences, 24, 309–82.CrossRefGoogle ScholarPubMed
Renfree, M. B., Russell, E. M. and Wooller, R. D. (1984). Reproduction and life history of the Honey Possum, Tarsipes rostratus. In Possums and Gliders, eds. Smith, A. P. and Hume, I. D.. Chipping Norton, NSW: Surrey Beatty and Sons, pp. 427–37.Google Scholar
Rexstad, E. A. and Anderson, D. R. (1988). Effect of the point system on redistributing hunting pressure on mallards. Journal of Wildlife Management, 52, 89–94.CrossRefGoogle Scholar
Reynolds, J. D. and Jennings, S. (2000). The role of animal behavior in marine conservation. In Behavior and Conservation, eds. Gosling, L. M. and Sutherland, W. J.. Cambridge, United Kingdom: Cambridge University Press, pp. 238–57.Google Scholar
Reznick, D. (1996). Life history evolution in guppies: A model system for the empirical study of adaptation. Netherlands Journal of Zoology, 46, 172–90.CrossRefGoogle Scholar
Rhine, R., Bloland, P. and Lodwick, L. (1985). Progression of adult male chacma baboons (Papio ursinus) in the Moremi Wildlife Reserve. International Journal of Primatology, 6, 115–22.CrossRefGoogle Scholar
Rice, D. W. (1998). Marine Mammals of the World, Systematics and Distribution. Lawrence, KA, USA: Society of Marine Mammalogy.Google Scholar
Rice, D. W. (1989). Sperm whale – Physeter macrocephalus (Linnaeus, 1758). In Handbook of Marine Mammals: River Dolphins and the Larger Toothed Whales, vol. 4, eds. Ridgway, S. H. and Harrison, R.. London: Academic Press, pp. 177–233.Google Scholar
Richard, A., Rakotomanaga, P. and Schwartz, M. (1993). Dispersal by Propithecus verrauxi at Beza Mahafaly Reserve. American Journal of Primatology, 30, 1–20.CrossRefGoogle Scholar
Richard, C. and Pépin, D. (1990). Seasonal variation in intragroup-spacing behavior of foraging isards (Rupicapra pyrenaica). Journal of Mammalogy, 71, 145–50.CrossRefGoogle Scholar
Richard, K. R., Dillon, M. C., Whitehead, H. and Wright, J. M. (1996). Patterns of kinship in groups of free-living sperm whales (Physeter macrocephalus) revealed by multiple molecular analyses. Proceedings of the Academy of Science of United States, 93, 8792–5.CrossRefGoogle Scholar
Richard-Hansen, C. (1992). Association between individually marked isards (Rupicapra pyrenaica): seasonal and inter-annual variations. In Ongulés/Ungulates 91, eds. Spitz, F., Janeau, G., Gonzalez, G., and Aulagnier, S.. Paris: S.F.E.P.M.-I.R.G.M., pp. 299–304.Google Scholar
Richards, A. F. (1996). Life history and behavior of female dolphins (Tursiops sp.) in Shark Bay, Western Australia. Ph.D thesis, Ann Arbor, Michigan: University of Michigan.Google Scholar
Richardson, A. J., Maharaj, G., Compagno, L. J. V.et al. (2000). Abundance, distribution, morphometrics, reproduction and diet of the Izak catshark. Journal of Fish Biology, 56, 552–76.CrossRefGoogle Scholar
Ridgway, S. H. and Harrison, R. (1989). Handbook of Marine Mammals: River Dolphins and the Larger Toothed Whales. London: Academic Press.Google Scholar
Ridgway, S. H. and Harrison, R. (1994). Handbook of Marine Mammals: The First Book of Dolphins. London: Academic Press.Google Scholar
Ridgway, S. H. and Harrison, R. (1999). Handbook of Marine Mammals: The Second Book of Dolphins and the Porpoises. London: Academic Press.Google Scholar
Riedman, M. (1990). The Pinnipeds: Seals, Sea Lions, and Walruses. Berkeley: University of California Press.Google Scholar
Rintamäki, P. T., Karvonen, E., Alatalo, R. V. and Lundberg, A. (1999). Why do black grouse males perform on lek sites outside the breeding season?Journal of Avian Biology, 30, 359–66.CrossRefGoogle Scholar
Ripley, W. E. (1946). The soupfin shark and the fishery. California Fish Bulletin, 64, 7–37.Google Scholar
Rivas, J. and Burghardt, G. M. (2001). Understanding sexual size dimorphism in snakes: wearing the snake's shoes. Animal Behaviour, 62, F1–F6.CrossRefGoogle Scholar
Robert, K. A. and Thompson, M. B. (2001). Viviparous lizard selects sex of embryos. Nature, 412, 698–9.CrossRefGoogle ScholarPubMed
Roberts, C., Brotherton, P., Luschekina, A. A., Kholodova, M. V. and Milner-Gulland, E. J. (2001). Gazelles, dwarf antelopes and saigas. In The New Encyclopedia of Mammals, ed. Macdonald, D.. Oxford: Oxford University Press, pp. 560–6.Google Scholar
Robichaud, D. and Rose, G. A. (2003). Sex differences in cod residency on a spawning ground. Fisheries Research, 60, 33–43.CrossRefGoogle Scholar
Rock, J. and Cree, A. (2003). Intraspecific variation in the effect of temperature on pregnancy in the viviparous gecko Hoplodactylus maculatus. Herpetologica, 59, 8–22.CrossRefGoogle Scholar
Rock, J., Andrews, R. M. and Cree, A. (2000). Effects of reproductive condition, season, and site on selected temperatures of a viviparous gecko. Physiological and Biochemical Zoology, 73, 344–55.CrossRefGoogle ScholarPubMed
Rodenhouse, N. L., Sherry, T. W. and Holmes, R. T. (1997). Site-dependent regulation of population size: a new synthesis. Ecology, 78, 2025–42.Google Scholar
Rodhouse, P. G., Elvidge, C. D. and Trathan, P. N. (2001). Remote sensing of the global light-fishing fleet: an analysis of interactions with oceanography, other fisheries and predators. Advances in Marine Biology, 39, 261–303.CrossRefGoogle Scholar
Rodman, P. (1988). Diversity and consistency in ecology and behavior. In Orang-Utan Biology, ed. Schwartz, J. H.. Oxford: Oxford University Press, pp. 31–51.Google Scholar
Rodman, P. S. (1979). Individual activity patterns and the solitary nature of orangutans. In The Great Apes, eds. Hamburg, D. A. and McKown, E. R.. Menlo Park: Benjamin Cummings, pp. 235–56.Google Scholar
Rodman, P. S. and Mitani, J. C. (1987). Orangutans: sexual dimorphism in a solitary species. In Primate Societies, eds. Smuts, B. B., Cheney, D. L., Seyfarth, R. M., Wrangham, R. W. and Struhsaker, T. T.. Chicago: University of Chicago Press, pp. 146–54.Google Scholar
Rodway, M. S., Regehr, H. M. and Cook, F. (2003). Sex and age differences in distribution, abundance, and habitat preferences of wintering harlequin ducks: implications for conservation and estimating recruitment rates. Canadian Journal of Zoology, 81, 492–503.CrossRefGoogle Scholar
Romeo, G., Lovari, S., Festa-Bianchet, M. (1997). Group leaving in mountain goats: are males ousted by adult females?Behavioural Processes, 40, 243–6.CrossRefGoogle ScholarPubMed
Rook, A. J. and Penning, P. D. (1991). Synchronization of eating, ruminating and idling activity by grazing sheep. Applied Animal Behaviour Science, 32, 157–66.CrossRefGoogle Scholar
Roosenburg, W. M., Halcy, K. L. and McGuire, S. (1999). Habitat selection and movements of diamondback terrapins, Malaclemys terrapin, in a Maryland estuary. Chelonian Conservation Biology, 3, 425–9.Google Scholar
Rootes, W. L. and Chabreck, R. H. (1993). Cannibalism in the American alligator. Herpetologica, 49, 99–107.Google Scholar
Rose, B. R. (1981). Factors affecting activity in Sceloporus virgatus. Ecology, 62, 706–16.CrossRefGoogle Scholar
Rose, F. L. and Judd, F. W. (1975). Activity and home range size of the Texas tortoise, Gopherus berlandieri, in South Texas. Herpetologica, 31, 448–56.Google Scholar
Rose, L. (1994). Sex differences in diet and foraging in Cebus capucinus. International Journal of Primatology, 15, 95–114.CrossRefGoogle Scholar
Ross, C. (1992). Basal metabolic rate, body weight, and diet in primates: an evaluation of the evidence. Folia Primatologica, 58, 7–23.CrossRefGoogle Scholar
Ross, D. A. (1989). Population ecology of painted and Blanding's turtles (Chrysemys picta and Emydoidea blandingii) in central Wisconsin. Transactions of the Wisconsin Academy of Science, Arts and Letters, 77, 77–84.Google Scholar
Ross, D. A., Brewster, K. N., Anderson, R. K., Ratner, N. and Brewster, C. M. (1991). Aspects of the ecology of wood turtles, Clemmys insculpta, in Wisconsin. Canadian Field-Naturalist, 105, 363–7.Google Scholar
Ross, P. I., Jalkotzy, M. G. and Festa-Bianchet, M. (1997). Cougar predation on bighorn sheep in southwestern Alberta during winter. Canadian Journal of Zoology, 75, 771–5.CrossRefGoogle Scholar
Rothstein, A. and Griswold, J. G. (1991). Age and sex preferences for social partners by juvenile bison bulls, Bison bison. Animal Behaviour, 41, 227–37.CrossRefGoogle Scholar
Rowe, J. W. and Moll, E. O. (1991). A radiotelemetric study of activity and movements of the Blanding's turtle (Emydoidea blandingii) in northeastern Illinois. Journal of Herpetology, 25, 178–85.CrossRefGoogle Scholar
Rowell, T. E. (1973). Social organization of wild talapoin monkeys. American Journal of Physical Anthropology, 38, 593–8.CrossRefGoogle ScholarPubMed
Rubin, E. S., Boyce, W. M. and Bleich, V. C. (2000). Reproductive strategies of desert bighorn sheep. Journal of Mammalogy, 81, 769–86.2.3.CO;2>CrossRefGoogle Scholar
Rubin, E. S., Boyce, W. M. and Caswell-Chen, E. P. (2002). Modeling demographic processes in an endangered population of bighorn sheep. Journal of Wildlife Management, 66, 796–810.CrossRefGoogle Scholar
Rubin, E. S., Boyce, W. M., Jorgensen, M. C.et al. (1998). Distribution and abundance of bighorn sheep in the Peninsular Ranges, California. Wildlife Society Bulletin, 26, 539–51.Google Scholar
Ruby, D. E. and Baird, D. I. (1994). Intraspecific variation in behavior: comparisons between populations at different altitudes of the lizard Sceloporus jarrovi. Journal of Herpetology, 28, 70–8.CrossRefGoogle Scholar
Ruckstuhl, K. E. (1998). Foraging behaviour and sexual segregation in bighorn sheep. Animal Behaviour, 56, 99–106.CrossRefGoogle ScholarPubMed
Ruckstuhl, K. E. (1999). To synchronise or not to synchronise: a dilemma for young bighorn males?Behaviour, 136, 805–18.CrossRefGoogle Scholar
Ruckstuhl, K. E. and Festa-Bianchet, M. (1998). Do reproductive status and lamb gender affect the foraging behavior of bighorn ewes?Ethology, 104, 941–54.CrossRefGoogle Scholar
Ruckstuhl, K. E. and Festa-Bianchet, M. (2001). Group choice by subadult male bighorn sheep: trade-offs between foraging efficiency and predator avoidance. Ethology, 107, 161–72.CrossRefGoogle Scholar
Ruckstuhl, K. E. and Kokko, H. (2002). Modelling sexual segregation in ungulates: effects of group size, activity budgets and synchrony. Animal Behaviour, 64, 909–14.CrossRefGoogle Scholar
Ruckstuhl, K. E. and Neuhaus, P. (2000). Sexual segregation in ungulates: a new approach. Behaviour, 137, 361–77.CrossRefGoogle Scholar
Ruckstuhl, K. E. and Neuhaus, P. (2001). Behavioral synchrony in ibex groups: effects of age, sex, and habitat. Behaviour, 138, 1033–46.CrossRefGoogle Scholar
Ruckstuhl, K. E. and Neuhaus, P. (2002). Sexual segregation in ungulates: a comparative test of three hypotheses. Biological Review, 77, 77–96.CrossRefGoogle ScholarPubMed
Ruckstuhl, K. E., Manica, A., MacColl, A. D. C., Pilkington, J. G., Clutton-Brock, T. H. (2005). The effects of castration, sex ratio and population density on social segregation and habitat use in Soay sheep. Behavioral Ecology and Sociobiology (in press).
Ruedas, L. A., Demboski, J. R. and Sison, R. V. (1994). Morphological and ecological variation in Otopteropus cartilagonodus Koch, 1969 (Mammalia: Chiroptera: Pteropodidae) from Luzon, Philippines. Proceedings of the Biological Society of Washington, 107, 1–16.Google Scholar
Ruibal, R. and Philibosian, R. (1974). The population ecology of the lizard Anolis acutus. Ecology, 55, 525–37.CrossRefGoogle Scholar
Russell, E. M. (1979). The size and composition of groups in the red kangaroo, Macropus rufus. Australian Wildlife Research, 6, 237–44.CrossRefGoogle Scholar
Russell, E. M. (1986). Observations on the behaviour of the honey possum, Tarsipes rostratus (Marsupialia: Tarsipedidae) in captivity. Australian Journal of Zoology, 121, 1–63.Google Scholar
Russo, D. (2002). Elevation affects the distribution of the two sexes in Daubenton's bats Myotis daubentonii (Chiroptera: Vespertilionidae) from Italy. Mammalia, 66, 543–51.CrossRefGoogle Scholar
Rutherford, P. L. and Gregory, P. T. (2003). How age, sex, and reproductive condition affect retreat-site selection and emergence, patterns in a temperate-zone lizard, Elgaria coerulea. Ecoscience, 10, 24–32.CrossRefGoogle Scholar
Ryan, P. G. and Boix-Hinzen, C. (1999). Consistent male-biased seabird mortality in the Patagonian toothfish longline fishery. Auk, 116, 851–4.CrossRefGoogle Scholar
Sabo, J. L. (2003). Hot rocks or no hot rocks: overnight retreat availability and selection by a diurnal lizard. Oecologia, 136, 329–35.CrossRefGoogle ScholarPubMed
Sachs, B. D. and Harris, V. S. (1978). Sex differences and developmental changes in selected juvenile activities (play) of domestic lambs. Animal Behaviour, 26, 678–84.CrossRefGoogle Scholar
Sailer, L. D., Gaulin, S. J. C., Boster, J. S. and Kurland, J. A. (1985). Measuring the relationship between dietary quality and body size in primates. Primates, 26, 14–27.CrossRefGoogle Scholar
Salamolard, M. and Weimerskirch, H. (1993). Relationship between foraging effort and energy requirement throughout the breeding season in the wandering albatross. Functional Ecology, 7, 643–52.CrossRefGoogle Scholar
Saltz, E., Dixon, D. and Johnson, J. (1977). Training disadvantaged preschoolers on various fantasy activities: Effects on cognitive functioning and impulse control. Child Development, 48, 367–80.CrossRefGoogle ScholarPubMed
Sanderson, R. A. (1974). Sexual dimorphism in the Barbour's map turtle, Malaclemys barbouri (Carr and Marchand). M.A. thesis, Tampa: University of South Florida.Google Scholar
Santos, X. and Llorente, G. A. (1998). Sexual and size-related differences in the diet of the snake Natrix maura from the Ebro Delta, Spain. Herpetological Journal, 8, 161–5.Google Scholar
Santos, X., Gonzalez-Solis, J. and Llorente, G. A. (2000). Variation in the diet of the viperine snake Natrix maura in relation to prey availability. Ecography, 23, 185–92.CrossRefGoogle Scholar
Sato, K., Mitani, Y., Cameron, M. F.et al. (2002). Deep foraging dives in relation to the energy depletion of Weddell seal (Leptonychotes weddellii) mothers during lactation. Polar Biology, 25, 696–702.Google Scholar
Savidge, J. A. (1988). Food habits of Boiga irregularis, an introduced preditor on Guam. Journal of Herpetology, 22, 275–82.CrossRefGoogle Scholar
Sayler, R. D. and Afton, A. D. (1981). Ecological aspects of common goldeneyes Bucephala clangula wintering on the Mississippi River, USA. Ornis Scandinavica, 12, 99–108.CrossRefGoogle Scholar
Schaefer, J. A. and Messier, F. (1995). Habitat selection as a hierarchy: the spatial scales of winter foraging by musk oxen. Ecography, 18, 333–44.CrossRefGoogle Scholar
Schaefer, R. J., Torres, S. G. and Bleich, V. C. (2000). Survivorship and cause-specific mortality in sympatric populations of mountain sheep and mule deer. California Fish and Game, 86, 127–35.Google Scholar
Schmid, J. and Kappeler, P. M. (1998). Fluctuating sexual dimorphism and differential hibernation by sex in a primate, the grey mouse lemur. Behavioral Ecology and Sociobiology, 43, 125–32.CrossRefGoogle Scholar
Schmid-Nielson, K. (1989). Scaling. Why is Animal Size so Important?Cambridge: Cambridge University Press.Google Scholar
Schmidt-Nielsen, K. (1972). Locomotion: energy cost of swimming, flying, and running. Science, 177, 222–8.CrossRefGoogle ScholarPubMed
Schoener, T. W. (1967). The ecological significance of sexual dimorphism in size in the lizard Anolis conspersus. Science, 155, 474–6.CrossRefGoogle ScholarPubMed
Schoener, T. W. (1968). The Anolis lizards of Bimini: resource partitioning in a complex fauna. Ecology, 49, 704–26.CrossRefGoogle Scholar
Schoener, T. W. and Gorman, G. C. (1968). Some niche differences in three lesser Antillean lizards of the genus Anolis. Ecology, 49, 819–30.CrossRefGoogle Scholar
Schoener, T. W., Slade, J. B. and Stinson, C. H. (1982). Diet and sexual dimorphism in the very catholic lizard genus, Leiocephalus of the Bahamas. Oecologia, 53, 160–9.CrossRefGoogle ScholarPubMed
Schreer, J. F. and Kovacs, K. M. (1997). Allometry of diving capacity in air-breathing vertebrates. Canadian Journal of Zoology, 75, 339–58.CrossRefGoogle Scholar
Schwaner, T. D. and Sarre, S. D. (1988). Body size of tiger snakes in southern Australia, with particular reference to Notechis ater serventyi (Elapidae) on Chappell Island. Journal of Herpetology, 22, 24–33.CrossRefGoogle Scholar
Schwartz, O. A., Bleich, V. C. and Holl, S. A. (1986). Genetics and the conservation of mountain sheep Ovis canadensis nelsoni. Biological Conservation, 37, 179–90.CrossRefGoogle Scholar
Schwarzkopf, L. (1994). Measuring trade-offs: a review of studies of costs of reproduction in lizards. In Lizard Ecology: Historical and Experimental Perspectives, eds. Vitt, L. J. and Pianka, E. R.. Princeton, NJ: Princeton University Press, pp. 7–30.CrossRefGoogle Scholar
Schwarzkopf, L. and Shine, R. (1991). Thermal biology of reproduction in viviparous skinks, Eulamprus tympanum: why do gravid females bask more?Oecologia, 88, 562–9.CrossRefGoogle ScholarPubMed
Schwarzkopf, L. and Shine, R. (1992). Costs of reproduction in lizards: escape tactics and susceptibility to predation. Behavioral Ecology and Sociobiology, 31, 17–25.CrossRefGoogle Scholar
Scott, M. D., Wells, R. S. and Irvine, A. B. (1990). A long-term study of bottlenose dolphin on the west coast of Florida. In The Bottlenose Dolphin, eds. Leatherwood, S. and Reeves, R. R.. San Diego: Academic Press, pp. 235–44.Google Scholar
Scott, N. J. Jr, Wilson, D. E., Jones, C. and Andrews, R. M. (1976). The choice of perch dimensions by lizards of the genus Anolis (Reptilia, Lacertilia, Iguanidae). Journal of Herpetology, 10, 75–84.CrossRefGoogle Scholar
Scudder, R. M. and Burghardt, G. M. (1983). A comparative study of defensive behavior in three sympatric species of water snakes (Nerodia). Zeitschrift für Tierpsychologie, 63, 17–26.CrossRefGoogle Scholar
Searcy, W. A. and Yasukawa, K. (1981). Sexual size dimorphism and survival of male and female blackbirds (Icteridae). Auk, 98, 457–65.Google Scholar
Secor, S. M. (1994). Ecological significance of movements and activity range for the sidewinder, Crotalus cerastes. Copeia, 1994, 631–45.CrossRefGoogle Scholar
Seebacher, F., Grigg, G. C. and Beard, L. A. (1999). Crocodiles as dinosaurs: behavioural thermoregulation in very large ectotherms leads to high and stable body temperatures. Journal of Experimental Biology, 202, 77–86.Google ScholarPubMed
Seghers, B. H. (1973). Zoology. The University of British Columbia.Google Scholar
Seghers, B. H. (1974). Schooling behaviour in the guppy (Poecilia reticulata): An evolutionary response to predation. Evolution, 28, 486–9.Google ScholarPubMed
Seib, R. L. (1981). Size and shape in a neotropical burrowing colubrid snake, Geophis nasalis, and its prey. American Zoologist, 21, 933.Google Scholar
Selander, R. K. (1966). Sexual dimorphism and differential niche utilization in birds. Condor, 68, 113–51.CrossRefGoogle Scholar
Selander, R. K. (1972). Sexual selection and dimorphism in birds. In Sexual Selection and the Descent of Man 1971–1971, ed. Campbell, B.. Chicago: Heinemann, pp. 180–229.Google Scholar
Senft, R. L., Coughenour, M. B., Bailey, D. W.et al. (1987). Large herbivore foraging and ecological hierarchies. Bioscience, 37, 789–99.CrossRefGoogle Scholar
Sergeant, D. E. (1973). Biology of white whales (Delphinapterus leucas) in Western Hudson Bay. Fisheries Research Board of Canada, 30, 1065–90.CrossRefGoogle Scholar
Sergeant, D. E. and Brodie, P. F. (1969). Body size in white whales, Delphinapterus leucas. Journal of the Fisheries Research Board of Canada, 26, 2561–80.CrossRefGoogle Scholar
Sexton, O. J. (1964). Differential predation by the lizard, Anolis carolinensis, upon unicolored and polycolored insects after an interval of no contact. Animal Behaviour, 12, 101–10.CrossRefGoogle Scholar
Sexton, O. J., Bauman, J. and Ortleb, E. (1972). Seasonal food habits of Anolis limifrons. Ecology, 53, 182–6.CrossRefGoogle Scholar
Shackleton, D. M. (1991). Social maturation and productivity in bighorn sheep: are young males incompetent. Applied Animal Behaviour Science, 29, 173–84.CrossRefGoogle Scholar
Shaffer, S. A., Weimerskirch, H. and Costa, D. P. (2001). Functional significance of sexual dimorphism in wandering albatrosses, Diomedea exulans. Functional Ecology, 15, 203–10.CrossRefGoogle Scholar
Shah, B., Shine, R., Hudson, S. and Kearney, M. (2003). Sociality in lizards: why do thick-tailed geckos (Nephrurus milii) aggregate?Behaviour, 140, 1039–52.CrossRefGoogle Scholar
Shank, C. C. (1982). Age-sex differences in the diets of wintering Rocky Mountain bighorn sheep. Ecology, 63, 627–33.CrossRefGoogle Scholar
Shank, C. C. (1985). Inter- and intra-sexual segregation of chamois (Rupicapra rupicapra) by altitude and habitat during summer. Zeitschrift für Säugetierkunde, 50, 117–25.Google Scholar
Shealer, D. A. (2002). Foraging behaviour and food of seabirds. In Biology of Marine Birds, eds. Schreiber, E. A. and Burger, J.. Boca Raton, Florida, USA: CRC Press, pp. 137–77.Google Scholar
Shealy, R. M. (1976). The natural history of the Alabama map turtle, Graptemys pulchra Baur, in Alabama. Bulletin of the Florida State Museum, Biological Sciences Series, 21, 47–111.Google Scholar
Sherry, T. W. and Holmes, R. T. 1996. Winter habitat quality, population limitation and conservation of Neotropical-Nearctic migrant birds. Ecology, 77, 36–48.CrossRefGoogle Scholar
Shetty, S. and Shine, R. (2002a). Activity patterns of yellow-lipped sea kraits (Laticauda colubrina) on a Fijian island. Copeia, 2002, 77–85.CrossRefGoogle Scholar
Shetty, S. and Shine, R. (2002b). Sexual divergence in diets and morphology in Fijian sea snakes Laticauda colubrina (Laticaudinae). Austral Ecology, 27, 77–84.CrossRefGoogle Scholar
Shimoka, Y. (2003). Seasonal variation in association patterns of wild spider monkeys (Ateles belzebuth belzebuth) at La Macarena, Columbia. Primates, 44, 83–90.Google Scholar
Shine, R. (1978). Sexual size dimorphism and male combat in snakes. Oecologia, 33, 269–78.CrossRefGoogle ScholarPubMed
Shine, R. (1979). Activity patterns in Australian elapid snakes (Squamata: Serpentes: Elapidae). Herpetologica, 35, 1–11.Google Scholar
Shine, R. (1980a). ‘Costs’ of reproduction in reptiles. Oecologia, 46, 92–100.CrossRefGoogle Scholar
Shine, R. (1980b). Ecology of the Australian death adder, Acanthophis antarcticus (Elapidae): evidence for convergence with the Viperidae. Herpetologica, 36, 281–9.Google Scholar
Shine, R. (1985). The evolution of viviparity in reptiles: an ecological analysis. In Biology of the Reptilia, vol. 15, eds. Gans, C. and Billett, F.. New York: John Wiley and Sons, pp. 605–94.Google Scholar
Shine, R. (1986). Sexual differences in morphology and niche utilization in an aquatic snake, Acrochordus arafurae. Oecologia, 69, 260–7.CrossRefGoogle Scholar
Shine, R. (1989a). Ecological causes for the evolution of sexual dimorphism: a review of the evidence. Quarterly Review of Biology, 64, 419–60.CrossRefGoogle Scholar
Shine, R. (1989b). Constraints, allometry and adaptation: food habits and reproductive biology of Australian brownsnakes (Pseudonaja, Elapidae). Herpetologica, 45, 195–207.Google Scholar
Shine, R. (1991a). Intersexual dietary divergence and the evolution of sexual dimorphism in snakes. American Naturalist, 138, 103–22.CrossRefGoogle Scholar
Shine, R. (1991b). Why do larger snakes eat larger prey?Functional Ecology, 5, 493–502.CrossRefGoogle Scholar
Shine, R. (1994a). Allometric patterns in the ecology of Australian snakes. Copeia, 1994, 851–67.CrossRefGoogle Scholar
Shine, R. (1994b). Sexual size dimorphism in snakes revisited. Copeia, 1994, 326–46.CrossRefGoogle Scholar
Shine, R. (1999). Why is sex determined by nest temperature in many reptiles?Trends in Ecology and Evolution, 14, 186–9.CrossRefGoogle ScholarPubMed
Shine, R. and Crews, D. (1988). Why male garter snakes have small heads: the evolution and endocrine control of sexual dimorphism. Evolution, 42, 1105–10.CrossRefGoogle ScholarPubMed
Shine, R. and Fitzgerald, M. (1995). Variation in mating systems and sexual size dimorphism between populations of the Australian python Morelia spilota (Serpentes: Pythonidae). Oecologia, 103, 490–8.CrossRefGoogle Scholar
Shine, R. and Fitzgerald, M. (1996). Large snakes in a mosaic rural landscape: the ecology of carpet pythons Morelia spilota (Serpentes: Pythonidae) in coastal eastern Australia. Biological Conservation, 76, 113–22.CrossRefGoogle Scholar
Shine, R. and Harlow, P. S. (1993). Maternal thermoregulation influences offspring viability in a viviparous lizard. Oecologia, 96, 122–7.CrossRefGoogle Scholar
Shine, R. and Wall, M. (2004). Why is intraspecific niche partitioning more common in snakes than in lizards? In Foraging Modes in Lizards, eds. Miles, D. B., Reilly, S. M. and McBrayer, L. D.. Cambridge University Press.Google Scholar
Shine, R., Barrott, E. G. and Elphick, M. J. (2002a). Some like it hot: effects of forest clearing on nest temperatures of montane reptiles. Ecology, 83, 2808–15.CrossRefGoogle Scholar
Shine, R., Branch, W. R., Harlow, P. S. and Webb, J. K. (1998a). Reproductive biology and food habits of horned adders, Bitis caudalis (Viperidae), from southern Africa. Copeia, 1998, 391–401.CrossRefGoogle Scholar
Shine, R., Cogger, H. G., Reed, R. N. S., Shetty, S. and Bonnet, X. (2002b). Aquatic and terrestrial locomotor speeds of amphibious sea-snakes (Serpentes, Laticaudidae). Journal of Zoology, London, 259, 261–8.CrossRefGoogle Scholar
Shine, R., Elphick, M. and Donnellan, S. (2002c). Co-occurrence of multiple, supposedly incompatible modes of sex determination in a lizard population. Ecology Letters, 5, 486–9.CrossRefGoogle Scholar
Shine, R., Haagner, G. V., Branch, W. R., Harlow, P. S. and Webb, J. K. (1996). Natural history of the African shieldnose snake Aspidelaps scutatus (Serpentes, Elapidae). Journal of Herpetology, 30, 361–6.CrossRefGoogle Scholar
Shine, R., Harlow, P. S., Keogh, J. S. and Boeadi, (1998b). The influence of sex and body size on food habits of a giant tropical snake, Python reticulatus. Functional Ecology, 12, 248–258.CrossRefGoogle Scholar
Shine, R., Langkilde, T. and Mason, R. T. (2003b). Cryptic forcible insemination: male snakes exploit female physiology, anatomy and behavior to obtain coercive matings. American Naturalist, 162, 653–67.CrossRefGoogle Scholar
Shine, R., O'Connor, D., LeMaster, M. P. and Mason, R. T. (2001). Pick on someone your own size: ontogenetic shifts in mate choice by male garter snakes result in size-assortative mating. Animal Behaviour, 61, 1–9.CrossRefGoogle Scholar
Shine, R., Olsson, M. M., Lemaster, M. P., Moore, I. T. and Mason, R. T. (2000). Effects of sex, body size, temperature, and location on the antipredator tactics of free-ranging gartersnakes (Thamnophis sirtalis, Colubridae). Behavioral Ecology, 11, 239–45.CrossRefGoogle Scholar
Shine, R., Reed, R. N., Shetty, S. and Cogger, H. G. (2002d). Relationships between sexual dimorphism and niche partitioning within a clade of sea-snakes (Laticaudinae). Oecologia, 133, 45–53.CrossRefGoogle Scholar
Shine, R., Shine, T. and Shine, B. G. (2003a). Intraspecific habitat partitioning by the sea snake Emydocephalus annulatus (Serpentes, Hydrophiidae): the effects of sex, body size, and colour pattern. Biological Journal of the Linnean Society, 80, 1–10.CrossRefGoogle Scholar
Shine, R., Sun, L. X., Fitzgerald, M. and Kearney, M. (2003c). A radiotelemetric study of movements and thermal biology of insular Chinese pit-vipers (Gloydius shedaoensis, Viperidae). Oikos, 100, 342–52.CrossRefGoogle Scholar
Shine, R., Phillips, B., Langkilde, T., Lutterschmidt, D. and Mason, R. T. (2004). Mechanisms and consequences of sexual conflict in garter snakes (Thamnophis sirtalis, Colubridae). Behavioral Ecology, 15, 654–60.CrossRefGoogle Scholar
Shively, S. H. (1982). Factors limiting the upstream distribution of the Sabine map turtle. M.Sc. thesis, Lafayette, LA: University of Southwestern Louisiana.Google Scholar
Shively, S. H. and Jackson, J. F. (1985). Factors limiting the upstream distribution of the Sabine map turtle. American Midland Naturalist, 114, 292–303.CrossRefGoogle Scholar
Short, H. L. (1963). Rumen fermentation and energy relationships in white-tailed deer. Journal of Wildlife Management, 27, 184–95.CrossRefGoogle Scholar
Shumway, C. A. (1999). A neglected science: applying behavior to aquatic conservation. Environmental Biology of Fishes, 55, 183–210.CrossRefGoogle Scholar
Silk, J. (1987). Activities and feeding behavior of free-ranging pregnant female baboons. International Journal of Primatology, 8, 596–613.CrossRefGoogle Scholar
Sillet, T. S. and Holmes, R. T. (2002). Variation in survivorship of a migratory songbird throughout its annual cycle. Journal of Animal Ecology, 71, 296–308.CrossRefGoogle Scholar
Simon, C. A. (1975). Size selection of prey by the lizard Sceloporus jarrovi. American Midland Naturalist, 96, 246–51.Google Scholar
Sims, D. W. (1996). The effect of body size on the standard metabolic rate of lesser spotted dogfish, Scyliorhinus canicula. Journal of Fish Biology, 48, 542–4.CrossRefGoogle Scholar
Sims, D. W. (2003). Tractable models for testing theories about natural strategies: foraging behaviour and habitat selection of free-ranging sharks. Journal of Fish Biology, 63 (Supplement A), 53–73.CrossRefGoogle Scholar
Sims, D. W. and Davies, S. J. (1994). Does specific dynamic action (SDA) regulate return of appetite in the lesser spotted dogfish, Scyliorhinus canicula?Journal of Fish Biology, 45, 341–8.Google Scholar
Sims, D. W. and Quayle, V. A. (1998). Selective foraging behaviour of basking sharks on zooplankton in a small-scale front. Nature, 393, 460–4.CrossRefGoogle Scholar
Sims, D. W., Davies, S. J. and Bone, Q. (1993). On the diel rhythms in metabolism and activity of post-hatching lesser spotted dogfish, Scyliorhinus canicula. Journal of Fish Biology, 43, 749–54.CrossRefGoogle Scholar
Sims, D. W., Genner, M. J., Southward, A. J. and Hawkins, S. J. (2001b). Timing of squid migration reflects North Atlantic climate variability. Proceedings of the Royal Society of London B, 268, 2607–11.CrossRefGoogle Scholar
Sims, D. W., Nash, J. P. and Morritt, D. (2001a). Movements and activity of male and female dogfish in a tidal sea lough: alternative behavioural strategies and apparent sexual segregation. Marine Biology, 139, 1165–75.Google Scholar
Sinclair, A. R. E. (1977). The African Buffalo. A Study of Resource Limitation of Populations. Chicago: University of Chicago Press.Google Scholar
Singleton, I. and Schaik, C. P. (2001). Orangutan home range size and its determinants. International Journal of Primatology, 22, 877–912.CrossRefGoogle Scholar
Siniff, D. B., DeMaster, D. P., Hofman, R. J. and Eberhardt, L. L. (1977). An analysis of the dynamics of a Weddell seal population. Ecological Monographs, 47, 319–35.CrossRefGoogle Scholar
Slip, D. J. and Shine, R. (1988a). Feeding habits of the diamond python, Morelia s. spilota: ambush predation by a boid snake. Journal of Herpetology, 22, 323–30.CrossRefGoogle Scholar
Slip, D. J. and Shine, R. (1988b). Habitat use, movements, and activity patterns of free-ranging diamond pythons, Morelia spilota spilota (Serpentes, Boidae): a radiotelemetric study. Australian Wildlife Research, 15, 515–31.CrossRefGoogle Scholar
Slip, D. J., Hindell, M. A. and Burton, H. R. (1994). Diving behaviour of southern elephant seals from Macquarie Island: An overview. In Elephant Seals: Population Ecology, Behaviour and Physiology, eds. Boeuf, B. J. and Laws, R. M.. Berkley: University of California Press, pp. 253–70.Google Scholar
Slotow, R., Dyk, G., Poole, J., Page, B. and Klocke, A. (2000). Older bull elephants control young males. Nature, 408, 425–6.CrossRefGoogle ScholarPubMed
Smith, A. P. and Broome, L. (1992). The effects of season, sex and habitat on the diet of the mountain pygmy-possum (Burramys parvus). Wildlife Research, 19, 755–68.CrossRefGoogle Scholar
Smith, C. (1977). Feeding and ranging behavior of mantled howler monkeys (Alouatta palliata). In Primate Ecology, ed. Clutton-Brock, T. H.. Cambridge: Cambridge University Press, pp. 183–222.Google Scholar
Smith, G. J. D. and Gaskin, D. E. (1983). An environmental index for habitat utilization by female harbour porpoises with calves near Deer Island, Bay of Fundy. Ophelia, 22, 1–13.CrossRefGoogle Scholar
Smith, G. R. (1996). Habitat use and fidelity in the striped plateau lizard Sceloporus virgatus. American Midland Naturalist, 135, 68–80.CrossRefGoogle Scholar
Smith, G. R. and Ballinger, R. E. (1994). Thermal ecology of Sceloporus virgatus from southeastern Arizona, with comparison to Urosaurus ornatus. Journal of Herpetology, 28, 65–9.CrossRefGoogle Scholar
Smith, G. R., Ballinger, R. E. and Congdon, J. D. (1993). Thermal ecology of the high-altitude bunch grass lizard, Sceloporus scalaris. Canadian Journal of Zoology, 71, 2152–5.CrossRefGoogle Scholar
Smith, M. S. R. (1966). Injuries as an indication of social behaviour in the Weddell seal (Leptonychotes wedellii). Mammalia, 30, 241–6.CrossRefGoogle Scholar
Smith, P. C. and Evans, P. R. (1973). Studies of shorebirds at Lindisfarne, Northumberland. 1. Feeding ecology and behaviour of the bar-tailed godwit. Wildfowl, 24, 135–9.Google Scholar
Smith, R. J. and Jungers, W. L. (1997). Body mass in comparative primatology. Journal of Human Evolution, 32, 523–59.CrossRefGoogle ScholarPubMed
Smith, T. G., Hammill, M. O. and Martin, A. R. (1994). Herd composition and behaviour of white whales (Delphinapterus leucas) in two Canadian arctic estuaries. Bioscience, 39, 175–84.Google Scholar
Snell, H. L., Jennings, R. D., Snell, H. M. and Harcourt, S. (1988). Intrapopulation variation in predator avoidance performance of Galapagos lava lizards: the interaction of sexual and natural selection. Evolutionary Ecology, 2, 353–69.CrossRefGoogle Scholar
Snelson, F. F., Mulligan, T. J. and Williams, S. E. (1984). Food habits, occurrence, and population structure of the bull shark, Carcharhinus leucas, in Florida coastal lagoons. Bulletin of Marine Science, 34, 71–80.Google Scholar
Southwell, C. J. (1984). Variability in grouping in the eastern grey kangaroo, Macropus giganteus I. Group density and group size. Australian Wildlife Research, 11, 423–35.CrossRefGoogle Scholar
Spaeth, D. F., Bowyer, R. T., Stephenson, T. R. and Barboza, P. S. (2004). Sexual segregation in moose Alces alces: an experimental manipulation of foraging behaviour. Wildlife Biology, 10, 59–72.CrossRefGoogle Scholar
Speakman, J. R. and Thomas, D. W. (2003). Physiological ecology and energetics of bats. In Bat Ecology, eds. Kunz, T. H. and Fenton, M. B.. Chicago: University of Chicago Press, pp. 430–90.Google Scholar
Speakman, J. R., Irwin, N., Tallach, N. and Stone, R. (1999). Effect of roost size on the emergence behaviour of pipistrelle bats. Animal Behaviour, 58, 787–95.CrossRefGoogle ScholarPubMed
Sprague, D. S., Suzuki, S., Takahashi, H. and Sato, S. (1998). Male life histories in natural populations of Japanese macaques: migration, dominance rank, and troop participation of males in two habitats. Primates, 39, 351–63.CrossRefGoogle Scholar
Springer, M. S., Kirsch, J. A. W. and Case, J. A. (1997). The chronicle of marsupial evolution. In Molecular Evolution and Adaptive Radiation, eds. Givnish, T. J. and Sytsma, K. J.. Cambridge: Cambridge University Press, pp. 129–61.Google Scholar
Springer, S. (1967). Social organization of shark populations. In Sharks, Skates and Rays, eds. Gilbert, P. W., Mathewson, R. F. and Rall, D. P.. Baltimore MD: Johns Hopkins University Press, pp. 149–74.Google Scholar
Sroufe, L. A., Bennett, C., Englund, M., Urban, J. and Shulman, S. (1993). The significance of cross-gender boundaries in preadolescence: contemporary correlates and antecedents of boundary violation and maintenance. Child Development, 64, 455–66.CrossRefGoogle Scholar
Aubin, St D. J., Smith, T. G. and Geraci, J. R. (1990). Seasonal epidermal molt in beluga whales, Delphinapterus leucas. Canadian Journal of Zoology, 68, 359–67.CrossRefGoogle Scholar
Stahl, J. C. and Sagar, P. M. (2000). Foraging strategies of southern Buller's albatrosses Diomedea b. bulleri breeding on The Snares, New Zealand. Journal of the Royal Society of New Zealand, 30, 299–318.CrossRefGoogle Scholar
Staines, B. W. (1976). The use of natural shelter by red deer (Cervus elaphus) in relation to weather in North-East Scotland. Journal of Zoology, 180, 1–8.CrossRefGoogle Scholar
Staines, B. W. (1977). Factors affecting the seasonal distribution of red deer (Cervus elaphus, L.) in Glen Dye, North-East Scotland. Annals of Applied Biology, 87, 495–512.CrossRefGoogle Scholar
Staines, B. W., Crisp, J. M. and Parish, T. (1982). Differences in the quality of food eaten by red deer (Cervus elaphus) stags and hinds in winter. Journal of Applied Ecology, 19, 65–77.CrossRefGoogle Scholar
Stammbach, E. (1987). Desert, forest, and montane baboons: multilevel societies. In Primate Societies, eds. Smuts, B. B., Cheney, D. L., Seyfarth, R. M., Wrangham, R. W. and Struhsaker, T. T.. Chicago: University of Chicago Press, pp. 112–20.Google Scholar
Stamps, J. A. (1977). The function of the survey posture in Anolis lizards. Copeia, 1977, 756–8.CrossRefGoogle Scholar
Stamps, J. A. (1983). Sexual selection, sexual dimorphism, and territoriality. In Lizard Ecology: Studies of a Model Organism, eds. Huey, R. B., Pianka, E. R. and Schoener, T. W.. Cambridge, MA: Harvard University Press, pp. 169–204.CrossRefGoogle Scholar
Staniland, I. J. and Boyd, I. L. (2003). Variation in the foraging location of Antarctic fur seals (Arctocephalus gazella), the effects on diving behaviour. Marine Mammal Science, 19, 331–43.CrossRefGoogle Scholar
Steenbeek, R., Sterck, E. H. M., de Vries, H. and van Hooff, J. A. R. A. M. (2000). Costs and benefits of the one-male, age-graded, and all-male phase in wild Thomas' langurs groups. In Primate Males, ed. Kappeler, P.. Cambridge: Cambridge University Press, pp. 130–45.Google Scholar
Sterck, E. H., Watts, D. P. and Schaik, C. P. (1997). The evolution of social relationships in female primates. Behavioral Ecology and Sociobiology, 41, 291–309.CrossRefGoogle Scholar
Stevens, J. D. (1974). The occurrence and significance of tooth cuts on the blue shark (Prionace glauca L.) from British waters. Journal of the Marine Biological Association of the United Kingdom, 54, 373–8.CrossRefGoogle Scholar
Stevens, J. D. (1976). First results of shark tagging in the north-east Atlantic, 1972–1975. Journal of the Marine Biological Association of the United Kingdom, 56, 929–37.CrossRefGoogle Scholar
Stevick, P. T., Allen, J., Berube, M.et al. (2003). Segregation of migration by feeding ground origin in North Atlantic humpback whales (Megaptera novaeangliae). Journal of Zoology, London, 259, 231–7.CrossRefGoogle Scholar
Stewart, B. S. and Delong, R. L. (1994). Postbreeding foraging migrations of northern elephant seals. In Elephant Seals: Population Ecology, Behaviour, and Physiology, eds. Boeuf, B. J. and Laws, R. M.. Berkley: University of California Press, pp. 290–309.Google Scholar
Stewart, B. S., Craig, M. P. and Antonelis, G. A. (1998). Characterization of Hawaiian Monk Seal (Monachus schauinslandi) pelagic habitat, home range and diving behaviour. Honolulu: National Marine Fisheries Service, Southwest Fisheries Science Center Honolulu Laboratory, 2570 Dole Street, Honolulu, Hawaii 96822–2902.
Stirling, I. (1969). Ecology of the Weddell seal in McMurdo Sound, Antarctica. Ecology, 50, 573–86.CrossRefGoogle Scholar
Stirling, I. (1983). The evolution of mating systems in pinnipeds. In Advances in the Study of Mammalian Behaviour, eds. Eisenberg, J. F. and Kleinmann, D. G.. Lawrence, Kansas: Allen Press, pp. 489–527.Google Scholar
Stoddart, D. and Braithwaite, R. (1979). A strategy for utilization of regenerating heathland habitat by the brown bandicoot (Isoodon obesulus; Marsupialia, Peramelidae). Journal of Animal Ecology, 48, 165–79.CrossRefGoogle Scholar
Stokke, S. (1999). Sex differences in feeding-patch choice in a megaherbivore: elephants in Chobe National Park, Botswana. Canadian Journal of Zoology, 77, 1723–32.CrossRefGoogle Scholar
Stokke, S. and du Toit, J. T. (2000). Sex and size related differences in the dry season feeding patterns of elephants in Chobe National Park, Botswana. Ecography, 23, 70–80.CrossRefGoogle Scholar
Stokke, S. and du Toit, J. T. (2002). Sexual segregation in habitat use by elephants in Chobe National Park, Botswana. African Journal of Ecology, 40, 360–71.CrossRefGoogle Scholar
Storz, J. F. and Williams, C. F. (1996). Summer population structure of sub-alpine bats in Colorado. Southwestern Naturalist, 41, 322–4.Google Scholar
Storz, J. F., Bhat, H. R. and Kunz, T. H. (2000). Social structure of a polygynous tent-making bat, Cynopterous sphinx (Megachiroptera). Journal of Zoology, London, 251, 151–65.CrossRefGoogle Scholar
Strahan, R. (1995). The Mammals of Australia, 2nd edn. Sydney: Reed New Holland.Google Scholar
Straits, B. C. (1998). Occupational sex segregation: the role of personal ties. Journal of Vocational Behavior, 52(2), 191–207.CrossRefGoogle Scholar
Strikwerda, T. E., Fuller, M. R., Seegar, W. S., Howey, P. W. and Black, H. D. (1986). Bird-borne satellite transmitter and location program. Johns Hopkins APL Technical Digest, 7, 203–8.Google Scholar
Suhonen, J. and Kuitunen, M. (1991). Intersexual foraging niche differentiation within the breeding pair in the common treecreeper Certhia familiaris. Ornis Scandinavica, 22, 313–18.CrossRefGoogle Scholar
Sukumar, R. and Gadgil, M. (1988). Male-female differences in foraging on crops by Asian elephants. Animal Behaviour, 36, 1233–5.CrossRefGoogle Scholar
Summers, R. W., Westlake, G. E. and Feare, C. J. (1987). Differences in the ages, sexes and physical condition of starlings Sturnus vulgaris at the centre and periphery of roosts. Ibis, 129, 96–102.CrossRefGoogle Scholar
Sun, L. X., Shine, R., Zhao, D. B. and Tang, Z. R. (2002). Low costs, high output: reproduction in an insular pit-viper (Gloydius shedaoensis, Viperidae) from north-eastern China. Journal of Zoology, London, 256, 511–21.Google Scholar
Sund, O. (1943). Et brugdebarsel. Naturen, 67, 285–6.Google Scholar
Swennen, C. (1984). Differences in quality of roosting flocks of oystercatchers. In Coastal Waders and Wildfowl in Winter, eds. Evans, P. R., Goss-Custard, J. D. and Hale, W. G.. Cambridge: Cambridge University Press.Google Scholar
Swennen, C., Bruijn, L. L. M., Duiven, P., Leopold, M. F. and Marteijn, E. C. L. (1983). Differences in bill form of the oystercatcher Haematopus ostralegus; a dynamic adaptation to specific foraging techniques. Netherlands Journal of Sea Research, 17, 57–83.CrossRefGoogle Scholar
Swingland, I. R. and Lessells, C. M. (1979). The natural regulation of giant tortoise populations on Aldabra Atoll: movement polymorphism, reproductive success and mortality. Journal of Animal Ecology, 48, 639–54.CrossRefGoogle Scholar
Sydeman, W. J. and Nur, N. (1994). Life history strategies of female northern elephant seals. In Elephant Seals: Population Ecology, Behaviour and Physiology, eds. Boeuf, B. J. and Laws, R. M.. Berkley: University of California Press, pp. 137–53.Google Scholar
Talbot, J. J. (1979). Time budget, niche overlap, inter- and intraspecific aggression in Anolis humilis and A. limifrons from Costa Rica. Copeia, 1979, 472–81.CrossRefGoogle Scholar
Tamisier, A. (1985). Hunting as a key environmental parameter for the Western Palearctic duck populations. Wildfowl, 36, 95–103.Google Scholar
Taylor, R. J. (1981). The comparative ecology of the Eastern grey kangaroo and wallaroo in the New England tablelands of New South Wales. Ph.D. thesis, University of New England.Google Scholar
Taylor, R. J. (1983). Association of social classes of the wallaroo, Macropus robustus (Marsupialia: Macropodidae). Australian Wildlife Research, 10, 39–45.CrossRefGoogle Scholar
Tedman, R. A. and Bryden, M. M. (1979). Cow-pup behaviour of the Weddell seal, Leptonychotes weddelli (Pinnipedia), in McMurdo Sound, Antarctica. Australian Journal of Wildlife Research, 6, 19–37.CrossRefGoogle Scholar
Terborgh, J. (1990). Where Have All the Birds Gone?Princeton: Princeton University Press.Google Scholar
Terranova, M. L. and Laviola, G. (1995). Individual differences in mouse behavioural development: effects of precious weaning and ongoing sexual segregation. Animal Behaviour, 50, 1261–71.CrossRefGoogle Scholar
Testa, J. W. (1994). Over winter movements and diving behaviour of female Weddell seal (Leptonychotes weddellii) in the Southwestern Ross Sea, Antarctica. Canadian Journal of Zoology, 72, 1700–10.CrossRefGoogle Scholar
Theodorakis, C. W. (1989). Size segregation and the effects of oddity on predation risk in minnow schools. Animal Behaviour, 38, 496–502.CrossRefGoogle Scholar
Thill, R. E., Martin, A. Jr, Morris, H. F. Jr and , McCune E. D. (1987). Grazing and burning impacts on deer diets on Louisiana pine-bluestem range. Journal of Wildlife Management, 51, 873–80.CrossRefGoogle Scholar
Thomas, C. D., Baguette, M. and Lewis, O. T. (2000). Butterfly movement and conservation in patchy landscapes. In Behavior and Conservation, eds. Gosling, L. M. and Sutherland, W. J.. Cambridge, United Kingdom: Cambridge University Press, pp. 85–104.Google Scholar
Thomas, D. W. and Cloutier, D. (1992). Evaporative water loss by hibernating little brown bats, Myotis lucifugus. Physiological Ecology, 65, 433–56.Google Scholar
Thomas, L. N. (1987). The effects of stress on some aspects of the demography and physiology of Isoodon obesulus (Shaw and Nodder). Unpublished Masters thesis, University of Western Australia.Google Scholar
Thompson, P. M., Fedak, M. A., McConnell, B. J. and Nicholas, K. S. (1989). Seasonal and sex-related variation in the activity patterns of common seals (Phoca vitulina). Journal of Applied Ecology, 27, 521–35.CrossRefGoogle Scholar
Thompson, P. M., Mackay, A., Tollit, D. J., Enderby, S. and Hammond, P. S. (1998). The influence of body size and sex on the characteristics of harbour seal foraging trips. Canadian Journal of Zoology – Revue Canadienne De Zoologie, 76, 1044–53.CrossRefGoogle Scholar
Thompson, P. M., Tollit, D. J., Wood, D.et al. (1997). Estimating harbour seal abundance and distribution in an esturine habitat in N.E. Scotland. Journal of Applied Ecology, 34, 43–52.CrossRefGoogle Scholar
Thorhallsdottir, A. G., Provenza, F. D. and Balph, D. F. (1990). Ability of lambs to learn about novel foods while observing or participating with social models. Applied Animal Behaviour Science, 25, 25–33.CrossRefGoogle Scholar
Tickell, W. L. N. (1968). The biology of the great albatrosses Diomedea exulans and Diomedea epomophora. Antarctic Research Series, 12, 1–55.Google Scholar
Tickell, W. L. N. (2000). Albatrosses. East Sussex, England: Pica Press.Google Scholar
Tieszen, L. L. and Imbamba, S. K. (1980). Photosynthetic systems, carbon isotope discrimination and herbivore selectivity in Kenya. African Journal of Ecology, 18, 237–42.CrossRefGoogle Scholar
Tolley, K. A., Read, A. J., Well, R. S.et al. (1995). Sexual dimorphism in wild bottlenose dolphins (Tursiops truncatus) from Sarasota, Florida. Journal of Mammalogy, 76, 1190–8.CrossRefGoogle Scholar
Townshend, D. J. (1981). The importance of field feeding to the survival of wintering male and female curlews Numenius arquata on the Tees estuary. In Feeding and Survival Strategies of Estuarine Organisms, eds. Jones, N. V. and Wolff, W. J.. New York: Plenum Press.CrossRefGoogle Scholar
Tricas, T. C. and Feuvre, E. M. (1985). Mating in the reef white-tip shark Triaenodon obesus. Marine Biology, 84, 233–7.CrossRefGoogle Scholar
Trivers, R. (1971). The evolution of reciprocal altruism. Quarterly Review of Biology, 46, 35–57.CrossRefGoogle Scholar
Trivers, R. L. (1972). Parental investment and sexual selection. In Sexual Selection and the Descent of Man, ed. Campbell, B.. Chicago: Aldine, pp. 1871–971.Google Scholar
Tuck, G. N., Polacheck, T. and Bulman, C. (2003). Spatio-temporal trends of longline fishing effort in the Southern Ocean and implications for seabird by-catch. Biological Conservation, 114, 1–27.CrossRefGoogle Scholar
Tucker, A. D., FitzSimmons, N. N. and Gibbons, J. W. (1995). Resource partitioning by the estuarine turtle Malaclemys terrapin: trophic, spatial, and temporal foraging constraints. Herpetologica, 51, 167–81.Google Scholar
Tucker, A. D., Limpus, C. J., McCallum, H. I. and McDonald, K. R. (1997). Movements and home ranges of Crocodylus johnstoni in the Lynd River, Queensland. Wildlife Research, 24, 379–96.CrossRefGoogle Scholar
Tucker, A. D., McCallum, H. I., Limpus, C. J. and McDonald, K. R. (1998). Sex-biased dispersal in a long-lived polygynous reptile (Crocodylus johnstoni). Behavioral Ecology and Sociobiology, 44, 85–90.CrossRefGoogle Scholar
Tuttle, M. D. (1979). Status, causes of decline, and management of endangered grey bats. Journal of Wildlife Management, 43, 1–17.CrossRefGoogle Scholar
Twente, J. W. (1955). Some aspects of habitat selection and other behaviour of cavern-dwelling bats. Ecology, 36, 706–32.CrossRefGoogle Scholar
Tyndale-Biscoe, H. and Renfree, M. (1987). Reproductive Physiology of Marsupials. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Uetz, P. (2000). How many reptile species?Herpetological Review, 31, 13–15.Google Scholar
Lichtenbelt, Marken W. D., Wesselingh, R. A., Vogel, J. T. and Albers, K. B. M. (1993). Energy budgets in free-living green iguanas in a seasonal environment. Ecology, 74, 1157–72.CrossRefGoogle Scholar
van Noordwijk, A. J. (1994). The interaction of inbreeding depression and environmental stochasticity in the risk of extinction of small populations. In Conservation Genetics, eds. Loeschcke, V., Tomiuk, J. and Jain, S. K.. Basel, Switzerland: Birkhauser Verlag, pp. 131–46.CrossRefGoogle Scholar
Noordwijk, M. and Schaik, C. P. (2002). Career moves: transfers and rank challenge decisions by male long-tailed macaques. Behaviour, 138, 359–95.CrossRefGoogle Scholar
van Noordwijk, M., Hemelrijk, C. K., Herremans, L. A. M. and Sterck, E. H. M. (1993). Spatial position and behavioral sex differences in juvenile long-tailed macaques. In Juvenile Primates, eds. Pereira, M. E. and Fairbanks, L. A.. Oxford: Oxford University Press, pp. 77–85.Google Scholar
Schaik, C. P. (1983). Why are diurnal primates living in groups?Behaviour, 37, 120–44.CrossRefGoogle Scholar
van Schaik, C. P. (1989). The ecology of social relationships amongst female primates. In Comparative Socioecology, eds. Standen, V. and Foley, R.. London: Blackwell, pp. 195–218.Google Scholar
Schaik, C. P. (1999). The fission-fusion social system of orangutans. Primates, 40, 69–86.CrossRefGoogle Scholar
van Schaik, C. P. (2000a). Vulnerability to infanticide by males: patterns among mammals. In Infanticide by Males and Its Implications, eds. Schaik, C. P. and Janson, C. H.. Cambridge: Cambridge University Press, pp. 61–71.CrossRefGoogle Scholar
van Schaik, C. P. (2000b). Infanticide by male primates: the sexual selection hypothesis revisited. In Infanticide by Males and Its Implications, eds. Schaik, C. P. and Janson, C. H.. Cambridge: Cambridge University Press, pp. 27–60.CrossRefGoogle Scholar
Schaik, C. P. and Kappeler, P. M. (1997). Infanticide risk and the evolution of male-female association in primates. Proceedings of the Royal Society, London, B, 264, 1687–94.CrossRefGoogle ScholarPubMed
Schaik, C. P. and Noordwijk, M. A. (1989). The special role of male Cebus monkeys in predation avoidance and its effects on group composition. Behavioral Ecology and Sociobiology, 24, 265–76.CrossRefGoogle Scholar
Soest, P. J. (1982). Nutritional Ecology of the Ruminant. Corvallis: O and B Books.Google Scholar
Soest, P. J. (1994). Nutritional Ecology of the Ruminant, 2nd edn. Ithaca: Cornell University Press.Google Scholar
Soest, P. J. (1996). Allometry and ecology of feeding behavior and digestive capacity in herbivores. Zoo Biology, 15, 455–79.3.0.CO;2-A>CrossRefGoogle Scholar
Vaughan, T. A. (1976). Nocturnal behaviour of the African false vampire bat (Cardioderma cor). Journal of Mammalogy, 57, 227–48.CrossRefGoogle Scholar
Vaughan, T. A. and O'Shea, T. J. (1976). Roosting ecology of the pallid bat, Antrozous pallidus. Journal of Mammalogy, 57, 227–48.CrossRefGoogle Scholar
Vaughan, T. A. and Vaughan, R. P. (1986). Seasonality and the behaviour of the African yellow-winged bat (Lavia frons). Journal of Mammalogy, 67, 91–102.CrossRefGoogle Scholar
Vaughan, T. A. and Vaughan, R. P. (1987). Parental behaviour in the African yellow-winged bat (Lavia frons). Journal of Mammalogy, 68, 217–23.CrossRefGoogle Scholar
Vehrencamp, S. L., Stiles, F. G. and Bradbury, J. W. (1977). Observations on the foraging behaviour and avian prey of the neotropical carnivorous bat, Vampyrum spectrum. Journal of Mammalogy, 58, 469–78.CrossRefGoogle Scholar
Verme, L. J. (1988). Niche selection by male white-tailed deer: an alternative hypothesis. Wildlife Society Bulletin, 16, 448–51.Google Scholar
Vernes, K. and Pope, L. C. (2001). Stability of nest range, home range and movements of the northern bettong (Bettongia tropica) following moderate-intensity fire in a tropical woodland, north-eastern Queensland. Wildlife Research, 28, 141–50.CrossRefGoogle Scholar
Vernes, K. and Pope, L. C. (2002). Fecundity, pouch young survivorship and breeding season of the northern bettong (Bettongia tropica) in the wild. Australian Mammalogy, 23, 95–100.CrossRefGoogle Scholar
Vidal, R. M., Macias-Caballero, C. and Duncan, C. D. (1994). The occurrence and ecology of the golden-cheeked warbler in the highlands of Northern Chiapas, Mexico. Condor, 96, 684–91.Google Scholar
Viitanen, P. (1967). Hibernation and seasonal movements of the viper, Vipera berus berus (L.) in southern Finland. Annales Zoologici Fennici, 4, 472–546.Google Scholar
Villard, M.-A., Martin, P. R. and Drummond, C. G. (1993). Habitat fragmentation and pairing success in the ovenbird, Seiurus aurocapillus. Auk, 110, 759–68.Google Scholar
Villaret, J. C. and Bon, R. (1995). Social and spatial segregation in alpine ibex (Capra ibex) in Bargy, French Alps. Ethology, 101, 291–300.CrossRefGoogle Scholar
Villaret, J. C. and Bon, R. (1998). Sociality and relationships in Alpine ibex (Capra ibex). Revue d'Ecologie (Terre Vie), 53, 153–70.Google Scholar
Villaret, J. C., Rivet, A. and Bon, R. (1997). Sexual segregation of habitat by the alpine ibex in the French Alps. Journal of Mammalogy, 78, 1273–81.CrossRefGoogle Scholar
Vincent, A. and Sadovy, Y. (1998). Reproductive ecology in the conservation and management of fishes. In Behavioral Ecology and Conservation Biology, ed. Caro, T.. New York: Oxford University Press, pp. 209–45.Google Scholar
Vincent, S. E., Herrel, A. and Irschick, D. J. (2004). Ontogeny of intersexual head shape and prey selection in the pitviper Agkistrodon piscivorus. Biological Journal of the Linnean Society, 81, 151–9.CrossRefGoogle Scholar
Visagie, L. (2001). Grouping behaviour in the armadillo lizard, Cordylus cataphractus. M.Sc. thesis, Matieland, South Africa: University of Stellenbosch.Google Scholar
Vitt, L. J. and Cooper, W. E. (1986). Foraging and diet of a diurnal predator (Eumeces laticeps) feeding on hidden prey. Journal of Herpetology, 20, 408–15.CrossRefGoogle Scholar
Vitt, L. J. and Zani, P. A. (1996). Ecology of the elusive tropical lizard Tropidurus [equals Uracentron] flaviceps (Tropiduridae) in lowland rain forest of Ecuador. Herpetologica, 52, 121–32.Google Scholar
Vladykov, V. D. (1946). Nourriture du marsouin blanc ou béluga (Delphinapterus leucas) du fleuve Saint-Laurent. Etudes sur les mammifères aquatiques (IV). Québec: Département des pêcheries, Province de Québec, pp. 129.Google Scholar
Vogt, R. C. (1981). Food partitioning in three sympatric species of map turtle, genus Graptemys (Testudinata, Emydidae). American Midland Naturalist, 105, 102–11.CrossRefGoogle Scholar
Voigt, C. C. and Streich, W. J. (2003). Queuing for harem access in colonies of the greater sac-winged bat. Animal Behaviour, 65, 149–56.CrossRefGoogle Scholar
Voisin, J. (1968). Les pétrels géants (Macronectes halli et Macronectes giganteus) de l'ile de la Possession. L'Oiseau, 38, 7–122.Google Scholar
Voisin, J. (1991). Sur le régime et l'écologie alimentaires des pétrels géants Macronectes halli et M. giganteus de l'archipel Crozet. L'Oiseau, 61, 39–49.Google Scholar
Voisin, J. and Bester, M. N. (1981). The specific status of Giant petrels Macronectes at Gough Island. In Proceedings of the Symposium on Birds on the Sea and Shore, 1979, ed. Cooper, J.. Cape Town: African Seabird Group, pp. 215–22.Google Scholar
Volkman, N. J., Presler, P. and Trivelpiece, W. (1980). Diets of pygoscelid penguins at King George Island, Antarctica. Condor, 82, 373–8.CrossRefGoogle Scholar
Wagenknecht, E. (1986). Rotwild. Malsungen: Neumann-Neudamm Verlag.Google Scholar
Wagner, R. H. (1997). Differences in prey species delivered to nestlings by male and female razorbills Alca torda. Seabird, 19, 58–9.Google Scholar
Walker, B. G. and Bowen, W. D. (1993). Changes in body-mass and feeding-behavior in male harbor seals, Phoca vitulina, in relation to female reproductive status. Journal of Zoology, 231, 423–36.CrossRefGoogle Scholar
Wanless, S., Corfield, T., Harris, M. P., Buckland, S. T. and Morris, J. A. (1993). Diving behaviour of the shag (Aves: Pelecaniformes) in relation to water depth and prey size. Journal of Zoology, London, 231, 11–25.CrossRefGoogle Scholar
Wanless, S., Finney, S. K., Harris, M. P. and McCafferty, D. J. (1999). Effect of the diel light cycle on the diving behaviour of two bottom feeding marine birds: the blue-eyed shag Phalacrocorax atriceps and the European shag P. aristotelis. Marine Ecology Progress Series, 188, 219–24.CrossRefGoogle Scholar
Wanless, S., Harris, M. P., Morris, J. A. (1995). Factors affecting daily activity budgets of South-Georgian shags during chick rearing at Bird Island, South Georgia. Condor, 97, 550–8.CrossRefGoogle Scholar
Wapstra, E., Olsson, M., Shine, R.et al. (2004). Maternal basking behaviour determines offspring sex in a viviparous reptile. Biology Letters, 271, S230–S232.Google Scholar
Ward, A. J. W. and Hart, P. J. B. (2003). The effects of kin and familiarity on interactions between fish. Fish and Fisheries, 4, 348–58.CrossRefGoogle Scholar
Ward, S. J. (1990). Life history of the feathertail glider, Acrobates pygmaeus (Acrobatidae: Marsupialia) in south-eastern Australia. Australian Journal of Zoology, 38, 503–17.CrossRefGoogle Scholar
Ward, S. J. and Renfree, M. B. (1988). Reproduction in females of the feathertail glider, Acrobates pygmaeus (Marsupialia). Journal of Zoology, London, 216, 225–39.CrossRefGoogle Scholar
Wardle, C. S. (1993). Fish behaviour and fishing gear. In Behaviour of Teleost Fishes, ed. Pitcher, T. J.. London: Chapman and Hall, pp. 607–43.CrossRefGoogle Scholar
Warham, J. (1990). The Petrels: Their Ecology and Breeding Systems. London: Academic Press.Google Scholar
Waser, P. M. (1977). Feeding, ranging, and group size in the mangabey, Cercocebus alibigena. In Primate Ecology, ed. Clutton-Brock, T. H.. Cambridge: Cambridge University Press, pp. 183–222.Google Scholar
Waser, P. M. (1985). Spatial structure in mangabey groups. International Journal of Primatology, 6, 569–80.CrossRefGoogle Scholar
Waterman, J. M. (1995). The social organization of the Cape ground squirrel (Xerus inauris; Rodentia: Sciuridae). Ethology, 101, 130–47.CrossRefGoogle Scholar
Waterman, J. M. (1997). Why do male Cape ground squirrels live in groups?Animal Behaviour, 53, 809–17.CrossRefGoogle Scholar
Waters, J. C. (1974). The biological significance of the basking habit in the black-knobbed sawback, Graptemys nigronoda Cagle. M.Sc. thesis, Auburn, AL: Auburn University.Google Scholar
Watkins, J. L., Buchholz, F., Priddle, J., Morris, D. J. and Ricketts, C. (1992). Variation in reproductive status of Antarctic krill swarms evidence for a size-related sorting mechanism?Marine Ecology Progress Series, 82, 163–74.CrossRefGoogle Scholar
Watson, A. and Staines, B. W. (1978). Differences in the quality of wintering areas used by male and female red deer (Cervus elaphus) in Aberdeenshire. Journal of Zoology, London, 186, 544–50.Google Scholar
Watt, E. M. and Fenton, M. B. (1995). DNA fingerprinting provides evidence of discriminant suckling and non-random mating in little brown bats, Myotis lucifugus. Molecular Ecology, 4, 261–4.CrossRefGoogle Scholar
Watts, D. P. (1984). Composition and variability of mountain gorilla diets in the central Virungas. American Journal of Primatology, 7, 325–56.CrossRefGoogle Scholar
Watts, D. P. (1988). Environmental influences on mountain gorilla time budgets. American Journal of Primatology, 15, 295–312.CrossRefGoogle Scholar
Watts, D. P. (1991). Ecology of gorillas and its relationship to female transfer in mountain gorillas. International Journal of Primatology, 11, 21–45.CrossRefGoogle Scholar
Watts, D. P. and Mitani, J. C. (2001). Boundary patrols and intergroup encounters in wild chimpanzees. Behaviour, 138, 299–327.CrossRefGoogle Scholar
Watts, P. S., Hansen, S. and Lavigne, D. M. (1993). Models of heat loss by marine mammals: Thermoregulation below the zone of irrelevance. Journal of Theoretical Biology, 163, 505–25.CrossRefGoogle Scholar
Webb, G. J. W. and Smith, A. M. A. (1984). Sex ratio and survivorship in the Australian freshwater crocodile, Crocodylus johnstoni. Symposia of the Zoological Society of London, 52, 319–55.Google Scholar
Webb, G. J. W., Buckworth, R. and Manolis, S. C. (1983). Crocodylus johnstoni in the McKinlay River area, N. T. III. Growth, movement and the population age structure. Australian Wildlife Research, 10, 383–401.CrossRefGoogle Scholar
Webb, J. K. and Shine, R. (1997). A field study of spatial ecology and movements of a threatened snake species, Hoplocephalus bungaroides. Biological Conservation, 82, 203–17.CrossRefGoogle Scholar
Webb, J. K., Brown, G. P. and Shine, R. (2001). Body size, locomotor speed and antipredator behaviour in a tropical snake (Tropidonophis mairii, Colubridae): the influence of incubation environments and genetic factors. Functional Ecology, 15, 561–8.CrossRefGoogle Scholar
Webb, R. G. (1961). Observations on the life histories of turtles (genus Pseudemys and Graptemys) in Lake Texoma, Oklahoma. American Midland Naturalist, 65, 193–214.CrossRefGoogle Scholar
Weckerly, F. W. (1993). Intersexual resource partitioning in black-tailed deer: a test of the body size hypothesis. Journal of Wildlife Management, 57, 475–94.CrossRefGoogle Scholar
Weckerly, F. W. (1998). Sexual-size dimorphism: influence of mass and mating systems in the most dimorphic mammals. Journal of Mammalogy, 79, 33–52.CrossRefGoogle Scholar
Weckerly, F. W., Ricca, M. A. and Meyer, K. P. (2001). Sexual segregation in Roosevelt elk: cropping rates and aggression in mixed-sex groups. Journal of Mammalogy, 82, 825–35.2.0.CO;2>CrossRefGoogle Scholar
Weeden, R. B. (1964). Spatial separation of sexes in rock and willow ptarmigan in winter. Auk, 81, 534–41.CrossRefGoogle Scholar
Weidt, A., Hagenah, N., Randrianambinina, B., Radespiele, U. and Zimmerman, E. (2004). Social organization of the golden brown mouse lemur (Microcebus ravelobensis). American Journal of Physical Anthropology, 123, 40–51.CrossRefGoogle Scholar
Weihs, D. and Webb, P. W. (1983). Optimisation of locomotion. In Fish Biomechanics, eds. Webb, P. W. and Weihs, D.. New York: Praeger, pp. 339–71.Google Scholar
Weilgart, L. S. and Whitehead, H. (1986). Observations of a sperm whale (Physeter catodon) birth. Journal of Mammalogy, 67, 399–401.CrossRefGoogle Scholar
Weilgart, L., Whitehead, H. and Payne, K. (1996). A colossal convergence. American Scientist, 84, 278–87.Google Scholar
Weimerskirch, H. (1991). Sex-specific differences in molt strategy in relation to breeding in the wandering albatross. Condor, 93, 731–7.CrossRefGoogle Scholar
Weimerskirch, H. (1992). Reproductive effort in long-lived birds: age-specific patterns of condition, reproduction and survival in the wandering albatross. Oikos, 64, 464–73.CrossRefGoogle Scholar
Weimerskirch, H. (1995). Regulation of foraging trips and incubation routine in male and female wandering albatrosses. Oecologia, 102, 37–43.CrossRefGoogle ScholarPubMed
Weimerskirch, H. (1998). Foraging strategies of Indian Ocean albatrosses and their relationships with fisheries. In Albatross Biology and Conservation, eds. Robertson, G. and Gales, R.. Chipping Norton, Australia: Surrey Beatty and Sons, pp. 168–79.Google Scholar
Weimerskirch, H. and Jouventin, P. (1987). Population dynamics of the wandering albatross, Diomedea exulans, of the Crozet Islands: causes and consequences of the population decline. Oikos, 49, 315–22.CrossRefGoogle Scholar
Weimerskirch, H. and Jouventin, P. (1998). Changes in population sizes and demographic parameters of six albatross species breeding on the French sub-Antarctic islands. In Albatross Biology and Conservation, eds. Robertson, G. and Gales, R.. Chipping Norton, Australia: Surrey Beatty and Sons, pp. 84–91.Google Scholar
Weimerskirch, H. and Wilson, R. P. (2000). Oceanic respite for wandering albatrosses. Nature, 406, 955–6.CrossRefGoogle ScholarPubMed
Weimerskirch, H., Barbraud, C. and Lys, P. (2000). Sex differences in parental investment and chick growth in Wandering Albatrosses: Fitness consequences. Ecology, 81, 309–18.CrossRefGoogle Scholar
Weimerskirch, H., Bonadonna, F., Bailleul, F.et al. (2002). GPS tracking of foraging albatrosses. Science, 295, 1259–69.CrossRefGoogle ScholarPubMed
Weimerskirch, H., Brothers, N. and Jouventin, P. (1997a). Population dynamics of wandering albatross Diomedea exulans and the Amsterdam albatross D. amsterdamensis in the Indian Ocean and their relationships with long-line fisheries: conservation implications. Biological Conservation, 79, 257–70.CrossRefGoogle Scholar
Weimerskirch, H., Cherel, Y., Cuenot-chaillet, F. and Ridoux, V. (1997b). Alternative foraging strategies and resource allocation by male and female wandering albatrosses. Ecology, 78, 2051–63.CrossRefGoogle Scholar
Weimerskirch, H., Lequette, B. and Jouventin, P. (1989). Development and maturation of plumage in the wandering albatross Diomedea exulans. Journal of Zoology, London, 219, 411–21.CrossRefGoogle Scholar
Weimerskirch, H., Salamolard, M., Sarrazin, F. and Jouventin, P. (1993). Foraging strategy of wandering albatrosses through the breeding season: a study using satellite telemetry. Auk, 110, 325–42.Google Scholar
Wells, R. S. (2003). Dolphin social complexity: lessons from long-term study and life history. In Animal Social Complexity: Intelligence, Culture, and Individualized Societies, eds. Waal, F. B. M. and Tyack, P. L.. Cambridge, Massachusetts: Harvard University Press, pp. 32–56.CrossRefGoogle Scholar
Wells, R. S. and Scott, M. S. (1999). Bottlenose dolphin – Tursiops truncatus (Montagu, 1821). In Handbook of Marine Mammals: The Second Book of Dolphins and Porpoises, vol. 6, eds. Ridgway, S. H. and Harrison, R.. London: Academic Press, pp. 137–82.Google Scholar
Wells, R. S., Irvine, A. B. and Scott, M. D. (1980). The social ecology of inshore odontocetes. In Cetacean Behavior: Mechanisms and Functions, ed. Herman, L. M.. New York: John Wiley and Sons, pp. 263–317.Google Scholar
Wells, R. S., Scott, M. D. and Irvine, A. B. (1987). The social structure of free-ranging Bottlenose dolphins. In Current Mammalogy, vol. 1, ed. Genoways, H. H.. New York: Plenum Press, pp. 247–305.CrossRefGoogle Scholar
Wells, R. T. (1978). Field observations of the hairy-nosed wombat, Lasiorhinus latifrons (Owen). Australian Wildlife Research, 5, 299–303.CrossRefGoogle Scholar
Wetherbee, B. M., Gruber, S. H. and Cortes, E. (1990). Diet, feeding habits, digestion and consumption in sharks, with special reference to the lemon shark, Negaprion brevirostris. In Elasmobranchs as Living Resources: Advances in the Biology, Ecology, Systematics and Status of the Fisheries, NOAA Technical Report 90, eds. Pratt, H. L., Gruber, S. H. and Taniuchi, T.. Seattle, WA: National Oceanic and Atmospheric Administration, pp. 29–47.Google Scholar
Whitaker, P. B. and Shine, R. (2001). Thermal biology and activity patterns of the eastern brownsnake (Pseudonaja textilis): a radiotelemetric study. Herpetologica, 58, 436–62.CrossRefGoogle Scholar
Whitaker, P. B. and Shine, R. (2003). A radiotelemetric study of movements and shelter-site selection by free-ranging brownsnakes (Pseudonaja textilis, Elapidae). Herpetological Monographs, 17, 130–44.CrossRefGoogle Scholar
White, M. and Kolb, J. A. (1974). A preliminary study of Thamnophis near Sagehagen Creek, California. Copeia, 1974, 126–36.CrossRefGoogle Scholar
Whitehead, H. (1993). The behaviour of mature male sperm whales on the Galapagos Islands breeding grounds. Canadian Journal of Zoology, 71, 689–99.CrossRefGoogle Scholar
Whitehead, H. (1996). Babysitting, dive synchrony, and indications of alloparental care in sperm whales. Behavioral Ecology and Sociobiology, 38, 237–44.CrossRefGoogle Scholar
Whitehead, H. (2003). Sperm Whales: Social Evolution in the Ocean. Chicago: The University of Chicago Press.Google Scholar
Whitehead, H. and Mann, J. (2000). Female reproductive strategies of cetaceans: Life histories and calf care. In Cetacean Societies: Field Studies of Dolphins and Whales, eds. Mann, J., Connor, R. C., Tyack, P. L. and Whitehead, H.. Chicago: University of Chicago Press, pp. 219–47.Google Scholar
Whitehead, H. and Weilgart, L. (2000). The sperm whale: Social females and roving males. In Cetacean Societies: Field Studies of Dolphins and Whales, eds. Mann, J., Connor, R. C., Tyack, P. L. and Whitehead, H.. Chicago: University of Chicago Press, pp. 154–73.Google Scholar
Whitehead, H., Waters, S. and Lyrholm, T. (1991). Social organization of female sperm whales and their offspring: constant companions and casual acquaintances. Behavioral Ecology and Sociobiology, 29, 385–9.CrossRefGoogle Scholar
Whiting, B. and Edwards, C. (1973). A cross-cultural analysis of sex-differences in the behavior of children age 3 through 11. Journal of Social Psychology, 91, 171–88.CrossRefGoogle Scholar
Whiting, M. J. and Greeff, J. M. (1997). Facultative frugivory in the Cape flat lizard, Platysaurus capensis (Sauria: Cordylidae). Copeia, 1997, 811–18.CrossRefGoogle Scholar
Wickings, E. J. and Dixon, A. F. (1992). Testicular function, secondary sexual development, and social status in male mandrills (Mandrillus sphinx). Physiology and Behavior, 52, 909–16.CrossRefGoogle Scholar
Wikelski, M. and Trillmich, F. (1994). Foraging strategies of the Galapagos marine iguana (Amblyrhynchus cristatus): adapting behavioral rules to ontogenic size change. Behaviour, 128, 255–79.CrossRefGoogle Scholar
Wilkinson, G. S. (1992). Information transfer at evening bat colonies. Animal Behaviour, 44, 501–18.CrossRefGoogle Scholar
Wilkinson, G. S. (1995). Information transfer in bats. In Ecology, Evolution and Behaviour of Bats, eds. Racey, P. A. and Swift, S. M.. Symposium of the Zoological Society of London 67. Oxford: Clarendon Press, pp. 345–360.Google Scholar
Wilkinson, L. C. and Barclay, R. M. R. (1997). Differences in the foraging behaviour of male and female big brown bats (Eptesicus fuscus) during the reproductive period. Ecoscience, 4, 279–85.CrossRefGoogle Scholar
Wilkinson, G. S. and Boughman, J. W. (1998). Social calls co-ordinate foraging in greater spear-nosed bats. Animal Behaviour, 55, 337–50.CrossRefGoogle Scholar
Williams, A. J. (1982). Sexual size-dimorphism in the growth of Macronectes. Acta XVIII Congressus Internationalis Ornithologici, 2, 1190–1.Google Scholar
Williams, J. M., Oehlert, G. W., Carlis, J. V. and Pusey, A. E. (2004). Why do male chimpanzees defend a group range?Animal Behaviour, 68, 523–32.CrossRefGoogle Scholar
Williams, J. M., Pusey, A. E., Carlis, J. V., , Goodall J. and Farm, B. E. (2002). Space use and group membership in female chimpanzees: alternative strategies. Animal Behaviour, 63, 347–60.CrossRefGoogle Scholar
Williams, T. A. and Christiansen, J. L. (1981). The niches of two sympatric softshell turtles, Trionyx muticus and Trionyx spiniferus, in Iowa. Journal of Herpetology, 15, 30–8.CrossRefGoogle Scholar
Williams, T. C., Ireland, L. C. and Williams, J. M. (1973). High altitude flights of the free-tailed bat, Tadarida brasiliensis observed with radar. Journal of Mammalogy, 54, 807–21.CrossRefGoogle Scholar
Williams, T. D. (1991). Foraging ecology and diet of gentoo penguins Pygoscelis papua at South Georgia during winter and an assessment of their winter prey consumption. Ibis, 133, 3–13.CrossRefGoogle Scholar
Williams, T. M. (1999). The evolution of low cost efficient swimming in marine mammals: limits to energetic optimization. Philosophical Transactions of the Royal Society of London, 354, 193–201.CrossRefGoogle Scholar
Wilson, B. (1995). The ecology of bottlenose dolphin in Moray Firth, Scotland: a population at the northern extreme of the species' range. Ph.D. thesis, Scotland: University of Aberdeen.
Wilson, E. O. (1980). Sociobiology: The Abridged Edition. Cambridge: Belknap Press.Google Scholar
Wilson, M. and Wrangham, R. W. (2003). Intergroup relations in chimpanzees. Annual Review of Anthropology, 32, 363–92.CrossRef
Wilson, R. P., Copper, J. and Plotz, J. (1992). Can we determine when marine endotherms feed? A case study of seabirds. Journal of Experimental Biology, 167, 267–75.Google Scholar
Wilson, M. L., Wallauer, W. R. and Pusey, A. E. (2004). Intergroup violence in chimpanzees: new cases from Gombe National Park, Tanzania. International Journal of Primatology, 25, 523–50.CrossRefGoogle Scholar
Wirtz, P. and Morato, T. (2001). Unequal sex ratios in longline catches. Journal of Marine Biology, 81, 1–2.Google Scholar
Witter, M. S. and Cuthill, I. C. (1993). The ecological costs of avian fat storage. Philosophical Transactions of the Royal Society of London B, 340, 73–90.CrossRefGoogle ScholarPubMed
Wolf, N. G. (1985). Odd fish abandon mixed-species groups when threatened. Behavioral Ecology and Sociobiology, 17, 47–52.CrossRefGoogle Scholar
Wood, A. G., Naef-Daenzer, B., Prince, P. A. and Croxall, J. P. (2000). Quantifying habitat use in satellite-tracked pelagic seabirds: application of kernel estimation to albatross locations. Journal of Avian Biology, 31, 278–86.CrossRefGoogle Scholar
Wood, C. M. and McDonald, D. G. (1997). Global Warming: Implications for Freshwater and Marine Fish. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Woodroffe, R. and Ginsberg, J. R. (2000). Ranging behaviour and vulnerability to extinction in carnivores. In Behavior and Conservation, eds. Gosling, L. M. and Sutherland, W. J.. Cambridge, United Kingdom: Cambridge University Press, pp. 125–40.Google Scholar
Woolington, D. W. (1993). Sex ratios of canvasbacks wintering in Louisiana. Journal of Wildlife Management, 57, 751–8.CrossRefGoogle Scholar
Wooller, R. D., Richardson, K. C., Garavanta, C. A. M., Saffer, V. M. and Bryant, K. A. (2000). Opportunistic breeding in the polyandrous honey possum, Tarsipes rostratus. Australian Journal of Zoology, 48, 669–80.CrossRefGoogle Scholar
Woolnough, A. P. and du Toit, J. T. (2001). Vertical zonation of browse quality in tree canopies exposed to a size-structured guild of African browsing ungulates. Oecologia, 129, 585–90.CrossRefGoogle ScholarPubMed
Wootton, R. J. (1976) The Biology of the Sticklebacks. London: Academic Press.Google Scholar
Wootton, R. J. (1998) Ecology of Teleost Fishes. Dordrecht, The Netherlands: Kluwer Academic Pulishers.Google Scholar
Worrell, E. (1958). Song of the Snake. Sydney: Angus and Robertson.Google Scholar
Wourms, J. P. and Demski, L. S. (1993). The reproduction and development of sharks, skates, rays and ratfishes: introduction, history, overview, and future prospects. Environmental Biology of Fishes, 38, 7–21.CrossRefGoogle Scholar
Wrangham, R. W. (1979). On the evolution of ape social systems. Social Science Information, 18, 335–68.CrossRefGoogle Scholar
Wrangham, R. W. (1999). Evolution of coalitionary killing. Yearbook of Physical Anthropology, 42, 1–30.3.0.CO;2-E>CrossRefGoogle Scholar
Wrangham, R. W. (2000). Why are male chimpanzees more gregarious than mothers? A scramble competition hypothesis. In Primate Males, ed. Kappeler, P. M.. Cambridge: Cambridge University Press, pp. 248–58.Google Scholar
Wrangham, R. W. and Rubenstein, D. I. (1986). Social evolution in birds and mammals. In Ecological Aspects of Social Evolution, ed. Wrangham, R.. Princeton, New Jersey: Princeton University Press, pp. 452–71.Google Scholar
Wrangham, R. W. and Smuts, B. B. (1980). Sex differences in the behavioural ecology of chimpanzees in the Gombe National Park, Tanzania. Journal of Reproduction and Fertility, Suppl. 28, S13–S31.Google ScholarPubMed
Wrangham, R. W., Chapman, C. A., Clark-Arcadi, A. P. and Isabirye-Basuta, G. (1996). Social ecology of Kanyawara chimpanzees: implications for understanding the costs of great ape groups. In Great Ape Societies, eds. McGrew, W. C., Marchant, L. F. and Toshisada, N.. Cambridge: Cambridge University Press, pp. 45–57.CrossRefGoogle Scholar
Wright, P. H. (1988). Interpreting research on gender differences in friendship: A case for moderation and plea for caution. Journal of Social and Personal Relationships, 5(3), 367–73.CrossRefGoogle Scholar
Wright, P. H. and Scanlon, M. B. (1991). Gender role orientations and friendship: some attenuation, but gender differences abound. Sex Roles, 24(9–10), 551–66.CrossRefGoogle Scholar
Wrona, F. J. and , Dixon W. J. (1991). Group size and predation risk: a field analysis of encounter and dilution effects. American Naturalist, 137, 186–201.CrossRefGoogle Scholar
Wunderle, J. M. (1995). Population characteristics of black-throated blue warblers wintering in three sites on Puerto Rico. Auk, 112, 931–46.CrossRefGoogle Scholar
Würsig, B. and Bastida, R. (1986). Long-range movement and individual associations of two dusky dolphins (Lagenorhynchus obscurus) off Argentina. Journal of Mammalogy, 67, 773–4.CrossRefGoogle Scholar
Würsig, B. and Würsig, M. (1980). Behavior and ecology of the Dusky dolphin, Lagenorhynchus obscurus, in the South Atlantic. Fishery Bulletin, 77, 871–90.Google Scholar
Xavier, J. C., Croxall, J. P., Trathan, P. N. and Rodhouse, P. G. (2003a). Inter-annual variation in the cephalopod component of the diet of wandering albatrosses Diomedea exulans breeding at Bird Island, South Georgia. Marine Biology, 142, 611–22.CrossRefGoogle Scholar
Xavier, J. C., Croxall, J. P. and Reid, K. (2003b). Interannual variation in the diets of two albatross species breeding at South Georgia: implications for breeding performance. Ibis, 145, 593–610.CrossRefGoogle Scholar
Xavier, J. C., Croxall, J. P., Trathan, P. N. and Wood, A. G. (2003c). Feeding strategies and diets of breeding grey-headed and wandering albatrosses at South Georgia. Marine Biology, 143, 221–32.CrossRefGoogle Scholar
Xavier, J. C., Trathan, P. N., Croxall, J. P.et al. (2004). Foraging ecology and interactions with fisheries of wandering albatrosses (Diomedea exulans) breeding at South Georgia. Fisheries Oceanography, 13(5), 324–44.CrossRefGoogle Scholar
Yamagiwa, J. (1987). Intra- and inter-group interactions of an all-male group of Virunga mountain gorillas (Gorilla gorilla beringei). Primates, 28, 1–30.CrossRefGoogle Scholar
Yearsley, J. M. and Pérez-Barbería, F. J. (2005). Does the activity budget hypothesis explain sexual segregation in ungulates?Animal Behaviour, 69, 257–67.CrossRefGoogle Scholar
Young, D. D. and Cockcroft, V. G. (1994). Diet of common dolphins (Delphinus delphis) off the south-east of southern Africa: opportunism or specialization?Journal of Zoology, London, 234, 41–53.CrossRefGoogle Scholar
Young, T. P. and Isbell, L. A. (1991). Sex-differences in giraffe feeding ecology – energetic and social constraints. Ethology, 87, 79–89.CrossRefGoogle Scholar
Yurk, H., Barrett-Lennard, L., Ford, J. K. B. and Matkins, C. O. (2002). Cultural transmission within maternal lineages: vocal clans in resident killer whales in southern Alaska. Animal Behaviour, 63, 1103–19.CrossRefGoogle Scholar
Zari, T. A. (1987). The energetics and thermal physiology of the Wiegmann's skink, Mabuya brevicollis. Ph.D. thesis, Nottingham, UK: University of Nottingham.Google Scholar
Zari, T. A. (1998). Effects of sexual condition on food consumption and temperature selection in the herbivorous desert lizard, Uromastyx philbyi. Journal of Arid Environments, 38, 371–7.CrossRefGoogle Scholar
Zharikov, Y. and Skilleter, G. A. (2002). Sex-specific intertidal habitat use in subtropically wintering bar-tailed godwits. Canadian Journal of Zoology, 80, 1918–29.CrossRefGoogle Scholar
Zinner, H. (1985). On behavioral and sexual dimorphism of Telescopus dhara Forscal 1776 (Reptilia: Serpentes, Colubridae). Journal of the Herpetological Association of Africa, 31, 5–6.CrossRefGoogle Scholar
Zuberbühler, J. K. D. and Bshary, R. (1999). The predator deterrence function of primate alarm calls. Ethology, 105, 477–90.CrossRefGoogle Scholar
Zucker, N. (1986). Perch height preferences of male and female tree lizards, Urosaurus ornatus: a matter of food competition or social role?Journal of Herpetology, 20, 547–53.CrossRefGoogle Scholar
Zuk, M. and McKean, K. A. (1996). Sex differences in parasite infections: patterns and processes. International Journal of Parasitology, 26, 1009–24.CrossRefGoogle ScholarPubMed
Zwarts, L. (1988). Numbers and distribution of coastal waders in Guinea-Bissau. Ardea, 76, 42–55.Google Scholar
Abernethy, K. A., White, L. J. T. and Wickings, E. J. (2002). Hordes of mandrills (Mandrillus sphinx): extreme group size and seasonal male presence. Journal of Zoology, London, 258, 131–7.CrossRefGoogle Scholar
Acuna, H. O. and Francis, J. M. (1995). Spring and summer prey of the Juan-Fernandez fur seal, Arctocephalus philippii. Canadian Journal of Zoology, 73, 1444–52.CrossRefGoogle Scholar
Adamopoulou, C. and Legakis, A. (2002). Diet of a lacertid lizard (Podarcis milensis) in an insular dune ecosystem. Israel Journal of Zoology, 48, 207–19.CrossRefGoogle Scholar
Afanasyev, V. (2004). A miniature daylight level and activity data recorder for tracking animals over long periods. Memoirs of the National Institute of Polar Research, Special Issue, 58, 227–233.Google Scholar
Agrimi, U. and Luiselli, L. (1992). Feeding strategies of the viper V. u. ursinii (Reptilia: Viperidae) in the Apennines. Herpetological Journal, 2, 37–42.Google Scholar
Alberts, S. A. and Altmann, J. (1995). Balancing costs and benefits: dispersal in male baboons. American Naturalist, 145, 279–306.CrossRefGoogle Scholar
Albon, S. D., Clutton-Brock, T. H. and Langvatn, R. (1992). Cohort variation in reproduction and survival: implications for population demography. In The Biology of Deer, ed. Brown, R. D.. Berlin: Springer Verlag.CrossRefGoogle Scholar
Alexander, C., Lynch, J. J. and Mottershead, B. E. (1979). Use of shelter and selection of lambing sites by shorn and unshorn ewes in paddocks with closely or widely spaced shelters. Applied Animal Ethology, 5, 51–69.CrossRefGoogle Scholar
Allen, B. A., Burghardt, G. M. and York, D. S. (1984). Species and sex differences in substrate preference and tongue-flick rate in three sympatric species of water snakes (Nerodia). Journal of Comparative Psychology, 98, 358–67.CrossRefGoogle Scholar
Altmann, J. (1980). Baboon Mothers and Infants. Cambridge, MA: Harvard University Press.Google Scholar
Altmann, J. (1990). Primate males go where the females are. Animal Behaviour, 39, 193–5.CrossRefGoogle Scholar
Altringham, J. D. (1996). Bats: Biology and Behaviour. Oxford: Oxford University Press.Google Scholar
Altringham, J. D. and Fenton, M. B. (2003). Sensory ecology and communication in the Chiroptera. In Bat Ecology, eds. Kunz, T. H. and Fenton, M. B.. Chicago: University of Chicago Press, pp. 90–127.Google Scholar
Anderson, D. J., Reeve, J., Gomez, J. E. M.et al. (1993). Sexual size dimorphism and food requirements of nestling birds. Canadian Journal of Zoology, 71, 2541–5.CrossRefGoogle Scholar
Anderson, D. J., Schwandt, A. J. and Douglas, H. D. (1998). Foraging ranges of waved albatrosses in the Eastern tropical Pacific. In Albatross Biology and Conservation, eds. Robertson, G. and Gales, R.. Chipping Norton: Surrey Beatty and Sons, pp. 180–5.Google Scholar
Andersson, M. (1994). Sexual Selection, Princeton: Princeton University Press.Google Scholar
Andrews, R. M. (1971). Structural habitat and time budget of a tropical Anolis lizard. Ecology, 52, 262–70.CrossRefGoogle Scholar
Andrews, R. M., Cruz, F. R. M. and Cruz, Santa M. V. (1997). Body temperatures of female Sceloporus grammicus: Thermal stress or impaired mobility? Copeia, 1997, 108–15.CrossRefGoogle Scholar
Angelici, F. M., Effah, C., Inyang, M. A. and Luiselli, L. (2000). A preliminary radiotracking study of movements, activity patterns and habitat use of free-ranging Gaboon vipers, Bitis gabonica. Terre et la Vie, 55, 45–55.Google Scholar
Anibaldi, C., Luiselli, L. and Angelici, F. M. (1998). Notes on the ecology of a suburban population of rainbow lizards in coastal Kenya. African Journal of Ecology, 36, 199–206.CrossRefGoogle Scholar
Antonelis, G. A., Lowry, M. S., Fiscus, C. H., Stewart, B. S. and Delong, R. L. (1994). Diet of the northern elephant seal. In Elephant Seals: Population Ecology, Behaviour, and Physiology, ed. Boeuf, B. J. and Laws, R. M.. Berkley: University of California Press, pp. 211–23.Google Scholar
Appleby, M. C. (1982). The consequences and causes of high social rank in red deer stags. Behaviour, 80, 259–73.CrossRefGoogle Scholar
Appleby, M. C. (1983). Competition in a red deer stag social group – rank, age and relatedness of opponents. Animal Behaviour, 31, 913–18.CrossRefGoogle Scholar
Arcese, P., Keller, L. F. and Cary, J. R. (1997). Why hire a behaviorist into a conservation or management team? In Behavioral Approaches to Conservation in the Wild, eds. Clemmons, J. R. and Buchholz, R.. Cambridge, United Kingdom: Cambridge University Press, pp. 48–71.Google Scholar
Archer, J. and Lloyd, B. (2002). Sex and Gender. Cambridge, UK: Cambridge University Press.CrossRefGoogle Scholar
Archibald, G. W. and Meine, C. D. (1996). Family Gruidae (Cranes). In Handbook of the Birds of the World, Vol. 3., eds. Hoyo, J. del, Elliot, A. and Sargatal, J.. Barcelona: Lynx Edicions.Google Scholar
Ardia, D. R. and Bildstein, K. L. (1997). Sex-related differences in habitat selection in wintering American kestrels, Falco sparverius. Animal Behaviour, 53, 1305–11.CrossRefGoogle ScholarPubMed
Ardia, D. R. and Bildstein, K. L. (2001). Sex-related differences in habitat use in wintering American kestrels. Auk, 118, 746–50.CrossRefGoogle Scholar
Arkhipkin, A. I. and Middletion, D. A. J. (2002). Sexual segregation in ontogenetic migrations by the squid Loligo gahi around the Falkland Islands. Bulletin of Marine Science, 71, 109–27.Google Scholar
Arlettaz, R. (1999). Habitat selection as a major resource partitioning mechanism between the two sympatric sibling bat species Myotis myotis and Myotis blythii. Journal of Animal Ecology, 68, 460–71.CrossRefGoogle Scholar
Arnbom, T. and Lundberg, S. (1995). Notes on Lepas australis (Cirripedia, Lepadidae) recorded on the skin of southern elephant seal (Mirounga leonina). Crustaceana, 68, 655–8.CrossRefGoogle Scholar
Arnbom, T., Papastavrou, V., Weilgart, L. S. and Whitehead, H. (1987). Sperm whales react to an attack by killer whale. Journal of Mammalogy, 68, 450–3.CrossRefGoogle Scholar
Arnold, G. W., Boundy, C. A. P., Morgan, P. D. and Bartle, G. (1975). The roles of sight and hearing in the lamb in the location and discrimination between ewes. Applied Animal Ethology, 5, 43–50.CrossRefGoogle Scholar
Arnold, G. W., Steven, D. E. and Grassia, A. (1990). Associations between individuals and classes in groups of different size in a population of western grey kangaroos, Macropus fuliginosus. Australian Wildlife Research, 17, 551–62.CrossRefGoogle Scholar
Arnold, G. W., Steven, D. E. and Weeldenberg, J. R. (1994). Comparative ecology of western grey kangaroos (Macropus fuliginosus) and Euros (M. robustus erubescens) in Durokoppin Nature Reserve, isolated in the central wheatbelt of Western Australia. Wildlife Research, 21, 307–22.CrossRefGoogle Scholar
Arnold, P. W. (1972). Predation on harbour porpoise, Phocoena phocoena, by a white shark, Carcharodon carcharias. Journal of Fisheries Research Board of Canada, 29, 1213–14.CrossRefGoogle Scholar
Arnold, S. J. (1993). Foraging theory and prey-size-predator-size relations in snakes. In Snakes: Ecology and Behavior, eds. Seigel, R. A. and Collins, J. T.. New York: McGraw-Hill, pp. 87–116.Google Scholar
Arnold, S. J. and Peterson, C. R. (2002). A model for optimal reaction norms: the case of the pregnant garter snake and her temperature-sensitive embryos. American Naturalist, 160, 306–16.Google ScholarPubMed
Arnould, J. P. Y. and Duck, C. D. (1997). The cost and benefits of territorial tenure, and factors affecting mating success in male Antarctic fur seals. Journal of Zoology, 241, 649–64.CrossRefGoogle Scholar
Arnould, J. P. Y., Green, J. A. and Rawlins, D. R. (2001). Fasting metabolism in Antarctic fur seal (Arctocephalus gazella) pups. Comparative Biochemistry and Physiology. Part A, Molecular and Integrative Physiology, 129, 829–41.CrossRefGoogle ScholarPubMed
Atsalis, K. (1999). The diet of the brown mouse lemur, Microcebus rufus, in Ranomafana National Park, Madagascar. International Journal of Primatology, 20, 193–229.CrossRefGoogle Scholar
Backus, R. H., Springer, S. and Arnold, E. L. (1956). A contribution to the natural history of the white-tip shark, Pterolamiops longimanus (Poey). Deep-Sea Research, 3, 178–88.CrossRefGoogle Scholar
Badyaev, A. V. (2002). Growing apart: an ontogenetic perspective on the evolution of sexual size dimorphism. Trends in Ecology and Evolution, 17, 369–78.CrossRefGoogle Scholar
Baird, R. W. (2000). The killer whale: Foraging specializations and group hunting. In Cetacean Societies: Field Studies of Dolphins and Whales, eds. Mann, J., Connor, R. C., Tyack, P. L. and Whitehead, H.. Chicago: Academic Press, pp. 127–53.Google Scholar
Baird, R. W. and Whitehead, H. (2000). Social organization of mammal-eating killer whales: group stability and dispersal patterns. Canadian Journal of Zoology, 78, 2096–105.CrossRefGoogle Scholar
Baker, J. R. and McCann, T. S. (1989). Pathology and bacteriology of adult male Antarctic fur seals, Arctocephalus gazella, dying at Bird Island, South Georgia. British Veterinary Journal, 145, 263–75.CrossRefGoogle ScholarPubMed
Baldassarre, G. A. and Bolen, E. G. (1994). Waterfowl Ecology and Management. New York: John Wiley and Sons.Google Scholar
Baldellou, M. and Aden, A. (1997). Time, gender, and seasonality in vervet activity: a chronobiological approach. Primates, 38, 33–43.CrossRefGoogle Scholar
Barboza, P. S. and Bowyer, R. T. (2000). Sexual segregation in dimorphic deer: a new gastrocentric hypothesis. Journal of Mammalogy, 81, 473–89.2.0.CO;2>CrossRefGoogle Scholar
Barclay, R. M. R. (1991). Population structure of temperate zone insectivorous bats in relation to foraging behaviour and energy demand. Journal of Animal Ecology, 60, 165–78.CrossRefGoogle Scholar
Barclay, R. M. R. and Harder, L. D. (2003). Life histories of bats: life in the slow lane. In Bat Ecology, eds. Kunz, T. H. and Fenton, M. B.. Chicago: University of Chicago Press, pp. 209–53.Google Scholar
Barrette, C. (1991). The size of Axis deer fluid groups in Wilpattu National Park, Sri Lanka. Mammalia, 2, 207–20.Google Scholar
Barros, N. B. and Odell, D. K. (1990). Food habits of bottlenose dolphins in Southeastern United States. In The Bottlenose Dolphin, eds. Leatherwood, S. and Reeves, R. R.. San Diego: Academic Press, pp. 309–28.Google Scholar
Barry, R. E., Botje, M. A. and Grantham, L. B. (1984). Vertical stratification of Peromyscus leucopus and P. maniculatus in southwestern Virginia. Journal of Mammalogy, 65, 145–8.CrossRefGoogle Scholar
Bartle, J. A. (1990). Sexual segregation of foraging zones in procellariiform birds: implications of accidental capture on commercial fishery longlines of grey petrels (Procellaria cinerea). Notornis, 37, 146–50.Google Scholar
Bass, A. J., D'Aubrey, J. and Kistnasamy, N. (1973). Sharks of the east coast of southern Africa. I. The genus Carcharhinus (Carcharhinidae). Investigation Report of the Oceanographic Research Institute, 33, 1–168.Google Scholar
Bateson, P. P. G. and Martin, P. (1999). Design for a Life: How Behaviour Develops. London: Jonathan Cape.Google Scholar
Bauwens, D. and Thoen, C. (1981). Escape tactics and vulnerability to predation associated with reproduction in the lizard Lacerta vivipara. Journal of Animal Ecology, 50, 733–43.CrossRefGoogle Scholar
Beamish, F. W. H. (1978). Swimming capacity of fish. In Fish Physiology, Vol. 7, ed. Randall, D. J.. New York: Academic Press, pp. 101–87.Google Scholar
Bearder, S. K. (1987). Lorises, bushbabies, and tarsiers: diverse societies in solitary foragers. In Primate Societies, eds. Smuts, B. B., Cheney, D. L., Seyfarth, R. M., Wrangham, R. W., and Struhsaker, T. T.. Chicago: University of Chicago Press, pp. 11–24.Google Scholar
Beck, C. A., Bowen, W. D. and Iverson, S. J. (2003a). Sex differences in the seasonal patterns of energy storage and expenditure in a phocid seal. Journal of Animal Ecology, 72, 280–91.CrossRefGoogle Scholar
Beck, C. A., Bowen, W. D., McMillan, J. I. and Iverson, S. J. (2003b). Sex differences in the diving behaviour of a size-dimorphic capital breeder: the grey seal. Animal Behaviour, 66, 777–89.CrossRefGoogle Scholar
Becker, C. D. and Ginsberg, J. R. (1990). Mother-infant behavior of wild Grevy's zebra: adaptation for survival in semi-desert East Africa. Animal Behaviour, 40, 1111–18.CrossRefGoogle Scholar
Becker, D. S. (1988). Relationship between sediment character and sex segregation in English sole, Parophrys vetulus. U.S. Fishery Bulletin, 86, 517–24.Google Scholar
Becker, P. H., González-Solís, J., Behrends, B. and Croxall, J. P. (2002). Feather mercury levels in seabirds at South Georgia: Influence of trophic position, sex and age. Marine Ecology Progress Series, 243, 261–9.CrossRefGoogle Scholar
Beer, J. R. and Richards, A. G. (1956). Hibernation of the big brown bat. Journal of Mammalogy, 37, 31–41.CrossRefGoogle Scholar
Beier, P. (1987). Sex differences in quality of white-tailed deer diets. Journal of Mammalogy, 68, 323–9.CrossRefGoogle Scholar
Beier, P. and McCullough, D. R. (1990). Factors influencing white-tailed deer activity patterns and habitat use. Wildlife Monographs, 109, 1–51.Google Scholar
Beissinger, S. R. (1997). Integrating behavior into conservation biology: potential and limitations. In Behavioral Approaches to Conservation in the Wild, eds. Clemmons, J. R. and Buchholz, R.. Cambridge, United Kingdom: Cambridge University Press, pp. 23–47.Google Scholar
Béland, P., Faucher, A. and Corbeil, P. (1990). Observations on the birth of a beluga whale (Delphinapterus leucas) in the St. Lawrence Estuary, Quebec, Canada. Canadian Journal of Zoology, 68, 1327–9.CrossRefGoogle Scholar
Belcher, C. A. (2003). Demographics of tiger quoll (Dasyurus maculatus maculatus) populations in south-eastern Australia. Australian Journal of Zoology, 51, 611–26.CrossRefGoogle Scholar
Bell, R. H. V. (1969). The use of the herbaceous layer by grazing ungulates in the Serengeti National Park, Tanzania. Ph.D. thesis, University of Manchester.
Bell, R. H. V. (1971). A grazing ecosystem in the Serengeti. Scientific American, 225, 86–93.CrossRefGoogle Scholar
Benham, P. F. J. (1982). Synchronization of behaviour in grazing cattle. Applied Animal Ethology, 8, 403–4.CrossRefGoogle Scholar
Bennett, P. M. and Owens, P. F. (2002). Evolutionary Ecology of Birds. Oxford: Oxford University Press.Google Scholar
Benshemesh, J. and Johnson, K. (2003). Biology and conservation of marsupial moles (Notoryctes). In Predators with Pouches: The Biology of Carnivorous Marsupials, eds. Jones, M. E., Dickman, C. and Archer, M.. Melbourne: CSIRO Publishing, pp. 464–74.Google Scholar
Bercovitch, F. B. (1983). Time budgets and consorts in olive baboons (Papio anubis). Folia Primatologica, 41, 180–90.CrossRefGoogle Scholar
Bercovitch, F. B. (1997). Reproductive strategies of rhesus macaques. Primates, 38, 247–63.CrossRefGoogle Scholar
Bergan, J. F. and Smith, L. M. (1989). Differential habitat use by diving ducks wintering in South Carolina. Journal of Wildlife Management, 53, 1117–26.CrossRefGoogle Scholar
Berger, J. (1986). Wild Horses of the Great Basin. Chicago: University of Chicago Press.Google Scholar
Berger, J. (1991). Pregnancy incentives, predation constraints, and habitat shifts: experimental and field evidence for wild bighorn sheep. Animal Behaviour, 41, 66–77.CrossRefGoogle Scholar
Berger, J. and Cunningham, C. (1988). Size-related effects on search times in North American grassland female ungulates. Ecology, 69, 177–83.CrossRefGoogle Scholar
Berger, J. and Cunningham, C. (1995). Predation, sensitivity, and sex: why female black rhinoceroses outlive males. Behavioral Ecology, 6, 57–64.CrossRefGoogle Scholar
Berger, J. and Gompper, M. E. (1999). Sex ratios in extant ungulates: products of contemporary predation or past life histories? Journal of Mammalogy, 80, 1084–113.CrossRefGoogle Scholar
Berger, J., Swenson, J. E. and Persson, I.-L. (2001). Recolonizing carnivores and naïve prey: conservation lessons from Pleistocene extinctions. Science, 291, 1036–9.CrossRefGoogle ScholarPubMed
Bergerud, A. T. (1974). Rutting behaviour of Newfoundland caribou. In The Behaviour of Ungulates and its Relationship to Management, eds. V. Giest and F. Walther, pp. 395–435. International Union for Conservation of Nature and Natural Resources: New Series, 24, 1–940.
Bergerud, A. T., Butler, H. E. and Miller, D. R. (1984). Antipredator tactics of calving caribou: dispersion in mountains. Canadian Journal of Zoology, 62, 1566–75.CrossRefGoogle Scholar
Bernard, H. J. and Hohn, A. A. (1989). Differences in feeding habits between pregnant and lactating spotted dolphins (Stenella attenuata). Journal of Mammalogy, 70, 211–15.CrossRefGoogle Scholar
Bernstein, N. P. and Maxson, S. J. (1984). Sexually distinct activity patterns of blue-eyed shags in Antarctica. Condor, 86, 151–6.CrossRefGoogle Scholar
Berrow, S. D. and Croxall, J. P. (2001). Provisioning rate and attendance patterns of wandering albatrosses at Bird Island, South Georgia. Condor, 103, 230–9.CrossRefGoogle Scholar
Berrow, S. D., Humpidge, R. and Croxall, J. P. (2000). Influence of adult breeding experience on growth and provisioning of wandering albatross Diomedea exulans chicks at South Georgia. Ibis, 142, 199–207.CrossRefGoogle Scholar
Berry, J. F. and Shine, R. (1980). Sexual size dimorphism and sexual selection in turtles (Order Chelonia). Oecologia, 44, 185–91.CrossRefGoogle Scholar
Berta, A. (2002). Pinnipedia, overview. In Encyclopedia of Marine Mammals, eds. Perrin, W. F., Wursig, B. and Thewissen, J. G. M.. San Diego: Academic Press, pp. 931–9.Google Scholar
Berteaux, D. (1993). Female-biased mortality in a sexually dimorphic ungulate: feral cattle of Amsterdam Island. Journal of Mammalogy, 74, 732–7.CrossRefGoogle Scholar
Best, P. B. (1979). Social organization in sperm whales, Physeter macrocephalus. In Behavior of Marine Mammal: Current Perspectives in Research, vol. 3, Cetaceans, eds. Winn, H. E. and Olla, B. L.. New York: Plenum Press, pp. 227–89.Google Scholar
Best, R. C. and da Silva, V. M. F. (1993). Inia geoffrensis. Mammalian Species, 423, 1–8.Google Scholar
Best, T. L. and Gennaro, A. L. (1984). Feeding ecology of the lizard, Uta stansburiana, in southwestern Mexico. Journal of Herpetology, 18, 291–301.CrossRefGoogle Scholar
Best, T. L. and Pfaffenberger, G. S. (1987). Age and sexual variation in the diet of collared lizards (Crotaphytus collaris). Southwestern Naturalist, 32, 415–26.CrossRefGoogle Scholar
Beuchat, C. A. (1986). Reproductive influences on the thermoregulatory behavior of a live-bearing lizard. Copeia, 1986, 971–9.CrossRefGoogle Scholar
Bibby, C. J. and Thomas, D. K. (1984). Sexual dimorphism in size, moult and movements of Cetti's warbler Cettia cetti. Bird Study, 31, 28–34.CrossRefGoogle Scholar
Bigg, M. A. (1973). Census of California sea lions on southern Vancouver Island, British Columbia. Journal of Mammal Research, 54, 285–7.CrossRefGoogle Scholar
Bigg, M. A. (1990). Migration of northern for seals (Callorhinus ursinus) off western North America. Canadian Technical Report of Fisheries and Aquatic Science, 1764, 1–64.Google Scholar
Bigg, M. A., Olesiuk, P. F., Ellis, G. M., Ford, J. K. B. and Balcomb, K. C. III (1990). Social organization and genealogy of resident killer whales (Orcinus orca) in the coastal waters of British Columbia and Washington State. Report of the International Whaling Commission, Special issue, 12, 383–405.Google Scholar
BirdLife International (2000). Threatened Birds of the World. Barcelona and Cambridge, UK: Lynx edicions and BirdLife International.
Blackburn, D. G. (1985). Evolutionary origins of viviparity in the Reptilia. II. Serpentes, Amphisbaenia, and Icthyosauria. Amphibia-Reptilia, 6, 259–91.CrossRefGoogle Scholar
Blatchford, P., Baines, , and Pellegrini, A. D. (2003). The social context of school playground games: sex and ethnic differences and changes over time after entry into junior school. British Journal of Developmental Psychology, 21, 459–71.CrossRefGoogle Scholar
Blaxter, J. H. S. and Holliday, F. G. T. (1969). The behaviour and physiology of herring and other clupeids. Advances in Marine Biology, 1, 261–393.CrossRefGoogle Scholar
Bleich, V. C., Bowyer, R. T. and Wehausen, J. D. (1997). Sexual segregation in mountain sheep: resources or predation? Wildlife Monographs, 134, 1–50.Google Scholar
Bleich, V. C., Wehausen, J. D. and Holl, S. A. (1990). Desert-dwelling mountain sheep: conservation implications of a naturally fragmented distribution. Conservation Biology, 4, 383–90.CrossRefGoogle Scholar
Bleich, V. C., Wehausen, J. D., Ramey, R. R. and Rechel, J. L. (1996). Metapopulation theory and mountain sheep: implications for conservation. In Metapopulations and Wildlife Conservation, ed. McCullough, D. R.. Covelo, California: Island Press, pp. 353–73.Google Scholar
Bleske, A. L. and Buss, D. M. (2000). Can men and women just be friends? Personal Relationships, 7, 131–51.CrossRefGoogle Scholar
Bleske-Rechek, A. L. and Buss, D. M. (2001). Opposite-sex friendship: sex differences and similarities in initiation, selection, and interpretation. Personality and Social Psychology Bulletin, 27, 1310–23.CrossRefGoogle Scholar
Blondel, J., Perret, P., Anstett, M.-C. and Thébaud, C. (2002). Evolution of sexual size dimorphism in birds: test of hypotheses using blue tits in contrasted Mediterranean habitats. Journal of Evolutionary Biology, 15, 440–50.CrossRefGoogle Scholar
Blouin-Demers, G. and Weatherhead, P. J. (2001). Thermal ecology of black rat snakes (Elaphe obsoleta) in a thermally challenging environment. Ecology, 82, 3025–43.CrossRefGoogle Scholar
Boase, H. (1926). Proportion of male and female duck on Tay estuary. British Birds, 20, 169–72.Google Scholar
Boesch, C. and Boesch-Achermann, H. (2000). The Chimpanzees of the Tai Forest. Oxford: Oxford University Press.Google Scholar
Boinski, S. (1987). Birth synchrony in squirrel monkeys: a strategy to reduce neonatal predation. Behavioral Ecology and Sociobiology, 21, 393–400.CrossRefGoogle Scholar
Boinski, S. (1988). Sex differences in the foraging behavior of squirrel monkeys. Behavioral Ecology and Sociobiology, 23, 177–86.CrossRefGoogle Scholar
Boinski, S. (1994). Affiliation patterns among Costa Rican squirrel monkeys. Behaviour, 130, 191–209.CrossRefGoogle Scholar
Boinski, S. (1999). The social organizations of squirrel monkeys: implications for ecological models of social evolution. Evolutionary Anthropology, 8, 101–12.3.0.CO;2-O>CrossRefGoogle Scholar
Boinski, S., Sughrue, K., Selvaggi, L.et al. (2002). An expanded test of the ecological model of primate social evolution: competitive regimes and female bonding in three species of squirrel monkeys (Saimiri oerstedi, Saimiri boliviensis, and Saimiri sciureus). Behaviour, 139, 227–61.CrossRefGoogle Scholar
Bon, R. (1991). Social and spatial segregation of males and females in polygamous ungulates: proximate factors. In Ongulés/Ungulates 1991, eds. Spitz, F., Janeau, G., Gonzalez, G., and Aulagnier, S.. Paris/Toulouse: SFEPM-IRGM, pp. 195–8.Google Scholar
Bon, R. (1998). Perspective éthologique de l'étude de la ségrégation sexuelle chez les ongulés. Habilitation à Diriger les Recherches. Université Paul Sabatier, Toulouse.
Bon, R. and Campan, R. (1989). Social tendencies of the Corsican mouflon Ovis ammon musimon in the Caroux-Espinouse (South of France). Behavioural Processes, 19, 57–78.CrossRefGoogle Scholar
Bon, R. and Campan, R. (1996). Unexplained sexual segregation in polygamous ungulates: a defense of an ontogenetic approach. Behavioural Processes, 38, 131–54.CrossRefGoogle ScholarPubMed
Bon, R., Dubois, M. and Maublanc, M. L. (1993). Does age influence between-rams companionship in mouflon (Ovis gmelini)? Revue d'Ecologie (Terre vie), 47, 57–64.Google Scholar
Bon, R., Joachim, J. and Maublanc, M. L. (1995). Do lambs affect feeding habitat use of lactating female mouflons in spring within an area free of predators? Journal of Zoology, London, 235, 43–51.CrossRefGoogle Scholar
, Bon R., Rideau, C., Villaret, J.-C. and Joachim, J. (2001). Segregation is not only a matter of sex in Alpine ibex (Capra ibex ibex). Animal Behaviour, 62, 495–504.Google Scholar
Boness, D. J., Clapham, P. J. and Mesnick, S. L. (2002). Life history and reproductive strategies. In Marine Mammal Biology: An Evolutionary Approach, ed. Hoelzel, A. R.. Oxford: Blackwell Science, pp. 278–324.Google Scholar
Bonner, W. N. (1968). The fur seal of South Georgia. British Antarctic Survey Scientific Report, 56, 81.Google Scholar
Bonnet, X., Guy, N. and Shine, R. (1999a). The dangers of leaving home: dispersal and mortality in snakes. Biological Conservation, 89, 39–50.CrossRefGoogle Scholar
Bonnet, X., Naulleau, G., Shine, R. and Lourdais, O. (1999b). What is the appropriate timescale for measuring costs of reproduction in a ‘capital breeder’ such as the aspic viper? Evolutionary Ecology, 13, 485–97.CrossRefGoogle Scholar
Bonnet, X., Shine, R., Naulleau, G. and Thiburce, C. (2001). Plastic vipers: genetic and environmental influences on the size and shape of Gaboon vipers, Bitis gabonica. Journal of Zoology, London, 255, 341–51.CrossRefGoogle Scholar
Booth, J. and Peters, J. A. (1972). Behavioural studies on the green turtle (Chelonia mydas) in the sea. Animal Behaviour, 20, 808–12.CrossRefGoogle Scholar
Bourne, W. R. P. and Warham, J. (1966). Geographical variation in the giant petrels of the genus Macronectes. Ardea, 54, 45–67.Google Scholar
Bowen, W. D., Oftedal, O. T. and Boness, D. J. (1985). Birth to weaning in 4 days: remarkable growth in the hooded seal, Cystophora cristata. Canadian Journal of Zoology, 63, 2841–6.CrossRefGoogle Scholar
Bowen, W. D., Read, A. J. and Estes, J. A. (2002). Feeding ecology. In Marine Mammal Biology: An Evolutionary Approach, ed. Hoelzel, A. R.. Oxford: Blackwell Science, pp. 217–46.Google Scholar
Bowers, M. A. (1994). Dynamics of age- and habitat-structured populations. Oikos, 69, 327–33.CrossRefGoogle Scholar
Bowman, J., Jaeger, J. A. G. and Fahrig, L. (2002). Dispersal distance of mammals is proportional to home range size. Ecology, 83, 2049–55.CrossRefGoogle Scholar
Bowyer, R. T. (1984). Sexual segregation in southern mule deer. Journal of Mammalogy, 65, 410–17.CrossRefGoogle Scholar
, Bowyer R. T., Kie, J. G. and Ballengerghe, V. (1996). Sexual segregation in black-tailed deer: effects of scale. Journal of Wildlife Management, 60, 10–17.Google Scholar
Bowyer, R. T., Pierce, B. M., Duffy, L. K. and Haggstrom, D. A. (2001). Sexual segregation in moose: effects of habitat manipulation. Alces, 37, 109–22.Google Scholar
Boyd, I. L. (1993a). Pup production and distribution of breeding Antarctic fur seals Arctocephalus gazella at South Georgia. Antarctic Science, 5, 17–24.CrossRefGoogle Scholar
Boyd, I. L. (1993b). Trends in marine mammal science. In Marine Mammals: Advances in Behavioural and Population Biology, ed. Boyd, I. L.. Oxford: Clarendon Press, pp. 1–12.Google Scholar
Boyd, I. L. and Croxall, J. P. (1996). Dive durations in pinnipeds and seabirds. Canadian Journal of Zoology, 74, 1696–1705.CrossRefGoogle Scholar
Boyd, I. L., Arnbom, T. A. and Fedak, M. A. (1994). Biomass and energy consumption of the South Georgia population of southern elephant seals. In Elephant Seals: Population Ecology, Behaviour, and Physiology, eds. Boeuf, B. J. and Laws, R. M.. Berkley: University of California Press.Google Scholar
Boyd, I. L., Lockyer, C. and Marsh, H. D. (1996). Reproduction in marine mammals. In Biology of Marine Mammals, eds. Reynolds, J. E. and Rommel, S. A.. Washington: Smithsonian Institution Press.Google Scholar
Boyd, I. L., McCafferty, D. J., Reid, K., Taylor, R. and Walker, T. R. (1998). Dispersal of male and female Antarctic fur seals Arctocephalus gazella. Canadian Journal of Fisheries and Aquatic Sciences, 55, 845–52.CrossRefGoogle Scholar
Boyd, I. L., Staniland, I. J. and Martin, A. R. (2002). Distribution of foraging by female Antarctic fur seals. Marine Ecology, Progress Series, 242, 285–94.CrossRefGoogle Scholar
Bradbury, J. W. (1977). Lek mating behaviour in the hammer-headed bat. Zeitschrift für Tierpsychologie, 45, 225–55.CrossRefGoogle Scholar
Bradbury, J. W. and Vehrencamp, S. L. (1976a). Social organisation and foraging in emballonurid bats. I. Field studies. Behavioral Ecology and Sociobiology, 1, 337–81.CrossRefGoogle Scholar
Bradbury, J. W. and Vehrencamp, S. L. (1976b). Social organisation and foraging in emballonurid bats. II. A model for the determination of group size. Behavioral Ecology and Sociobiology, 1, 383–404.CrossRefGoogle Scholar
Bradbury, J. W. and Vehrencamp, S. L. (1977). Social organisation and foraging in emballonurid bats. III. Mating systems. Behavioral Ecology and Sociobiology, 2, 1–17.CrossRefGoogle Scholar
Bradshaw, S. D. and Bradshaw, F. J. (2002). Short-term movements and habitat use of the marsupial honey possum (Tarsipes rostratus). Journal of Zoology, 258, 343–8.CrossRefGoogle Scholar
Brana, F. (1993). Shifts in body temperature and escape behaviour of female Podarcis muralis during pregnancy. Oikos, 66, 216–22.CrossRefGoogle Scholar
Brana, F. (1996). Sexual dimorphism in lacertid lizards: male head increase vs. female abdomen increase? Oikos, 75, 511–23.CrossRefGoogle Scholar
Breitwisch, R. (1989). Mortality patterns, sex ratios and parental investment in monogamous birds. In Current Ornithology 6, ed. Power, D.. New York and London: Plenum Press.CrossRefGoogle Scholar
Bried, J. and Jouventin, P. (2002). Site and mate choice in seabirds: an evolutionary approach. In Biology of Marine Birds, eds. Schreiber, E. A. and Burger, J.. Boca Raton, Florida: CRC Press, pp. 263–305.Google Scholar
Brodie, E. D. I. (1989). Behavioural modification as a means of reducing the cost of reproduction. American Naturalist, 134, 225–38.CrossRefGoogle Scholar
Brodie, P. F. (1971). A reconsideration of aspects of growth, reproduction, and behavior of the white whale (Delphinapterus leucas), with reference to the Cumberland Sound, Baffin Island, population. Journal of the Fisheries Research Board of Canada, 28, 1309–18.CrossRefGoogle Scholar
Brooke, A. P. (1990). Tent selection, roosting ecology and social organization of the tent-making bat, Ectophylla alba, in Costa Rica. Journal of Zoology, London, 221, 11–19.CrossRefGoogle Scholar
Brooks, G. R. and Mitchell, J. C. (1989). Predator-prey size relations in three species of lizard from Sonora, Mexico. Southwestern Naturalist, 34, 541–6.CrossRefGoogle Scholar
Broome, L. (2001). Density, home range, seasonal movements and habitat use of the mountain pygmy-possum Burramys parvus (Marsupialia: Burramyidae) at Mount Blue Cow, Kosciuszko National Park. Austral Ecology, 26, 275–92.CrossRefGoogle Scholar
Brothers, N. (1990). Albatross mortality and associated bait loss in the Japanese longline fishery in the Southern Ocean. Biological Conservation, 55, 255–68.CrossRefGoogle Scholar
Brothers, N., Gales, R. and Reid, T. (1999). The influence of environmental variables and mitigation measures on seabird catch rates in the Japanese tuna longline fishery within the Australian Fishing Zone, 1991–1995. Biological Conservation, 88, 85–101.CrossRefGoogle Scholar
Brown, D. R., Strong, C. M. and Stouffer, P. C. (2002). Demographic effects of habitat selection by hermit thrushes wintering in a pine plantation landscape. Journal of Wildlife Management, 66, 407–16.CrossRefGoogle Scholar
Brown, G. P. and Brooks, R. J. (1993). Sexual and seasonal differences in activity in a northern population of snapping turtles, Chelydra serpentina. Herpetologica, 49, 311–18.Google Scholar
Brown, G. P. and Shine, R. (2004). Effects of reproduction on the antipredator tactics of snakes (Tropidonophis mairii, Colubridae). Behavioral Ecology and Sociobiology, 56, 257–62.CrossRefGoogle Scholar
Brown, G. P. and Weatherhead, P. J. (2000). Thermal ecology and sexual size dimorphism in northern water snakes, Nerodia sipedon. Ecological Monographs, 70, 311–30.CrossRefGoogle Scholar
Brownell Jr, R. L. and Cipriano, F. (1989). Dusky dolphin – Lagenorhynchus obscurus (Gray, 1828). In Handbook of Marine Mammals: The Second Book of Dolphins and the Porpoises, vol. 6, eds. Ridgway, S. H. and Harrison, R.. London: Academic Press, pp. 85–104.Google Scholar
Bubenik, A. B. (1984). Ernaehrung, Verhalten und Umwelt des Schalenwildes. München: BLV Verlagsgesellschaft.Google Scholar
Bull, J. J. (1983). Evolution of Sex Determining Mechanisms. Menlo Park, California: Benjamin/Cummings Publ. Co.Google Scholar
Bull, J. J. and Charnov, E. L. (1989). Enigmatic reptilian sex ratios. Evolution, 43, 1561–6.CrossRefGoogle ScholarPubMed
Bullis, H. R. (1967). Depth segregations and distribution of sex-maturity groups in the marbled catshark, Galeus arae. In Sharks, Skates and Rays, eds. Gilbert, P. W., Mathewson, R. F. and Rall, D. P.. Baltimore, MD: Johns Hopkins University Press, pp. 141–8.Google Scholar
Bunnell, F. L. and Gillingham, M. P. (1985). Foraging behavior: dynamics of dining out. In Bioenergetics of Wild Herbivores, eds. Hudson, R. J. and White, R. G.. Boca Raton: CRC Press, Inc., pp. 53–75.Google Scholar
Bunnell, F. L. and Harestad, A. S. (1983). Dispersal and dispersion of black-tailed deer: models and observations. Journal of Mammalogy, 64, 201–9.CrossRefGoogle Scholar
Burger, A. E. (1991). Maximum diving depths and underwater foraging in alcids and penguins. In Studies of High Latitude Seabirds I. Behavioural, Energetic and Oceanographic Aspects of Seabird Feeding Ecology, eds. W. A. Montevecchi A. J. Gaston. Canadian Wildlife Service Occasional Papers, pp. 9–15.
Burgman, M. A., Ferson, S. and Akcakaya, H. R. (1993). Risk Assessment in Conservation Biology. London: Chapman and Hall.Google Scholar
Burke, R. L. and Mercurio, R. J. (2002). Food habits of a New York population of Italian wall lizards, Podarcis sicula (Reptilia, Lacertidae). American Midland Naturalist, 147, 368–75.CrossRefGoogle Scholar
Burland, T. M., Barratt, E. M., Beaumont, M. A. and Racey, P. A. (1999). Population genetic structure and gene flow in a gleaning bat, Plecotus auritus. Proceedings of the Royal Society, London B, 266, 975–80.CrossRefGoogle Scholar
Burland, T. M., Barratt, E. M., Nichols, R. A. and Racey, P. A. (2001). Mating patterns, relatedness and the basis of natal philopatry in the brown long-eared bat, Plecotus auritus. Molecular Ecology, 10, 1309–21.CrossRefGoogle ScholarPubMed
Burns, J. M., Trumble, S. J., Castellini, M. A. and Testa, J. W. (1998). The diet of Weddell seals in McMurdo Sound, Antarctica as determined from scat collections and stable isotope analysis. Polar Biology, 19, 272–82.CrossRefGoogle Scholar
Bury, R. B. (1972). Habits and home range of the Pacific pond turtle, Clemmys marmorata, in a stream community. Ph.D. Thesis, Berkeley, CA: University of California.Google Scholar
Bury, R. B. (1986). Feeding ecology of the turtle, Clemmys marmorata. Journal of Herpetology, 20, 515–21.CrossRefGoogle Scholar
Busack, S. D. and Jaksic, F. M. (1982). Autecological observations of Acanthodactylus erythrurus (Sauria: Lacertidae) in southern Spain. Amphibia-Reptilia, 3, 237–55.CrossRefGoogle Scholar
Busse, C. D. (1984). Spatial structure of chacma baboon groups. International Journal of Primatology, 5, 247–61.CrossRefGoogle Scholar
Butler, P. J. and Taylor, E. W. (1975). The effect of progressive hypoxia on respiration in the dogfish (Scyliorhinus canicula) at different seasonal temperatures. Journal of Experimental Biology, 63, 117–30.Google ScholarPubMed
Buttemer, W. A. and Dawson, W. R. (1993). Temporal pattern of foraging and microhabitat use by Galapagos marine iguanas, Amblyrhynchus cristatus. Oecologia, 96, 56–64.CrossRefGoogle ScholarPubMed
Byers, J. A. (1997a). American Pronghorn: Social Adaptations and the Ghosts of Predators Past. Chicago, University of Chicago Press.Google Scholar
Byers, J. A. (1997b). Mortality risk to young pronghorns from handling. Journal of Mammalogy, 78, 894–9.CrossRefGoogle Scholar
Byers, J. A. and Walker, C. (1995). Refining the motor training hypothesis for the evolution of play. American Naturalist, 146, 25–40.CrossRefGoogle Scholar
Caister, L. E., Shields, W. M. and Gosser, A. (2003). Female tannin avoidance: a possible explanation for habitat and dietary segregation of giraffes (Giraffa camelopardalis peralta) in Niger. African Journal of Ecology, 41, 201–10.CrossRefGoogle Scholar
Cameron, E., Linklater, W. L., Stafford, K. J. and Minot, E. O. (2003). Social grouping and maternal behaviour in feral horses (Equus caballus): the influence of males on maternal protectiveness. Behavioral Ecology and Sociobiology, 53, 92–101.Google Scholar
Cameron, E. Z. and du Toit, J. T. (2005). Social influences on vigilance behaviour in giraffes (Giraffa camelopardalis). Animal Behaviour 69, 1337–44.CrossRefGoogle Scholar
Camilleri, C. and Shine, R. (1990). Sexual dimorphism and dietary divergence: differences in trophic morphology between male and female snakes. Copeia, 1990, 649–58.CrossRefGoogle Scholar
Campagna, C., Werner, R., Karesh, W.et al. (2001). Movements and location at sea of South American sea lions (Otaria flavescens). Journal of Zoology, 2, 205–20.CrossRefGoogle Scholar
Campbell, A. (1999). Staying alive: evolution, culture, and women's intrasexual aggression. Behavioral and Brain Sciences, 22, 203–52.CrossRefGoogle ScholarPubMed
Campbell, A., Shirley, L., Heywood, C. and Crook, C. (2000). Infants' visual preference for sex-congruent babies, children, toys, and activities: a longitudinal study. British Journal of Developmental Psychology, 18, 479–98.CrossRefGoogle Scholar
Campbell, D. W. and Eaton, W. O. (1999). Sex differences in the activity level of infants. Infant and Child Development, 8, 1–17.3.0.CO;2-O>CrossRefGoogle Scholar
Campredon, P. (1983). Sexe et age ratios chez le canard siffleur Anas penelope en periode hivernale en Europe de l'ouest. Revue d'Ecologie (Terre et Vie), 37, 117–28.Google Scholar
Carbone, C. and Owen, M. (1995). Differential migration of the sexes of pochard Aythya ferina: results from a European survey. Wildfowl, 46, 99–108.Google Scholar
Caro, T. (1998). Behavioral Ecology and Conservation Biology. New York: Oxford University Press.Google Scholar
Carpenter, C. C. (1956). Body temperatures of three species of Thamnophis. Ecology, 37, 732–5.CrossRefGoogle Scholar
Carrier, J. C., Pratt, H. L. and Martin, L. K. (1994). Group reproductive behaviours in free-living nurse sharks, Ginglymostoma cirratum. Copeia, 1994, 646–56.CrossRefGoogle Scholar
Carson, J., Burks, V. and Parke, R. (1993). Parent-child physical play: determination and consequences. In Parent-child Play, ed. MacDonald, K.. Albany: State University of New York Press, pp. 197–220.Google Scholar
Carter, S. L., Haas, C. A. and Mitchell, J. C. (1999). Home range and habitat selection of bog turtles in southwestern Virginia. Journal of Wildlife Management, 63, 853–60.CrossRefGoogle Scholar
Catry, P., Campos, A., Almada, V. and Cresswell, W. (2004). Winter segregation of migrant European robins Erithacus rubecula in relation to sex, age and size. Journal of Avian Biology, 35, 204–9.CrossRefGoogle Scholar
Catry, P., Phillips, R. A. and Furness, R. W. (1999). Evolution of reversed sexual size dimorphism in skuas and jaegers. Auk, 116, 158–68.CrossRefGoogle Scholar
Caughley, G. and Sinclair, A. R. E. (1994). Wildlife Ecology and Management. Cambridge, MA: Blackwell Science.Google Scholar
Cederlund, B.-M. (1987). Parturition and early development of moose (Alces alces L.) calves. Swedish Wildllife Research Supplement, 1, 399–422.Google Scholar
Cederlund, G. and Sand, H. K. G. (1992). Dispersal of subadult moose (Alces alces) in nonmigratory population. Canadian Journal of Zoology, 70, 1309–14.CrossRefGoogle Scholar
Cederlund, G., Sandegren, F. and Larsson, K. (1987). Summer movements of female moose and dispersal of their offspring. Journal of Wildlife Management, 51, 342–52.CrossRefGoogle Scholar
Censky, E. J. (1995). The evolution of sexual size dimorphism in the teiid lizard Ameiva plei. Behaviour, 132, 529–57.CrossRefGoogle Scholar
Cerling, T. E., Harris, J. M. and Passey, B. H. (2003). Diets of East African Bovidae based on stable isotope analysis. Journal of Mammalogy, 84, 456–70.2.0.CO;2>CrossRefGoogle Scholar
Cerling, T. E., Harris, J. M. and Passey, B. H.Chambers (The) Encyclopedic English Dictionary. (1994). Edinburgh: Chambers.Google Scholar
Chapman, C. A. (1990). Association patterns of spider monkeys: the influence of ecology and sex on social organization. Behavioral Ecology and Sociobiology, 26, 409–14.CrossRefGoogle Scholar
Chapman, C. A. and Wrangham, R. W. (1993). Range use of the forest chimpanzees of Kibale: implications for the understanding of chimpanzee social organization. American Journal of Primatology, 31, 263–73.CrossRefGoogle Scholar
Chapman, C. A., Wrangham, R. W. and Chapman, L. J. (1995). Ecological constraints on group size: an analysis of spider monkey and chimpanzee subgroups. Behavioral Ecology and Sociobiology, 32, 199–209.Google Scholar
Charland, M. B. and Gregory, P. T. (1990). The influence of female reproductive status on thermoregulation in a viviparous snake, Crotalus viridis. Copeia, 1990, 1089–98.CrossRefGoogle Scholar
Chase, J. D., Dixon, K. R., Gates, J. E., Jacobs, D. and Taylor, G. J. (1989). Habitat characteristics, population size, and home range of the bog turtle, Clemmys muhlenbergii, in Maryland. Journal of Herpetology, 23, 356–62.CrossRefGoogle Scholar
Cheney, D. L. (1978). The play partners of immature baboons. Animal Behaviour, 26, 1038–50.CrossRefGoogle Scholar
Chism, J. and Rowell, T. (1986). Mating and residence patterns of male patas monkeys (Erythrocebus patas). Ethology, 103, 109–26.CrossRefGoogle Scholar
Chivers, D. J. (1974). The Siamang in Malaysia. Basel: S. Karger.Google Scholar
Chivers, D. P., Brown, G. E. and Smith, R. J. F. (1995). Familiarity and shoal cohesion in fathead minnows (Pimephales promelas) – implications for antipredator behaviour. Canadian Journal of Zoology – Revue Canadienne De Zoologie, 73, 955–60.CrossRefGoogle Scholar
Choudhury, S. and Black, J. M. (1991). Testing the behavioural dominance and dispersal hypothesis in pochard. Ornis Scandinavica, 22, 155–9.CrossRefGoogle Scholar
Christal, J. and Whitehead, H. (1997). Aggregations of mature male sperm whales on the Galápagos Islands breeding ground. Marine Mammal Science, 13, 59–69.CrossRefGoogle Scholar
Christal, J., Whitehead, H. and Lettevall, E. (1998). Sperm whale social units: variation and change. Canadian Journal of Zoology, 76, 1431–40.CrossRefGoogle Scholar
Ciofi, C. and Chelazzi, G. (1994). Analysis of homing pattern in the colubrid snake Coluber viridiflavus. Journal of Herpetology, 28, 477–84.CrossRefGoogle Scholar
Clapham, P. J. (1996). The social and reproductive biology of humpback whales: an ecological perspective. Mammal Review, 26, 27–49.CrossRefGoogle Scholar
Clarke, J., Manly, B., Kerry, K. R.et al. (1998). Sex differences in Adélie penguin foraging strategies. Polar Biology, 20, 248–58.CrossRefGoogle Scholar
Clauss, M., Frey, R., Kiefer, B.et al. (2003). The maximum attainable body size of herbivorous mammals: morphophysiological constraints on foregut, and adaptations of hindgut fermenters. Oecologia, 136, 14–27.CrossRefGoogle ScholarPubMed
Clemmons, J. R. and Buchholz, R. (1997). Behavioural Approaches to Conservation in the Wild. Cambridge: Cambridge University Press.Google Scholar
Clinton, W. L. (1994). Sexual selection and growth in male northern elephant seals. In Elephant Seals: Population Ecology, Behaviour and Physiology, eds. Boeuf, B. J. and Laws, R. M.. Berkley: University of California Press, pp. 154–68.Google Scholar
Clutton-Brock, J., Dennis-Bryan, K., Armitage, P. L. and Jewell, P. A. (1990). Osteology of the Soay sheep. Bulletin of the British Museum of Natural History, 56, 1–56.Google Scholar
Clutton-Brock, T. H. (1977). Some aspects of intra-specific variation in feeding and ranging behaviour in primates. In Primate Ecology, ed. Clutton-Brock, T. H.. Cambridge: Cambridge University Press, pp. 539–56.Google Scholar
Clutton-Brock, T. H. (1983). Selection in relation to sex. In Evolution: From Molecules to Men, ed. Bendall, D. S.: Cambridge University Press, pp. 457–81.Google Scholar
Clutton-Brock, T. H. (1988). Reproductive success. In Reproductive Success, ed. Clutton-Brock, T. H.. Chicago: University of Chicago Press, pp. 472–85.Google Scholar
Clutton-Brock, T. H. (1989). Mammalian mating systems. Proceedings of the Royal Society of London, Series B, 236, 339–72.CrossRefGoogle ScholarPubMed
Clutton-Brock, T. H., Albon, S. D. and , Guinness F. E. (1985). Parental investment and sex differences in juvenile mortality in birds and mammals. Nature, 313, 131–3.CrossRefGoogle Scholar
Clutton-Brock, T. H., Albon, S. D. and , Guinness F. E. (1987b). Interactions between population density and maternal characteristics affecting fecundity and juvenile survival in red deer (Cervus elaphus). Journal of Animal Ecology, 56, 857–71.CrossRefGoogle Scholar
Clutton-Brock, T. H., Albon, S. D. and Guinness F. E. (1988). Reproductive success in male and female red deer. In Reproductive Success, ed. Clutton-Brock, T. H.. Chicago: University of Chicago Press, pp. 325–43.Google Scholar
Clutton-Brock, T. H., Guinness, F. E. and Albon, S. D. (1982). Red Deer: behavior and ecology of two sexes. Chicago: University of Chicago Press.Google Scholar
Clutton-Brock, T. H., Iason, G. R. and Guinness, F. E. (1987a). Sexual segregation and density-related changes in habitat use in male and female red deer (Cervus elaphus). Journal of Zoology, London, 211, 275–89.CrossRefGoogle Scholar
Clutton-Brock, T. H., Major, M. and Guinness, F. E. (1985). Population regulation in male and female red deer. Journal of Animal Ecology, 54, 831–46.CrossRefGoogle Scholar
Cochran, P. A. and McConville, D. R. (1983). Feeding by Trionyx spiniferus in backwaters of the upper Mississippi River. Journal of Herpetology, 17, 82–6.CrossRefGoogle Scholar
Cockburn, A. and Lazenby-Cohen, K. A. (1992). Use of nest trees by Antechinus stuartii a semelparous lekking marsupial. Journal of Zoology, London, 226, 657–80.CrossRefGoogle Scholar
Cockcroft, V. G. and Ross, G. J. B. (1990). Food and feeding of the Indian Ocean bottlenose dolphin off Southern Natal, South Africa. In The Bottlenose Dolphin, eds. Leatherwood, S. and Reeves, R. R.. San Diego: Academic Press, pp. 295–308.Google Scholar
Cogger, H. G. (2000). Reptiles and Amphibians of Australia. Sydney: Reed New Holland.Google Scholar
Collar, N. J. (1996). Family Otididae (Bustards). In Handbook of the Birds of the World, vol. 3., eds. Hoyo, J. del, Elliot, A. and Sargatal, J.. Barcelona: Lynx Edicions.Google Scholar
Compagno, L. J. V. (1984). FAO Species Catalogue. Volume 4 Sharks of the World, Parts 1 and 2. Rome: Food and Agriculture Organization of the United Nations.Google Scholar
Compagno, L. J. V. (1999). Checklist of living elasmobranchs. In Sharks, Skates and Rays: The Biology of Elasmobranch Fishes, ed. Hamlett, W. C.. Baltimore, MD: Johns Hopkins University Press, pp. 471–98.Google Scholar
Connor, R. C., Mann, J., Tyack, P. L. and Whitehead, H. (1998). Social evolution in toothed whales. Trends in Ecology and Evolution, 13, 228–32.CrossRefGoogle ScholarPubMed
Connor, R. C., Read, A. J. and Wrangham, R. (2000). Male reproductive strategies and social bonds. In Cetacean Societies: Field Studies of Dolphins and Whales, eds. Mann, J., Connor, R. C., Tyack, P. L. and Whitehead, H.. Chicago: University of Chicago Press, pp. 247–70.Google Scholar
Connor, R. C., Richards, A. F., Smolker, R. A. and Mann, J. (1996). Patterns of female attractiveness in Indian Ocean bottlenose dolphins. Behaviour, 133, 37–69.CrossRefGoogle Scholar
Connor, R. C., Smolker, R. A. and Richards, A. F. (1992). Two levels of alliance formation among male bottlenose dolphins (Tursiops sp.). Proceedings of the Academy of Science of United States, 89, 987–90.CrossRefGoogle Scholar
Conradt, L. (1997). Causes of sex differences in habitat use in red deer (Cervus elaphus, L.). Ph.D. thesis, University of Cambridge.
Conradt, L. (1998a). Could asynchrony in activity between the sexes cause intersexual social segregation in ruminants? Proceedings of the Royal Society of London, Series B, Biological Sciences, 265, 1359–63.CrossRefGoogle Scholar
Conradt, L. (1998b). Measuring the degree of sexual segregation in group-living animals. Journal of Animal Ecology, 67, 217–26.CrossRefGoogle Scholar
Conradt, L. (1999). Social segregation is not a consequence of habitat segregation in red deer and feral Soay sheep. Animal Behaviour, 57, 1151–7.CrossRefGoogle ScholarPubMed
Conradt, L. (2000). Use of a seaweed habitat by red deer (Cervus elaphus L.). Journal of Zoology, 250, 541–9.CrossRefGoogle Scholar
Conradt, L. and Roper, T. J. (2000). Activity synchrony and social cohesion: a fisson-fusion model. Proceedings of the Royal Society of London, Series B, 267, 2213–18.CrossRefGoogle ScholarPubMed
Conradt, L. and Roper, T. J. (2003). Group decision-making in animals. Nature, 421, 155–8.CrossRefGoogle ScholarPubMed
Conradt, L., Clutton-Brock, T. H. and Guinness, F. E. (1999a). The relationship between habitat choice and lifetime reproductive success in female red deer. Oecologia, 120, 218–24.CrossRefGoogle Scholar
Conradt, L., Clutton-Brock, T. H. and Guinness, F. E. (2000). Sex differences in weather sensitivity can cause habitat segregation: red deer as an example. Animal Behaviour, 59, 1049–60.CrossRefGoogle Scholar
Conradt, L., Clutton-Brock, T. H. and Thomson, D. (1999b). Habitat segregation in ungulates: are males forced into suboptimal foraging habitats through indirect competition by females? Oecologia, 119, 367–77.CrossRefGoogle Scholar
Conradt, L., Gordon, I. J., Clutton-Brock, T. H., Thomson, D. and Guinness, F. E. (2001). Could the indirect competition hypothesis explain inter-sexual site segregation in red deer (Cervus elaphus L.)? Journal of Zoology, London, 254, 185–93.CrossRefGoogle Scholar
Conroy, J. W. H. (1972). Ecological aspects of the biology of the giant petrel Macronectes giganteus (Gmelin) in the maritime Antarctic. Scientific Report of the British Antarctic Survey, 75, 1–74.Google Scholar
Constable, A. J. and Nicol, S. (2002). Defining smaller-scale management units to further develop the ecosystem approach in managing large-scale pelagic krill fisheries in Antarctica. CCAMLR Science, 9, 117–31.Google Scholar
Conway, C. J., Powell, G. V. N. and Nichols, J. D. (1995). Overwinter survival of neotropical migratory birds in early-successional and mature tropical forests. Conservation Biology, 9, 855–64.CrossRefGoogle Scholar
Cooper, J., Brooke, M. L., Burger, A. E., Crawford, R. J. M., Hunter, S. and Williams, T. (2001). Aspects of the breeding biology of the Northern Giant Petrel (Macronectes halli) and the Southern Giant Petrel (M. giganteus) at sub-Antarctic Marion Island. International Journal of Ornithology, 4, 53–68.Google Scholar
Cooper, J., Henley, S. and Klages, N. (1992). The diet of the wandering albatross Diomedea exulans at sub-Antarctic Marion Island. Polar Biology, 12, 477–84.CrossRefGoogle Scholar
Cooper, J., Wilson, R. P. and Adams, N. J. (1993). Timing of foraging by the wandering albatross Diomedea exulans. Proceedings National Institute Polar Research Symposium on Polar Biology, 6, 55–61.Google Scholar
Cooper, W. E. (2003). Sexual dimorphism in distance from cover but not escape behavior by the keeled earless lizard Holbrookia propinqua. Journal of Herpetology, 37, 374–8.CrossRefGoogle Scholar
Cords, M. (1984). Mating patterns and social structure in redtail monkeys (Cercopithecus ascanius). Zietschrift für Tierpsycholgie, 64, 213–39.Google Scholar
Corfield, T. (1973). Elephant mortality in Tsavo National Park, Kenya. East African Wildlife Journal, 11, 339–68.CrossRefGoogle Scholar
Cork, S. J. (1991). Meeting the energy-requirements for lactation in a macropodid marsupial – current nutrition versus stored body reserves. Journal of Zoology, 225, 567–76.CrossRefGoogle Scholar
Corkeron, P. J. and Connor, R. C. (1999). Why do baleen whales migrate? Marine Mammal Science, 15, 1228–45.CrossRefGoogle Scholar
Côté, S. D., Schaefer, J. A. and Messier, F. (1997). Time budgets and synchrony of activities in muskoxen: the influence of sex, age, and season. Canadian Journal of Zoology, 75, 1628–35.CrossRefGoogle Scholar
Coulson, G. (1993). The influence of population density and habitat on grouping in the western grey kangaroo, Macropus fuliginosus. Wildlife Research, 20, 151–62.CrossRefGoogle Scholar
Coulson, J. C. (2002). Colonial breeding in seabirds. In Biology of Marine Birds, eds. Schreiber, E. A. and Burger, J.. Boca Raton, Florida: CRC Press.Google Scholar
Couzin, I. D. and Krause, J. (2003). Self-organization and collective behaviour in vertebrates. Advances in the Study of Behaviour, 32, 1–75.CrossRefGoogle Scholar
Craig, M. J. (1992). Radio-telemetry and tagging study of movement patterns, activity cycles, and habitat utilization in Cagle's map turtle, Graptemys caglei. M. Sc. thesis, Canyon, TX: West Texas State University.
Cransac, N., Gerard, J.-F., Maublanc, M. L. and Pépin, D. (1998). An example of segregation between age and sex classes only weakly related to habitat use in mouflon sheep (Ovis gmelini). Journal of Zoology, London, 244, 371–8.CrossRefGoogle Scholar
Crawley, M. C. (1973). A live-trapping study of Australian brush-tailed possums, Trichosurus vulpecula (Kerr), in the Orongorongo Valley, Wellington, New Zealand. Australian Journal of Zoology, 21, 75–90.CrossRefGoogle Scholar
Creer, S., Chou, W. H., Malhotra, A. and Thorpe, R. S. (2002). Offshore insular variation in the diet of the Taiwanese bamboo viper Trimeresurus stejnegeri (Schmidt). Zoological Science, 19, 907–13.CrossRefGoogle Scholar
Cristol, D. A., Baker, M. B. and Carbone, C. (1999). Differential migration revisited. Latitudinal segregation by age and sex class. In Current Ornithology, vol. 15, eds. , V. Nolan Jr, Ketterson, E. D. and Thompson, C. F.. New York: Kluwer Academic/Plenum Publishers.Google Scholar
Croft, D. B. (1981). Social behaviour of the euro, Macropus robustus (Gould), in the Australian Arid Zone. Australian Wildlife Research, 8, 13–49.CrossRefGoogle Scholar
Croft, D. B. (1987). Socio-ecology of the antilopine wallaroo, Macropus antilopinus, in the Northern Territory, with observations on sympatric M. robustus woodwardii and M. agilis. Australian Wildlife Research, 14, 243–55.CrossRefGoogle Scholar
Croft, D. B. (1989). Social organisation of the Macropodoidea. In Kangaroos, Wallabies and Rat-kangaroos, eds. Grigg, G., Jarman, P. and Hume, I.. New South Wales, Australia: Surrey Beatty and Sons, pp. 505–25.Google Scholar
Croft, D. B. (1991a). Home range of the euro, Macropus robustus erubescens. Journal of Arid Environments, 20, 99–111.Google Scholar
Croft, D. B. (1991b). Home range of the red kangaroo, Macropus rufus. Journal of Arid Environments, 20, 83–98.Google Scholar
Croft, D. P., Arrowsmith, B. J., Bielby, J.et al. (2003). Mechanisms underlying shoal composition in the Trinidadian guppy (Poecilia reticulata). Oikos, 100, 429–38.CrossRefGoogle Scholar
Croft, D. P., Botham, M. S. and Krause, J. (2004). Is sexual segregation in the guppy, Poecilia reticulata, consistent with the predation risk hypothesis? Environmental Biology of Fishes, 71(2), 127–33.CrossRefGoogle Scholar
Crow, J. F. and Kimura, M. (1970). An Introduction to Population Genetics Theory. New York: Harper and Row.Google Scholar
Croxall, J. P. (1991). Constraints on reproduction in albatrosses. Proceedings of the 20th International Ornithological Congress, 281–302.Google Scholar
Croxall, J. P. (1995). Sexual size dimorphism in seabirds. Oikos, 73, 399–403.CrossRefGoogle Scholar
Croxall, J. P. (1998). Research and conservation: a future for albatrosses? In Albatross Biology and Conservation, eds. Robertson, G. and Gales, R.. Chipping Norton, Australia: Surrey Beatty and Sons, pp. 267–88.Google Scholar
Croxall, J. P. and Gales, R. (1998). An assessment of the conservation status of albatrosses. In Albatross Biology and Conservation, eds. Robertson, G. and Gales, R.. Chipping Norton, Australia: Surrey Beatty and Sons, pp. 46–65.Google Scholar
Croxall, J. P. and Prince, P. A. (1990). Recoveries of wandering albatrosses Diomedea exulans ringed at South Georgia (1958–1986). Ringing and Migration, 11, 43–51.CrossRefGoogle Scholar
Croxall, J. P. and Prince, P. A. (1996). Potential interactions between wandering albatrosses and longline fisheries for Patagonian toothfish at South Georgia. CCAMLR Science, 3, 101–10.Google Scholar
Croxall, J. P., Black, A. D. and Wood, A. G. (1999). Age, sex and status of wandering albatrosses Diomedea exulans L. in Falkland Islands waters. Antarctic Science, 11, 150–6.CrossRefGoogle Scholar
Croxall, J. P., Prince, P. A. and Reid, K. (1997). Dietary segregation of krill-eating South Georgia seabirds. Journal of Zoology, London, 242, 531–56.CrossRefGoogle Scholar
Croxall, J. P., Prince, P. A., Rothery, P. and Wood, A. D. (1998). Population changes in albatrosses at South Georgia. In Albatross Biology and Conservation, eds. Robertson, G. and Gales, R.. Chipping Norton: Surrey Beatty and Sons, pp. 69–83.Google Scholar
Croxall, J. P., Rothery, P., Pickering, S. P. C. and Prince, P. A. (1990). Reproductive performance, recruitment and survival of wandering albatrosses Diomedea exulans at Bird Island, South Georgia. Journal of Animal Ecology, 59, 775–96.CrossRefGoogle Scholar
Croxall, J. P., Silk, J. R. D., Phillips, R. A., Afanasyev, V. and Briggs, D. R. (2005). Global circumnavigations: tracking year-round ranges of nonbreeding albatrosses. Science, 307, 249–50.CrossRefGoogle ScholarPubMed
Cryan, P. M., Bogan, M. A. and Altenbach, J. S. (2000). Effect of elevation on distribution of female bats in the Black Hills, South Dakota. Journal of Mammalogy, 81, 719–25.2.3.CO;2>CrossRefGoogle Scholar
Cuadrado, M. (1997). Why are migrant Robins Erithacus rubecula territorial in winter?: the importance of the anti-predation behaviour. Ethology, Ecology and Evolution, 9, 77–88.CrossRefGoogle Scholar
Curlewis, J. D. (1989). The breeding season of Bennett's wallaby (Macropus rufogriseus rufogriseus) in Tasmania. Journal of Zoology, London, 218, 337–9.CrossRefGoogle Scholar
Cutter, J. (2002). Dietary Segregation in the Western Grey Kangaroo, Macropus fuliginosus, at Hattah-Kulkyne National Park. Unpublished honours thesis, University of Melbourne.Google Scholar
D'Onghia, G., Matarrese, A., Tursi, A. and Sion, L. (1995). Observations on the depth distribution pattern of the small-spotted catshark in the north Aegean Sea. Journal of Fish Biology, 47, 421–6.CrossRefGoogle Scholar
Dahlheim, M. E. and Heyning, J. E. (1999). Killer whale – Orcinus orca (Linnaeus, 1758). In Handbook of Marine Mammals: The Second Book of Dolphins and the Porpoises, vol. 6, eds. Ridgway, S. H. and Harrison, R.. London: Academic Press, pp. 281–322.Google Scholar
Dalrymple, G. H. (1977). Intraspecific variation in the cranial feeding mechanism of the turtles of the genus Trionyx (Reptilia, Testudines, Trionychidae). Journal of Herpetology, 11, 255–85.CrossRefGoogle Scholar
Daltry, J. C., Wuster, W. and Thorpe, R. S. (1998). Intraspecific variation in the feeding ecology of the crotaline snake Calloselasma rhodostoma in Southeast Asia. Journal of Herpetology, 32, 198–205.CrossRefGoogle Scholar
Dalziell, J. and Poorter, M. (1993). Seabird mortality in longline fisheries around South Georgia. Polar Record, 29, 143–5.CrossRefGoogle Scholar
Daneri, G. A. and Coria, N. R. (1992). The diet of Antarctic fur seals, Arctocephalus gazella, during the summer-autumn period at Mossman Peninsula, Laurie Island (South Orkneys). Polar Biology, 11, 565–6.CrossRefGoogle Scholar
Darwin, C. (1871). The Decent of Man and Selection in Relation to Sex, London: Murray.Google Scholar
Daunt, F., Monaghan, P., Wanless, S., Harris, M. P. and Griffiths, R. (2001). Sons and daughters: age-specific differences in parental rearing capacities. Functional Ecology, 15, 211–16.CrossRefGoogle Scholar
Daut, E. F. and Andrews, R. M. (1993). The effect of pregnancy on thermoregulatory behavior of the viviparous lizard Chalcides ocellatus. Journal of Herpetology, 27, 6–13.CrossRefGoogle Scholar
Davis, L. S. and Speirs, E. A. H. (1990). Mate choice in penguins. In Penguin Biology, eds. Davis, S. and Darby, J. T.. San Diego, California: Academic Press, pp. 377–97.Google Scholar
Davis, R. B., Herreid, C. F. and Short, H. L. (1962). Mexican free-tailed bats in Texas. Ecological Monographs, 32, 311–46.CrossRefGoogle Scholar
Dawbin, W. H. (1982). The tuatara Sphenodon punctatus (Reptilia: Rhynchocephalia): a review. In New Zealand Herpetology, ed. Newman, D. G.. Wellington, New Zealand: New Zealand Wildlife Service, pp. 149–81.Google Scholar
Day, T., O'Connor, C. and Matthews, L. (2000). Possum social behaviour. In The Brushtail Possum: Biology, Impact and Management of an Introduced Marsupial, ed. Montague, T. L.. Lincoln, NZ: Manaaki Whenua Press, pp. 35–46.Google Scholar
Deere, J. A. (2001). The use of black rhinoceros (Diceros bicornis) dung sampling to investigate sexual differences in diet quality and midden site selection in Pilanesberg National Park. Unpublished dissertation, University of Pretoria.
del Hoyo, J. (1994). Family Cracidae (Chachalacas, Guans and Curassows). In Handbook of the Birds of the World, vol. 2., eds. Hoyo, J. del, Elliot, A. and Sargatal, J.. Barcelona: Lynx Edicions.Google Scholar
Delgado, R. A. and Schaik, C. P. (2000). The behavioral ecology and conservation of the orangutan (Pongo pygmaeus): a tale of two islands. Evolutionary Anthropology, 9, 201–18.3.0.CO;2-Y>CrossRefGoogle Scholar
Demment, M. W. and Soest, P. J. (1985). A nutritional explanation for body-size patterns of ruminant and nonruminant herbivores. American Naturalist, 125, 641–72.CrossRefGoogle Scholar
Demski, L. S. and Northcutt, R. G. (1996). The brain and cranial nerves of the white shark: an evolutionary perspective. In Great White Sharks: The Biology of Carcharodon carcharias, eds. Klimley, A. P. and Ainley, D. G.. San Diego, CA: Academic Press, pp. 121–30.Google Scholar
Deneubourg, J. L. and Goss, S. (1989). Collective patterns and decision making. Ethology, Ecology and Evolution, 1, 295–311.CrossRefGoogle Scholar
Deneubourg, J. L., Goss, S., Franks, N. et al. (1991). The dynamics of collective sorting robot-like ants and ant-like robots. In Simulation of Adaptive Behavior: From Animals to Animats, eds. Meyer, J. A. and Wilson, S. W.. Cambridge, MA: MIT Press/Bradford Books, pp. 356–65.Google Scholar
Dennis, A. J. and Marsh, H. (1997). Seasonal reproduction in musky rat-kangaroos, Hypsiprymnodon moschatus: a response to changes in resource availability. Wildlife Research, 24, 561–78.CrossRefGoogle Scholar
Desrochers, A. (1989). Sex, dominance, and microhabitat use in wintering black-capped chickadees: a field experiment. Ecology, 70, 636–45.CrossRefGoogle Scholar
Deutsch, C. J., Crocker, D. E., Costa, D. P. and Le Boeuf, B. J. (1994). Sex- and age-related variation in reproductive effort of northern elephant seals. In Elephant Seals: Population Ecology, Behaviour and Physiology, eds. Boeuf, B. J. and Laws, R. M.. Berkley: University of California Press, pp. 169–210.Google Scholar
Deutsch, C. J., Haley, M. P. and Leboeuf, B. J. (1990). Reproductive effort of male northern elephant seals – estimates from mass-loss. Canadian Journal of Zoology, 68, 2580–93.CrossRefGoogle Scholar
Di Bitetti, M. S. and Janson, C. H. (2000). When will the stork arrive? Patterns of birth seasonality in neotropic primates. American Journal of Primatology, 50, 109–30.3.0.CO;2-W>CrossRefGoogle Scholar
Dickman, C. R. (1995). Agile Antechinus. In The Mammals of Australia, ed. Strahan, R.. Sydney: Reed New Holland, pp. 99–101.Google Scholar
Diego-Rasilla, F. J. and Perez-Mellado, V. (2003). Home range and habitat selection by Podarcis hispanica (Squamata, Lacertidae) in western Spain. Folia Zoologica, 52, 87–98.Google Scholar
DiFiore, A. and Rodman, P. S. (2001). Time allocation patterns of lowland woolly monkeys (Lagothrix lagotricha poepigii) in a neotropical terra firma forest. International Journal of Primatology, 22, 449–80.Google Scholar
Dobson, A. and Poole, J. (1998). Conspecific aggregation and conservation biology. In Behavioral Ecology and Conservation Biology, ed. Caro, T.. New York: Oxford University Press, pp. 193–208.Google Scholar
, Dodd C. K. Jr, Enge, K. M. and Stuart, J. N. (1988). Aspects of the biology of the flattened musk turtle, Sternotherus depressus, in northern Alabama. Bulletin of the Florida State Museum, Biological Sciences Series, 34, 1–64.Google Scholar
Dodd, J. M. (1983). Reproduction in cartilaginous fishes (Chondrichthyes). In Fish Physiology, Vol. 9 – Reproduction: Endocrine Tissues and Hormones, eds. Hoar, W. S., Randall, D. J. and Donaldson, D. M.. New York: Academic Press, pp. 31–87.Google Scholar
Doidge, D. W. (1990). Integumentary heat loss and blubber distribution in the beluga, Delphinapterus leucas, with comparisons to narwhal, Monodon monoceros. Canadian Bulletin of Fisheries and Aquatic Sciences, 224, 129–40.Google Scholar
Doidge, D. W., McCann, T. S. and Croxall, J. P. (1986). Attendance behavior of Antarctic fur seals. In Fur Seals: Maternal Strategies on Land and at Sea, eds. Gentry, R. L. and Kooyman, G. L.. Princeton: Princeton University Press, pp. 102–14.CrossRefGoogle Scholar
Doody, J. S., Young, J. E. and Georges, A. (2002). Sex differences in activity and movements in the pig-nosed turtle, Carettochelys insculpta, in the wet-dry tropics of Australia. Copeia, 2002, 93–103.CrossRefGoogle Scholar
Dressen, W. (1993). On the behaviour and social organisation of agile wallabies, Macropus agilis (Gould, 1842) in two habitats of northern Australia. Zeitschrift für Säugetierkunde, 58, 201–11.Google Scholar
Du, W. G., Yan, S. J. and Ji, X. (2000). Selected body temperature, thermal tolerance and thermal dependence of food assimilation and locomotor performance in adult blue-tailed skinks, Eumeces elegans. Journal of Thermal Biology, 25, 197–202.CrossRefGoogle Scholar
du Toit, J. T. (1988). Patterns of resource use within the browsing ruminant guild in the central Kruger National Park. Unpublished Ph.D. thesis, Johannesburg: University of the Witwatersrand.
du Toit, J. T. (1990). Feeding height stratification among African browsing ruminants. African Journal of Ecology, 28, 55–61.CrossRefGoogle Scholar
du Toit, J. T. (1995). Sexual segregation in kudu: sex differences in competitive ability, predation risk or nutritional needs? South African Journal of Wildlife Research, 25, 127–32.Google Scholar
du Toit, J. T. (2003). Large herbivores and savanna heterogeneity. In The Kruger Experience: Ecology and Management of Savanna Heterogeneity, eds. du Toit, J. T., Rogers, K. H. and Biggs, H. C.. Washington, DC: Island Press, pp. 292–309.Google Scholar
Dubois, M., Bon, R., Cransac, N. and Maublanc, M. L. (1994). Dispersion patterns among ewes of Corsican mouflon: importance and proximate influences. Applied Animal Behaviour Science, 42, 29–40.CrossRefGoogle Scholar
Dubois, M., Quenette, P. Y., Bideau, E. and Magnac, M. P. (1993). Seasonal range use by European mouflon rams in medium altitude mountains. Acta Theriology, 38, 185–98.CrossRefGoogle Scholar
Duellman, W. E. and Trueb, L. (1986). Biology of Amphibians. New York: McGraw-Hill.Google Scholar
Dufault, S. and Whitehead, H. (1995a). An encounter with recently wounded sperm whales (Physeter macrocephalus). Marine Mammal Science, 11, 560–3.CrossRefGoogle Scholar
Dufault, S. and Whitehead, H. (1995b). The geographic stock structure of female and immature sperm whales in the South Pacific. Report of the International Whaling Commission, 45, 401–5.Google Scholar
Dulvy, N. K. and Reynolds, J. D. (1997). Evolutionary transitions among egg-laying, live-bearing and maternal inputs in sharks and rays. Proceedings of the Royal Society of London B, 264, 1309–15.CrossRefGoogle Scholar
Dunbar, R. I. M. (1977). Feeding ecology of gelada baboons: a preliminary report. In Primate Ecology, ed. Clutton-Brock, T. H.. Cambridge: Cambridge University Press, pp. 251–73.Google Scholar
Dunbar, R. I. M. (1984). Reproductive Decisions: An Economic Analysis of Gelada Baboon Social Organization. Princeton: Princeton University Press.Google Scholar
Dunbar, R. I. M. and Dunbar, P. (1988). Maternal time budgets of gelada baboons. Animal Behaviour, 76, 970–80.CrossRefGoogle Scholar
Dunn, D. G., Barco, S. G., Pabst, D.-A. and McLellan, W. A. (2002). Evidence of infanticide in bottlenose dolphins of the Western North Atlantic. Journal of Wildlife Diseases, 38, 505–10.CrossRefGoogle ScholarPubMed
Durell, S. E. A. Le V. dit (2000). Individuals feeding specialisation in shorebirds: population consequences and conservation implications. Biological Reviews of the Cambridge Philosophical Society, 75, 503–18.CrossRefGoogle Scholar
Durell, S. E. A. Le V. dit and Goss-Custard, J. D. (1996). Oystercatcher Haematopus ostralegus sex ratios on the wintering grounds: the case of the Exe estuary. Ardea, 84A, 373–81.Google Scholar
Durell, S. E. A. Le V. dit, Goss-Custard, J. D. and Caldow, R. W. G. (1993). Sex-related differences in diet and feeding method in the oystercatcher Haematopus ostralegus. Journal of Animal Ecology, 62, 205–15.CrossRefGoogle Scholar
Durell, S. E. A. Le V. dit, Goss-Custard, J. D., Caldow, R. W. G., Malcolm, H. M. and Osborn, D. (2001a). Sex, diet and feeding method-related differences in body condition in the oystercatcher Haematopus ostralegus. Ibis, 143, 107–19.CrossRefGoogle Scholar
Durell, S. E. A. Le V. dit, Goss-Custard, J. D. and Clarke, R. T. (2001b). Modelling the population consequences of age- and sex-related differences in winter mortality in the oystercatcher, Haematopus ostralegus. Oikos, 95, 69–77.CrossRefGoogle Scholar
Durell, S. E. A. Le V. dit, Ormerod, S. J. and Dare, P. J. (1996). Differences in population structure between two oystercatcher Haematopus ostralegus roosts on the Burry Inlet, South Wales. Ardea, 84A, 383–8.Google Scholar
Durner, G. M. and Gates, J. E. (1993). Spatial ecology of black rat snakes on Remington Farms, Maryland. Journal of Wildlife Management, 57, 812–26.CrossRefGoogle Scholar
Durtsche, R. D. (1992). Feeding time strategies of the fringe-toed lizard, Uma inornata, during breeding and non-breeding seasons. Oecologia, 89, 85–9.CrossRefGoogle ScholarPubMed
Duvall, D., King, M. B. and Gutzwiller, M. J. (1985). Behavioral ecology and ethology of the prairie rattlesnake. National Geographic Research, 1, 80–111.Google Scholar
Dwyer, D. P. (1966). The population pattern of Miniopterus schreibersii (Chiroptera) in north-eastern New South Wales. Australian Journal of Zoology, 14, 1073–137.CrossRefGoogle Scholar
Eaton, W. C. and Enns, L. R. (1986). Sex differences in human motor activity level. Psychology Bulletin, 100, 19–28.CrossRefGoogle ScholarPubMed
Eaton, W. C. and Yu, A. (1989). Are sex function of sex differences in maturational status? Child Development, 60, 1005–11.CrossRefGoogle ScholarPubMed
Eaton, W. C., Enns, L. R. and Presse, M. (1987). Scheme for observing activity. Journal of Psychoeducational Assessment, 3, 273–80.CrossRefGoogle Scholar
Eberle, M. and Kappeler, P. (2002). Mouse lemurs in space and time: a test of the socioecological model. Behavioral Ecology and Sociobiology, 51, 131–9.CrossRefGoogle Scholar
Ebert, D. A. (2002). Ontogenetic changes in the diet of the sevengill shark (Notorynchus cepedianus). Marine and Freshwater Research, 53, 517–23.CrossRefGoogle Scholar
Economakis, A. E. and Lobel, P. S. (1998). Aggregation behaviour of the grey reef shark, Carcharhinus amblyrhynchos, at Johnston Atoll, Central Pacific Ocean. Environmental Biology of Fishes, 51, 129–39.CrossRefGoogle Scholar
Edwards, J. (1983). Diet shifts in moose due to predator avoidance. Oecologia, 60, 185–9.CrossRefGoogle ScholarPubMed
Eifler, D. A. and Eifler, M. A. (1999a). Foraging behavior and spacing patterns of the lizard Oligosoma grande. Journal of Herpetology, 33, 632–9.CrossRefGoogle Scholar
Eifler, D. A. and Eifler, M. A. (1999b). The influence of prey distribution on the foraging strategy of the lizard Oligosoma grande (Reptilia: Scincidae). Behavioral Ecology and Sociobiology, 45, 397–402.CrossRefGoogle Scholar
Eisenberg, J. F. (1981). The Mammalian Radiations: An Analysis of Trends in Evolution, Adaptation, and Behavior. Chicago: The University of Chicago Press.Google Scholar
Ellis, J. R., Pawson, M. G. and Shackley, S. E. (1996). The comparative feeding ecology of six species of shark and four species of ray (Elasmobranchii) in the north-east Atlantic. Journal of the Marine Biological Association of the United Kingdom, 76, 89–106.CrossRefGoogle Scholar
Emlen, S. T. and Oring, L. W. (1977). Ecology, sexual selection, and the evolution of mating systems. Science, 197, 215–23.CrossRefGoogle ScholarPubMed
Ena, V., Lucio, A. and Purroy, F. J. (1985). The great bustard in Leon, Spain. Bustard Studies, 2, 35–52.Google Scholar
Endler, J. A. (1980). Natural selection on colour patterns in Poecillia reticulata. Evolution, 34, 76–91.CrossRefGoogle Scholar
Entwistle, A. C., Racey, P. A. and Speakman, J. R. (1996). Habitat exploitation by a gleaning bat, Plecotus auritus. Philosophical Transactions of the Royal Society London, B, 351, 921–31.CrossRefGoogle Scholar
Entwistle, A. C., Racey, P. A. and Speakman, J. R. (2000). Social and population structure of a gleaning bat, Plecotus auritus. Journal of Zoology, London, 252, 11–17.CrossRefGoogle Scholar
Ernst, C. H., Lovich, J. E. and Barbour, R. W. (1994). Turtles of the United States and Canada. Washington, D.C.: Smithsonian Institution Press.Google Scholar
Estep, D. Q., Crowell-Davis, S. L., Earl-Costello, S.-A. and Beatey, S. A. (1993). Changes in the social behaviour of drafthorse (Equus caballus) mares coincident with foaling. Applied Animal Behaviour Science, 35, 199–213.CrossRefGoogle Scholar
Estes, R. D. (1991a). The Behavior Guide to African Mammals: including hoofed mammals, carnivores, primates. Berkley: University of California Press.Google Scholar
Estes, R. D. (1991b). The significance of horns and other male secondary sexual characters in female bovids. Applied Animal Behaviour Science, 29, 403–51.CrossRefGoogle Scholar
Fabes, R. A. (1994). Physiological, emotional, and behavioral correlates of gender segregation. In Childhood Gender Segregation: Causes and Consequences, ed. Leaper, C.. San Francisco: Jossey-Bass, pp. 33–50.Google Scholar
Fagot, B. I. (1994). Peer relations and the development of competence in boys and girls. In Childhood Gender Segregation: Causes and Consequences, ed. Leaper, C.. San Francisco: Jossey-Bass, pp. 53–66.Google Scholar
Fairbanks, L. A. (1993). Juvenile vervet monkeys: establishing relationships and practicing skills for the future. In Juvenile Primates, eds. Pereira, M. E. and Fairbanks, L. A.. Oxford: Oxford University Press, pp. 211–27.Google Scholar
Fashing, P. (2001). Activity and ranging patterns of guerezas in the Kakamega Forest: inter-group variation and implications for intra-group feeding competition. International Journal of Primatology, 22, 549–78.CrossRefGoogle Scholar
Feare, C. J., Gill, E. L., McKay, H. V. and Bishop, J. D. (1995). Is the distribution of starlings Sturnus vulgaris within roosts determined by competition? Ibis, 137, 379–82.CrossRefGoogle Scholar
Fein, G. (1981). Pretend play in childhood: an integrative review. Child Development, 52, 1095–118.CrossRefGoogle Scholar
Feldheim, K. A., Gruber, S. H. and Ashley, M. V. (2002). The breeding biology of lemon sharks at a tropical nursery lagoon. Proceedings of the Royal Society of London B, 269, 1655–61.CrossRefGoogle Scholar
Fenton, M. B. (1990). The foraging behaviour and ecology of animal-eating bats. Canadian Journal of Zoology, 68, 411–22.CrossRefGoogle Scholar
Fenton, M. B., Rautenbach, I. L., Smith, S. E., Swanepoel, C. M., Grosell, J. and Jaarsveld, J. (1994). Raptors and bats: threats and opportunities. Animal Behaviour, 48, 9–18.CrossRefGoogle Scholar
Festa-Bianchet, M. (1986). Site fidelity and seasonal range use by bighorn rams. Canadian Journal of Zoology, 64, 2126–32.CrossRefGoogle Scholar
Festa-Bianchet, M. (1988). Seasonal range selection in bighorn sheep: conflicts between forage quality, forage quantity, and predator avoidance. Oecologia, 75, 580–6.CrossRefGoogle ScholarPubMed
Festa-Bianchet, M. (1991). The social system of bighorn sheep: grouping patterns, kinship and female dominance rank. Animal Behaviour, 42, 71–82.CrossRefGoogle Scholar
Festa-Bianchet, M. and Apollonio, M. (2003). Animal Behavior and Wildlife Conservation. Washington, D.C.: Island Press.Google Scholar
Fietz, C. (1999). Mating systems of Microcebus murinus. American Journal of Primatology, 48, 127–33.3.0.CO;2-4>CrossRefGoogle ScholarPubMed
Fisher, D. O. and Owens, I. P. F. (2000). Female home range size and the evolution of social organization in macropod marsupials. Journal of Animal Ecology, 69, 1083–98.CrossRefGoogle Scholar
Fisher, D. O., Blomberg, S. P. and Owens, I. P. F. (2002). Convergent maternal care strategies in ungulates and macropods. Evolution, 56, 167–76.CrossRefGoogle ScholarPubMed
Fitch, H. S. (1960). Autecology of the copperhead. University of Kansas Publications, Museum of Natural History, 13, 85–288.Google Scholar
Fitch, H. S. (1982). Resources of a snake community in prairie-woodland habitat of northeastern Kansas. In Herpetological Communities, ed. , N. J. Scott Jr.Washington, D.C.: U.S. Department of the Interior, Fish and Wildlife Service, pp. 83–98.Google Scholar
Fitch, H. S. (1999). A Kansas Snake Community: Composition and Changes Over 50 Years. Malabar, Florida: Krieger.Google Scholar
Fitch, H. S. and Shirer, H. W. (1971). A radiotelemetric study of spatial relationships in some common snakes. Copeia, 1971, 118–28.CrossRefGoogle Scholar
Fleming, M. R. and Frey, H. (1984). Aspects of the natural history of Feathertail Gliders Acrobates pygmaeus (Marsupialia: Burramyidae) in Victoria. In Possums and Gliders, eds. Smith, A. P. and Hume, I. D.. Chipping Norton, NSW: Surrey Beatty and Sons, pp. 403–8.Google Scholar
Fleming, T. H. and Eby, P. (2003). Ecology of bat migration. In Bat Ecology, eds. Kunz, T. H. and Fenton, M. B.. Chicago: University of Chicago Press, pp. 156–208.Google Scholar
Fleming, T. H. and Hooker, R. S. (1975). Anolis cupreus: the response of a lizard to tropical seasonality. Ecology, 56, 1243–61.CrossRefGoogle Scholar
Fletcher, T. and Selwood, L. (2000). Possum reproduction and development. In The Brushtail Possum: Biology, Impact and Management of an Introduced Marsupial, ed. Montague, T. L.. Lincoln, NZ: Manaaki Whenua Press, pp. 62–81.Google Scholar
Ford, D. (1999). Foraging ecology and demography of Sternotherus odoratus in a southwestern Missouri population. M.Sc. thesis, Springfield: Southwest Missouri State University.
Ford, E. (1921). A contribution to our knowledge of the life-histories of the dogfishes landed at Plymouth. Journal of the Marine Biological Association of the United Kingdom, 12, 468–505.CrossRefGoogle Scholar
Forero, M. G. and Hobson, K. A. (2003). Using stable isotopes of Nitrogen and Carbon to study seabird ecology: applications in the Mediterranean seabird community. Scientia Marina, 67, 23–32.CrossRefGoogle Scholar
Forero, M. G., Hobson, K. A., Bortolotti, G. R.et al. (2002). Food resource utilisation by the Magellanic penguin evaluated through stable-isotope analysis: segregation by sex and age and influence on offspring quality. Marine Ecology Progress Series, 234, 289–99.CrossRefGoogle Scholar
Fox, S. F., Conder, J. M. and Smith, A. E. 1998. Sexual dimorphism in the case of tail autotomy: Uta stansburiana with and without previous tail loss. Copeia, 1998, 376–82.CrossRefGoogle Scholar
Fragaszy, D. M. and Boinski, S. (1995). Patterns of individual diet choice and efficiency of foraging of wedge-capped capuchin monkeys (Cebus olivaceous). Journal of Comparative Psychology, 109, 339–48.CrossRefGoogle Scholar
Fraker, M. A., Gordon, C. D., McDonald, J. W., Ford, J. K. B. and Cambers, G. (1979). White whale (Delphinapterus leucas) distribution and abundance and the relationship to physical and chemical characteristics of the Mackenzie Estuary. Canadian Fisheries and Marine Service Technical Report, 863, v–56.Google Scholar
Francisci, F., Focardi, S. and Boitani, L. (1985). Male and female Alpine ibex: phenology of space use and herd size. In The Biology and Management of Mountain Ungulates, ed. Lovari, S.. London: Croom Helm, pp. 124–33.Google Scholar
Freake, M. J. (1998). Variation in homeward orientation performance in the sleepy lizard (Tiliqua rugosa): effects of sex and reproductive period. Behavioral Ecology and Sociobiology, 43, 339–44.CrossRefGoogle Scholar
Frid, A. (1999). Huemul (Hippocamelus bisulcus) sociality at a periglacial site: sexual segregation and habitat effects on group size. Canadian Journal of Zoology, 77, 1083–91.CrossRefGoogle Scholar
Friend, J. (1989). Myrmecobiidae: In Fauna of Australia, eds. Walton, D. and Richardson, B.. Canberra: Australian Government Publishing Service, pp. 583–90.Google Scholar
Friend, J. (1995). Numbat: In The Mammals of Australia, ed. Strahan, R.. Sydney: Reed New Holland, pp. 160–2.Google Scholar
Friend, J. (1996). Numbats on a junk food diet. Nature Australia, 29, 40–9.Google Scholar
Friend, J. and Whitford, D. (1993). Maintenance and breeding of the numbat (Myrmecobius fasciatus) in captivity. In Biology and Management of Australasian Carnivorous Marsupials, eds. Roberts, M., Carnio, J., Crawshaw, G. and Hutchins, M.. Toronto: Metropolitan Toronto Zoo and Monotreme and Marsupial Advisory Group of AAZPA, pp. 103–24.Google Scholar
Frischknecht, M. (1993). The breeding coloration of male 3-spined sticklebacks (Gasterosteus aculeatus) as an indicator of energy investment in vigor. Evolutionary Ecology, 7, 439–50.CrossRefGoogle Scholar
Frost, K. J., Russel, R. B. and Lowry, L. F. (1992). Killer whales, Orcinus orca, in the Southeastern Bearing Sea: recent sightings and predation on other marine mammals. Marine Mammal Science, 8, 110–19.CrossRefGoogle Scholar
Furness, R. W. (1993). Birds as monitors of pollutants. In Birds as Monitors of Environment Change, eds. Furness, R. W. and Greenwood, J. J. D.. London: Chapman and Hall, pp. 86–143.CrossRefGoogle Scholar
Gadsen, H. and Palacios-Orona, L. E. (1997). Seasonal dietary patterns of the Mexican fringe-toed lizard (Uma paraphygas). Journal of Herpetology, 31, 1–9.Google Scholar
Galan, P. (1999). Demography and population dynamics of the lacertid lizard Podarcis bocagei in north-west Spain. Journal of Zoology, London, 249, 203–18.CrossRefGoogle Scholar
Galdikas, B. (1988). Orangutan diet, range, and activity at Tanjung Putting, Central Borneo. International Journal of Primatology, 9, 1–35.CrossRefGoogle Scholar
Gales, N. (2002). New Zealand sea lion. In Encyclopedia of Marine Mammals, eds. Perrin, W. F., Wursig, B. and Thewissen, J. G. M.. San Diego: Academic Press.Google Scholar
Gales, R. (1998). Albatross populations: status and threats. In Albatross Biology and Conservation, eds. Robertson, G. and Gales, R.. Chipping Norton, Australia: Surrey Beatty and Sons, pp. 20–45.Google Scholar
Gales, R. (1993). Co-operative Mechanisms for the Conservation of Albatrosses. Hobart, Australia: Australian Nature Conservation Agency and Australian Antarctic Foundation.
Gales, R., Brothers, N. and Reid, T. (1998). Seabird mortality in the Japanese tuna longline fishery around Australia. Biological Conservation, 86, 37–56.CrossRefGoogle Scholar
Galois, P., Leveille, M., Bouthillier, L., Daigle, C. and Parren, S. (2002). Movement patterns, activity and home range of the eastern spiny softshell turtle (Apalone spinifera) in northern Lake Champlain, Quebec, Vermont. Journal of Herpetology, 36, 402–11.CrossRefGoogle Scholar
Gannon, V. P. J. and Secoy, D. M. (1985). Seasonal and daily activity patterns in a Canadian population of the prairie rattlesnake, Crotalus viridis viridis. Canadian Journal of Zoology, 63, 86–91.CrossRefGoogle Scholar
Garrick, L. D. (1974). Reproductive influences on behavioral thermoregulation in the lizard, Sceloporus cyanogenys. Physiology and Behaviour, 12, 85–91.CrossRefGoogle Scholar
Garstka, W. R., Camazine, B. and Crews, D. (1982). Interactions of behavior and physiology during the annual reproductive cycle of the red-sided garter snake (Thamnophis sirtalis parietalis). Herpetologica, 38, 104–23.Google Scholar
Garton, J. S. and Dimmick, R. W. (1969). Food habits of the copperhead in middle Tennessee. Journal of the Tennessee Academy of Science, 44, 113–17.Google Scholar
Gates, S. (2002). Review of methodology of quantitative reviews using meta-analysis in ecology. Journal of Animal Ecology, 71, 547–57.CrossRefGoogle Scholar
Gauthreaux, S. A. Jr (1978). The ecological significance of behavioural dominance. Perspectives in Ethology, 3, 17–54.CrossRefGoogle Scholar
Gauthreaux, S. A. Jr (1982). The ecology and evolution of avian migration systems. In Avian Biology Vol. 6, eds. Farner, D. S., King, J. R. and Parkes, K. C.. New York: Academic Press.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, 237–69.CrossRefGoogle Scholar
Gehrt, S. D. and Fritzell, E. K. (1996). Sex-biased response of raccoons (Procyon lotor) to live traps. American Midland Naturalist, 135, 23–32.CrossRefGoogle Scholar
Geisler, J. H. and Uhen, M. D. (2003). Morphological support for a close relationship between hippo and whales. Journal of Vertebrate Paleontology, 23, 991–6.CrossRefGoogle Scholar
Geist, V. (1968). On delayed social and physical maturation in mountain sheep. Canadian Journal of Zoology, 46, 899–904.CrossRefGoogle ScholarPubMed
Geist, V. (1971). Mountain Sheep: a Study in Behavior and Evolution. Chicago: University Chicago Press.Google Scholar
Geist, V. (1974). On the relationship of social evolution and ecology in ungulates. American Zoologist, 14, 205–20.CrossRefGoogle Scholar
Geist, V. and Bromley, P. T. (1978). Why deer shed antlers. Zeitschrift für Säugetierkunde – International Journal of Mammalian Biology, 43, 223–31.Google Scholar
Geist, V. and Petocz, R. G. (1977). Bighorn sheep in winter: do rams maximize reproductive fitness by spatial and habitat segregation from ewes? Canadian Journal of Zoology, 55, 1802–10.CrossRefGoogle Scholar
Gerard, J.-F. and Loisel, P. (1995). Spontaneous emergence of a relationship between habitat openness and mean group size and its possible evolutionary consequences in large herbivores. Journal of Theoretical Biology, 176, 511–22.CrossRefGoogle Scholar
Gerard, J.-F., Bideau, E., Maublanc, M.-L., Loisel, P. and Marchal, C. (2002). Herd size in large herbivores: encoded in the individual or emergent? Biological Bulletin, 202, 275–82.CrossRefGoogle ScholarPubMed
Gerard, J.-F., Pendu, Y., Maublanc, M.-L., Vincent, J.-P., Poulle, M.-L. and Cibien, C. (1995). Large group formation in European roe deer, an adaptive feature? Revue d'Ecologie (Terre Vie), 50, 391–401.Google Scholar
Gerpe, M. S., Moreno, J. E., Moreno, V. J. and Patat, M. L. (2000). Cadmium, zinc and copper accumulation in the squid Illex argentinus from the Southwest Atlantic Ocean. Marine Biology, 136, 1039–44.CrossRefGoogle Scholar
Gibbons, J. W. (1986). Movement patterns among turtle populations: applicability to management of the desert tortoise. Herpetologica, 42, 104–13.Google Scholar
Gibson, R. and Falls, J. B. (1979). Thermal biology of the common garter snake Thamnophis sirtalis (L.). 1. Temporal variation, environmental effects and sex differences. Oecologia, 43, 79–97.CrossRefGoogle Scholar
Gilardi, J. D. (1992). Sex-specific foraging distributions of brown boobies in the eastern tropical Pacific. Colonial Waterbirds, 15, 148–51.CrossRefGoogle Scholar
Gillis, R. (1991). Thermal biology of two populations of red-chinned lizards (Sceloporus undulatus erythrocheilus) living in different habitats in south-central Colorado. Journal of Herpetology, 25, 18–23.CrossRefGoogle Scholar
Gillooly, J. F., Brown, J. H., West, G. B., Savage, V. M. and Charnov, E. L. (2001). Effects of size and temperature on metabolic rate. Science, 293, 2248–51.CrossRefGoogle ScholarPubMed
Gilmore, R. G., Dodrill, J. W. and Linley, P. A. (1983). Reproduction and embryonic development of the sand tiger shark, Odontaspis taurus. U.S. Fishery Bulletin, 81, 201–25.Google Scholar
Ginnett, T. F. and Demment, M. W. (1997). Sex differences in giraffe foraging behavior at two spatial scales. Oecologia, 110, 291–300.CrossRefGoogle ScholarPubMed
Ginnett, T. F. and Demment, M. W. (1999). Sexual segregation by Masai giraffes at two spatial scales. African Journal of Ecology, 37, 93–106.CrossRefGoogle Scholar
Gjerde, I. (1991). Cues in winter habitat selection by capercaillie. I. Habitat characteristics. Ornis Scandinavica, 22, 197–204.CrossRefGoogle Scholar
Goldfoot, D. A., Wallen, K., Neff, K., McBrair, D. A. and Goy, M. C. (1984). Social influences on the display of sexually dimorphic behavior in rhesus monkeys: isosexual rearing. Archives of Sexual Behaviour, 13, 395–412.CrossRefGoogle Scholar
Goldingay, R. L. and Kavanagh, R. P. (1990). Socioecology of the yellow-bellied glider (Petaurus australis) at Waratah Creek, NSW. Australian Journal of Zoology, 38, 327–41.CrossRefGoogle Scholar
Goldsworthy, S. (1995). Differential expenditure of maternal resources in Antarctic fur seals, Arctocephalus gazella, at Heard Island, southern Indian Ocean. Behavioral Ecology, 6, 218–28.CrossRefGoogle Scholar
Gompper, M. E. (1996). Sociality and asociality in white-nosed coatis (Nasua narica): foraging costs and benefits. Behavioral Ecology, 7, 254–63.CrossRefGoogle Scholar
González-Solís, J. (2004a). Sexual size dimorphism in northern giant petrels: ecological correlates and scaling. Oikos, 105(2), 247–54.CrossRefGoogle Scholar
González-Solís, J. (2004b). The regulation of incubation shifts near hatching: a timed mechanism, embryonic signalling or food availability? Animal Behaviour, 67, 663–71.CrossRefGoogle Scholar
González-Solís, J., Croxall, J. P. and Briggs, D. R. (2002a). Activity patterns of giant petrels Macronectes spp. using different foraging strategies. Marine Biology, 140, 197–204.Google Scholar
González-Solís, J., Croxall, J. P. and Wood, A. G. (2000a). Sexual dimorphism and sexual segregation in foraging strategies of northern giant petrels, Macronectes halli, during incubation. Oikos, 90, 390–8.CrossRefGoogle Scholar
González-Solís, J., Croxall, J. P. and Wood, A. G. (2000b). Foraging partitioning between giant petrels Macronectes spp. and its relationship with breeding population changes at Bird Island, South Georgia. Marine Ecology Progress Series, 204, 279–288.CrossRefGoogle Scholar
González-Solís, J., Sanpera, C. and Ruiz, X. (2002b). Metals and selenium as bioindicators of geographic and trophic segregation in giant petrels Macronectes spp. Marine Ecology Progress Series, 244, 257–64.CrossRefGoogle Scholar
Goodall, J. (1986). The Chimpanzees of Gombe. Cambridge, MA: Harvard University Press.Google Scholar
Goodman-Lowe, G. D. (1998). Diet of the Hawaiian monk seal (Monachus schauinslandi) from the Northwestern Hawaiian islands during 1991 to 1994. Marine Biology, 132, 535–46.CrossRefGoogle Scholar
Gordon, D. M. and MacCulloch, R. D. (1980). An investigation of the ecology of the map turtle, Graptemys geographica (Le Seur), in the northern part of its range. Canadian Journal of Zoology, 58, 2210–19.CrossRefGoogle Scholar
Gordon, I. J. and Illius, A. W. (1988). Incisor arcade structure and diet selection in ruminants. Functional Ecology, 2, 15–22.CrossRefGoogle Scholar
Gosler, A. G. (1987). Pattern and process in the bill morphology of the great tit Parus major. Ibis, 129, 451–76.CrossRefGoogle Scholar
Gosler, A. G. (1990). The variable niche hypothesis revisited; an analysis of intra- and inter-specific differences in bill variation in Parus. In Population Biology of Passerine Birds, eds. Blondel, J., Gosler, A. G., Lebreton, J. and McCleery, R.. Berlin: Springer-Verlag, pp. 167–74.CrossRefGoogle Scholar
Gosler, A. G. (1991). On the use of greater covert moult and pectoral muscle as measures of condition in passerines with data for the great tit Parus major. Bird Study, 31, 1–9.CrossRefGoogle Scholar
Gosler, A. G. (1996). Environmental and social determinants of winter fat storage in the great tit Parus major. Journal of Animal Ecology, 65, 1–17.CrossRefGoogle Scholar
Gosler, A. G. and Carruthers, T. D. (1994). Bill size and niche breadth in the Irish coal tit Parus ater hibernicus. Journal of Avian Biology, 25, 171–7.CrossRefGoogle Scholar
Gosler, A. G. and Carruthers, T. D. (1999). Body reserves and social dominance in the great tit Parus major in relation to winter weather in Southwest Ireland. Journal of Avian Biology, 30, 447–59.CrossRefGoogle Scholar
Gosling, L. M. (1969). Parturition and related behaviour in Coke's Hartebeest, Alcelaphus buselaphus cokei Günther. Journal of Reproduction and Fertility, Supplement, 6, 265–86.Google Scholar
, Gosling L. M. and Sutherland, W. J. (2000). Behaviour and Conservation. Cambridge, United Kingdom: Cambridge University Press.Google Scholar
Gottman, J. M. (1983). How children become friends. Monographs of the Society for Research in Child Development, 48 (3) Serial No. 201.CrossRefGoogle Scholar
Gowans, S., Whitehead, H. and Hooker, S. K. (2001). Social organization in northern bottlenose whales, Hyperoodon ampullatus: not driven by deep-water foraging? Animal Behaviour, 62, 369–77.CrossRefGoogle Scholar
Graham, T. E. and Graham, A. A. (1992). Metabolism and behavior of wintering common map turtles, Graptemys geographica, in Vermont. Canadian Field-Naturalist, 104, 517–19.Google Scholar
Graham, T. E. and Graham, A. A.Grand Usuel Larousse (1997). Paris: Larousse-Bordas.Google Scholar
Grant, G. S. and Banak, S. A. (1995). Harem structure and reproductive behaviour of Pteropus tonganus in American Samoa. Department of Marine and Wildlife Research, American Samoan Government. Biological Reports, 69, 214–44.Google Scholar
Grassi, C. (2002). Sex differences in feeding, height, and space use in Hapalemur griseus. International Journal of Primatology, 23, 677–93.CrossRefGoogle Scholar
Graves, B. M. and Duvall, D. (1987). An experimental study of aggregation and thermoregulation in prairie rattlesnakes (Crotalus viridis viridis). Herpetologica, 43, 259–64.Google Scholar
Graves, B. M. and Duvall, D. (1995). Aggregation of squamate reptiles associated with gestation, oviposition, and parturition. Herpetological Monographs, 9, 102–19.CrossRefGoogle Scholar
Greenberg, R. (1986). Competition in migrant birds in the nonbreeding season. In Current Ornithology 3, ed. Johnston, R. F.. New York and London: Plenum Press, pp. 281–307.CrossRefGoogle Scholar
Greene, H. W. (1997). Snakes: The Evolution of Mystery in Nature. Berkeley, CA: University of California Press.Google Scholar
Gregory, P. T. (1974). Patterns of spring emergence of the red-sided garter snake (Thamnophis sirtalis parietalis) in the Interlake region of Manitoba. Canadian Journal of Zoology, 52, 1063–9.CrossRefGoogle Scholar
Gregory, P. T. and Isaac, L. A. (2004). Food habits of the grass snake in southeastern England: is Natrix natrix a generalist predator? Journal of Herpetology, 38, 88–95.CrossRefGoogle Scholar
Gregory, P. T., Crampton, L. H. and Skebo, K. M. (1999). Conflicts and interactions among reproduction, thermoregulation and feeding in viviparous reptiles: are gravid snakes anorexic? Journal of Zoology, London, 248, 231–41.CrossRefGoogle Scholar
Gregory, P. T., Macartney, J. M. and Larsen, K. W. (1987). Spatial patterns and movements. In Snakes: Ecology and Evolutionary Biology, eds. Seigel, R. A., Collins, J. T. and Novak, S. S.. New York: Macmillan, pp. 366–95.Google Scholar
Griffiths, R. (1992). Sex-biased mortality in the Lesser Black-backed Gull Larus fuscus during the nestling stage. Ibis, 134, 237–44.CrossRefGoogle Scholar
Griffiths, S. W. (2003). Learned recognition of conspecifics by fishes. In Fish are Smarter than You Think: Learning in Fishes, eds. C. Brown, K. N. Laland and J. Krause. Fish and Fisheries, Special edn., 4, 256–68.CrossRef
Griffiths, S. W. and Magurran, A. E. (1998). Sex and schooling behaviour in the Trinidadian guppy. Animal Behaviour, 56, 698–93.CrossRefGoogle ScholarPubMed
Grindal, S. D., Morissette, J. L. and Brigham, R. M. (1999). Concentration of bat activity in riparian habitats over an elevational gradient. Canadian Journal of Zoology, 77, 972–7.CrossRefGoogle Scholar
Gross, J. E. (1998). Sexual segregation in ungulates: a comment. Journal of Mammalogy, 79, 1404–9.CrossRefGoogle Scholar
Gross, J. E., Alkon, P. U. and Demment, M. W. (1996). Nutritional ecology of dimorphic herbivores: digestion and passage rates in Nubian ibex. Oecologia, 107, 170–8.CrossRefGoogle ScholarPubMed
Gross, J. E., Alkon, P. U. and Demment, M. W. (1995a). Grouping patterns and spatial segregation by Nubian ibex. Journal of Arid Environments, 30, 423–39.CrossRefGoogle Scholar
Gross, J. E., Demment, M. W., Alkon, P. U. and Kotzman, M. (1995b). Feeding and chewing behaviours of Nubian ibex: compensation for sex-related differences in body size. Functional Ecology, 9, 385–93.CrossRefGoogle Scholar
Grubb, P. (1974). Mating activity and the social significance of rams in a feral sheep community. In The Behaviour of Ungulates and its Relation to Management, vol. 2, eds. Geist, V. and Walther, F.. Morges, Switzerland: IUCN No 50, pp. 457–76.Google Scholar
Grubb, P. and Jewell, P. A. (1966). Social grouping and home range in feral Soay sheep. Symposium of Zoology, Society of London, 18, 179–210.Google Scholar
Grubb, T. C. and Woodrey, M. S. (1990). Sex, age, intraspecific dominance status, and the use of food by birds wintering in temperate-deciduous and cold-coniferous woodlands: a review. Studies in Avian Biology, 13, 270–9.Google Scholar
Gruber, S. H., Nelson, D. R. and Morrissey, J. F. (1988). Patterns of activity and space utilization of lemon sharks, Negaprion brevirostris, in a shallow Bahamian lagoon. Bulletin of Marine Science, 43, 61–76.Google Scholar
Gruys, R. C. (1993). Autumn and winter movements and sexual segregation of willow ptarmigan. Arctic, 46, 228–39.CrossRefGoogle Scholar
Gueron, S., Levin, S. A. and Rubenstein, D. I. (1996). The dynamics of herds: from individuals to aggregations. Journal of Theoretical Biology, 182, 85–98.CrossRefGoogle Scholar
Gustin, K. and McCracken, G. F. (1987). Scent recognition in the Mexican free-tailed bat, Tadarida brasiliensis mexicana. Animal Behaviour, 35, 13–19.CrossRefGoogle Scholar
Hamilton, I. A. and Barclay, R. M. R. (1994). Patterns of daily torpor and day-roost selection by male and female big brown bats (Eptesicus fuscus). Canadian Journal of Zoology, 72, 744–9.CrossRefGoogle Scholar
Hamilton, W. D. (1971). Geometry of the selfish herd. Journal of Theoretical Biology, 31, 295–311.CrossRefGoogle ScholarPubMed
Hammerson, G. A. (1978). Observations on the reproduction, courtship, and aggressive behavior of the striped racer, Masticophis lateralis euryxanthus (Reptilia, Serpentes, Colubridae). Journal of Herpetology, 12, 253–5.CrossRefGoogle Scholar
Hammond, K. A., Spotila, J. R. and Standora, E. A. (1988). Basking behavior of the turtle Pseudemys scripta: effects of digestive state, acclimation temperature, sex, and season. Physiological Zoology, 61, 69–77.CrossRefGoogle Scholar
Handasyde, K. A. and Martin, R. W. (1996). Field observations on the common striped possum (Dactylopsila trivirgata) in North Queensland. Wildlife Research, 23, 755–66.CrossRefGoogle Scholar
Hanley, T. A. (1982). The nutritional basis for food selection by ungulates. Journal of Range Management, 35, 146–51.CrossRefGoogle Scholar
Hanski, I. and Gilpin, M. (1991). Metapopulation dynamics: brief history and conceptual domain. Biological Journal of the Linnean Society, 42, 3–16.CrossRefGoogle Scholar
Haramis, G. M., Nichols, J. D., Pollock, K. H. and Hines, J. E. (1986). The relationship between body mass and survival of wintering canvasbacks. Auk, 103, 506–14.Google Scholar
Harestad, A. S. and Bunnell, F. L. (1979). Home range and body weight – a reevaluation. Ecology, 60, 389–402.CrossRefGoogle Scholar
Harper, P. C. (1987). Feeding behaviour and other notes on 20 species of Procellariiformes at sea. Notornis, 34, 169–92.Google Scholar
Harrel, J. B., Allen, C. M. and Hebert, S. J. (1996). Movements and habitat use of subadult alligator snapping turtles (Macroclemys temminckii) in Louisiana. American Midland Naturalist, 135, 60–7.CrossRefGoogle Scholar
Harris, J. E. (1952). A note on the breeding season, sex ratio and embryonic development of the dogfish Scyliorhinus canicula (L.). Journal of the Marine Biological Association of the United Kingdom, 31, 269–70.CrossRefGoogle Scholar
Hart, D. R. (1983). Dietary and habitat shift with size of red-eared turtles (Pseudemys scripta) in a southern Louisiana population. Herpetologica, 39, 285–90.Google Scholar
Hashimoto, C., Furuichi, T., Tashiro, Y. (2001). What factors influence the size of chimpanzee parties in the Kalinzu Forest, Uganda? Examination of fruit abundance and number of estrous females. International Journal of Primatology, 22, 947–59.CrossRefGoogle Scholar
Hasegawa, T. (1990). Sex differences in ranging patterns. In The Chimpanzees of the Mahale Mountains, ed. Nishida, T.. Tokyo: University of Tokyo Press, pp. 99–114.Google Scholar
Hass, C. C. and Jenni, D. A. (1993). Social play among juvenile bighorn sheep: structure, development, and relationship to adult behavior. Ethology, 93, 105–16.CrossRefGoogle Scholar
Hauksson, E. and Bogason, V. (1997). Comparative feeding of grey (Halichoerus grypus) and common seals (Phoca vitulina) in coastal waters of iceland, with a note on the diet of hooded (Cystophora cristata) and harp seals (Phoca groenlandica). Journal of Northwest Atlantic Fishery Science, 22, 125–35.CrossRefGoogle Scholar
Hayes, C. L., Rubin, E. S., Jorgensen, M. C., Botta, R. A. and Boyce, W. M. (2000). Mountain lion predation of bighorn sheep in the Peninsular Ranges, California. Journal of Wildlife Management, 64, 954–9.CrossRefGoogle Scholar
Heatwole, H. (1968). Relationship of escape behavior and camouflage in anoline lizards. Copeia, 1968, 109–13.CrossRefGoogle Scholar
Hebrard, J. J. and Madsen, T. (1984). Dry season intersexual habitat partitioning by flap-necked chameleons (Chamaeleo dilepis) in Kenya. Biotropica, 16, 69–72.CrossRefGoogle Scholar
Hedd, A., Gales, R. and Brothers, N. (2001). Foraging strategies of shy albatross Thalassarche cauta breeding at Albatross Island, Tasmania, Australia. Marine Ecology Progress Series, 224, 267–82.CrossRefGoogle Scholar
Hedrick, A. V. and Temeles, E. J. (1989). The evolution of sexual dimorphism in animals: hypotheses and tests. Trends in Ecology and Evolution, 4, 136–8.CrossRefGoogle ScholarPubMed
Heinsohn, G. E. (1966). Ecology and reproduction of the Tasmanian bandicoots (Perameles gunni and Isoodon obesulus). University of California Publications in Zoology, 80, 1–96.Google Scholar
Helfman, G. S., Collette, B. B. and Facey, D. E. (1997). The Diversity of Fishes. Malden, MA: Blackwell Science.Google Scholar
Henderson, R. W. (1993). Foraging and diet in West Indian Corallus enydris (Serpentes: Boidae). Journal of Herpetology, 27, 24–8.CrossRefGoogle Scholar
Henderson, R. W. and Binder, M. H. (1980). The ecology and behavior of vine snakes (Ahaetulla, Oxybelis, Thelotornis, Uromacer): a review. Milwaukee Public Museum Contributions in Biology and Geology, 37, 1–38.Google Scholar
Henry, S. R. (1984). Social organization of the greater glider (Petauroides volans) in Victoria. In Possums and Gliders, eds. Smith, A. P. and Hume, I. D.. Chipping Norton, NSW: Surrey Beatty and Sons, pp. 221–8.Google Scholar
Hepp, G. R. and Hair, J. D. (1984). Dominance in wintering waterfowl (Anatini): effects on distribution of sexes. Condor, 86, 251–7.CrossRefGoogle Scholar
Heppel, S. S., Walters, J. R. and Crowder, L. B. (1994). Evaluating management alternatives for red-cockaded woodpeckers: a management approach. Journal of Wildlife Management, 58, 479–87.CrossRefGoogle Scholar
Herbinger, I., Boesch, C. and Rothe, H. (2001). Territory characteristics among three neighboring chimpanzee communities in the Taï National Park, Côte d'Ivoire. International Journal of Primatology, 22, 143–67.CrossRefGoogle Scholar
Herremans, M. (1997). Habitat segregation of male and female red-backed shrikes Lanius collurio and lesser grey shrikes Lanius minor in the Kalahari basin, Botswana. Journal of Avian Biology, 28, 240–8.CrossRefGoogle Scholar
Hertz, P. E., Huey, R. B. and Stevenson, R. D. (1993). Evaluating temperature regulation by field-active ectotherms: the fallacy of the inappropriate question. American Naturalist, 142, 796–818.CrossRefGoogle ScholarPubMed
Heulin, B., Surget-Groba, Y., Guiller, A., Guillaume, C. P. and Deunff, J. (1999). Comparisons of mitochondrial DNA (mtDNA) sequences (16S rRNA gene) between oviparous and viviparous strains of Lacerta vivipara: a preliminary study. Molecular Ecology, 8, 1627–31.CrossRefGoogle ScholarPubMed
Hickey, M. B. C. and Fenton, M. B. (1990). Foraging by red bats (Lasiurus borealis) – do intraspecific chases mean territoriality?Canadian Journal of Zoology, 68, 2477–82.CrossRefGoogle Scholar
Hickling, C. F. (1930). A contribution towards the life-history of the spur-dog. Journal of the Marine Biological Association of the United Kingdom, 16, 529–76.CrossRefGoogle Scholar
Hillman, J. C. (1987). Group size and association patterns of the common eland (Tragelaphus oryx). Journal of Zoology, London, 213, 641–63.CrossRefGoogle Scholar
Hinch, G. N., Lynch, J. J., Elwin, R. L. and Green, G. C. (1990). Long-term associations between Merino ewes and their offspring. Applied Animal Behaviour Science, 27, 93–103.CrossRefGoogle Scholar
Hinde, R. A. (1974). Biological bases of human social behaviour. New York: McGraw-Hill.Google Scholar
Hirth, D. H. (1977). Social behavior of white-tailed deer in relation to habitat. Wildlife Monographs, 53, 1–55.Google Scholar
Hjelm, J. and Persson, L. (2001). Size-dependent attack rate and handling capacity: inter-cohort competition in zooplantivorous fish. Oikos, 95, 520–32.CrossRefGoogle Scholar
Hjelset, A. M., Andersen, M., Gjertz, I., Lydersen, C. and Gulliksen, B. (1999). Feeding habits of bearded seals (Erignathus barbatus) from the Svalbard area, Norway. Polar Biology, 21, 186–93.CrossRefGoogle Scholar
Hobson, E. S. (1968). Predatory behaviour of some shore fishes in the Gulf of California. United States Bureau of Sport Fisheries and Wildlife Research Report, 73, 1–91.Google Scholar
Hobson, K. A., Piatt, J. F. and Pitocchelli, J. (1994). Using stable isotopes to determine seabird trophic relationships. Journal of Animal Ecology, 63, 786–98.CrossRefGoogle Scholar
Hobson, K. A., Sease, J. L., Merrick, R. L. and Piatt, J. F. (1997). Investigating trophic relationships of pinnipeds in Alaska and Washington using stable isotope ratios of nitrogen and carbon. Marine Mammal Science, 13, 114–32.CrossRefGoogle Scholar
Hocking, G. J. (1981). The population ecology of the brush-tail possum, Trichosurus vulpecula (Kerr) in Tasmania. Unpublished Masters thesis, University of Tasmania.Google Scholar
Hodum, P. J. and Hobson, K. A. (2000). Trophic relationships among Antarctic fulmarine petrels: insights into dietary overlap and chick provisioning strategies inferred from stable-isotope (d15N and d13C) analyses. Marine Ecology Progress Series, 198, 273–81.CrossRefGoogle Scholar
Hoek, W. (1992). An unusual aggregation of harbor porpoises (Phocoena phocoena). Marine Mammal Science, 8, 152–4.CrossRefGoogle Scholar
Hoelzel, A. R., Boeuf, B. J., Reiter, J. and Campagna, C. (1999). Alpha-male paternity in elephant seals. Behavioral Ecology and Sociobiology, 46, 298–306.CrossRefGoogle Scholar
Hofmann, R. R. (1989). Evolutionary steps of ecophysiological adaptation and diversification of ruminants: a comparative view of their digestive system. Oecologia, 78, 443–57.CrossRefGoogle ScholarPubMed
Hoge, A. R. and Federsoni, P. A. J. (1977). Observations on a brood of Bothrops atrox (Linnaeus, 1758): (Serpentes: Viperidae: Crotalinae). Memorias do Instituto Butantan (Sao Paulo), 40/41, 19–36.Google Scholar
Hogg, J. T. (1984). Mating in bighorn sheep: multiple creative male strategies. Science, 225, 526–9.CrossRefGoogle ScholarPubMed
Hohn, A. A. and Brownell, R. L. (1990). Harbor porpoise in central Californian waters: life history and incidental catches. Paper SC/42/SM47 presented at 42nd meeting of the scientific committee, International Whaling Commission, Nordwijk, Holland.Google Scholar
Holekamp, K. E. and Sherman, P. W. (1989). Why do male ground squirrels disperse?American Scientist, 77, 232–9.Google Scholar
Holmes, R. T. (1986). Foraging patterns of forest birds: male-female differences. Wilson Bulletin, 98, 196–213.Google Scholar
Hölzenbein, S. and Marchinton, L. (1992). Spatial integration of maturing-male white-tailed deer into the adult population. Journal of Mammalogy, 73, 326–34.CrossRefGoogle Scholar
Honda, K., Yamamoto, Y. and Tatsukawa, R. (1987). Distribution of heavy metals in Antarctic marine ecosystems. Proceedings of the NIPR Symposium of Polar Biology, 1, 184–97.Google Scholar
Hooge, P. N. and Eichenlaub, B. (1997). Animal movement extension to Arcview ver. 1.1. Alaska Science Center – Biological Science Center, U.S. Geological Survey, Anchorage, AK, USA.Google Scholar
Horak, I. G., Boomker, J. and Flamand, J. R. B. (1995). Parasites of domestic and wild animals in South Africa. XXXIV. Arthropod parasites of nyalas in north-eastern KwaZulu-Natal. Onderstepoort Journal of Veterinary Research, 62, 171–9.Google ScholarPubMed
Houde, A. E. (1997). Sex, Color, and Mate Choice in Guppies. Princeton: Princeton University Press.Google Scholar
Houston, D. L. and Shine, R. (1993). Sexual dimorphism and niche divergence: feeding habits of the Arafura filesnake. Journal of Animal Ecology, 62, 737–49.CrossRefGoogle Scholar
Houston, D. L. and Shine, R. (1994). Population demography of Arafura filesnakes (Serpentes, Acrochordidae) in tropical Australia. Journal of Herpetology, 28, 273–80.CrossRefGoogle Scholar
How, R. A. (1972). The ecology and management of Trichosurus species (Marsupialia) in New South Wales. Unpublished Ph.D. thesis, University of New England.Google Scholar
How, R. A., Barnett, J. L., Bradley, A. J., Humphreys, W. F. and Martin, R. (1984). The population biology of Pseudocheirus peregrinus in a Leptospermum laevigatum thicket. In Possums and Gliders, eds. Smith, A. P. and Hume, I. D.. Chipping Norton, NSW: Surrey Beatty and Sons, pp. 261–8.Google Scholar
Howell, D. J. (1979). Flock-foraging in nectar-feeding bats: advantages to the bats and to the host plants. American Naturalist, 114, 23–49.CrossRefGoogle Scholar
Hoyer, R. E. and Stewart, G. R. (2000). Biology of the rubber boa (Charina bottae) with emphasis on C. b. umbratica. Part I: Capture, size, sexual dimorphism, and reproduction. Journal of Herpetology, 34, 348–54.CrossRefGoogle Scholar
Hrabar, H. and du Toit, J. T. (in press). Dynamics of an introduced population of black rhinoceros (Diceros bicornis): Pilanesberg National Park, South Africa. Animal Conservation.Google Scholar
Hudson, R. J. (1985). Body size, energetics, and adaptive radiation. In Bioenergetics of Wild Herbivores, eds. Hudson, R. J. and White, R. G.. Boca Raton: CRC Press, Inc., pp. 1–24.Google Scholar
Hudson, R. J. and White, R. G. (1985). Bioenergetics of Wild Herbivores. Boca Raton: CRC Press Inc.Google Scholar
Huey, R. and Slatkin, M. (1976). Costs and benefits of lizard thermoregulation. Quarterly Review of Biology, 51, 363–84.CrossRefGoogle ScholarPubMed
Hulscher, J. B. (1985). Growth and abrasion of the oystercatcher bill in relation to dietary switches. Netherlands Journal of Zoology, 35, 124–54.CrossRefGoogle Scholar
Humple, D. L., Nur, N., Geupel, G. R. and Lynes, M. P. (2001). Female-biased sex ratio in a wintering population of ruby-crowned kinglets. Wilson Bulletin, 113, 419–24.CrossRefGoogle Scholar
Hungate, R. E. (1975). The rumen microbial ecosystem. Annual Review of Ecology and Systematics, 6, 39–66.CrossRefGoogle Scholar
Hunter, S. (1983). The food and feeding of the giant petrels Macronectes halli and M. giganteus at South Georgia. Journal of Zoology, London, 200, 521–38.CrossRefGoogle Scholar
Hunter, S. (1984). Breeding biology and population dynamics of giant petrels Macronectes at South Georgia (Aves: Procellariiformes). Journal of Zoology, London, 203, 441–60.CrossRefGoogle Scholar
Hunter, S. (1985). The role of giant petrels in the Southern Ocean ecosystem. In Antarctic Nutrient Cycles and Food Webs, eds. Siegfried, W. R., Laws, R. M. and Condy, P. R.. Berlin: Springer-Verlag, pp. 534–42.CrossRefGoogle Scholar
Hunter, S. (1987). Species and sexual isolation mechanisms in sibling species of giant petrels Macronectes. Polar Biology, 7, 295–301.CrossRefGoogle Scholar
Hunter, S. and Brooke, M. L. (1992). The diet of giant petrels Macronectes spp. at Marion Island, Southern Indian Ocean. Colonial Waterbirds, 15, 56–65.CrossRefGoogle Scholar
Huntingford, F. A. and Turner, A. K. (1987). Animal Conflict. London: Chapman and Hall.CrossRefGoogle Scholar
Hyrenbach, K. D., Fernández, P. and Anderson, D. J. (2002). Oceanographic habitats of two sympatric North Pacific albatrosses during the breeding season. Marine Ecology Progress Series, 233, 283–301.CrossRefGoogle Scholar
Illius, A. W. and Gordon, I. J. (1987). The allometry of food intake in grazing ruminants. Journal of Animal Ecology, 56, 989–99.CrossRefGoogle Scholar
Illius, A. W. and Gordon, I. J. (1992). Modelling the nutritional ecology of ungulate herbivores: evolution of body size and competitive interactions. Oecologia, 89, 428–34.CrossRefGoogle ScholarPubMed
Irwin, L. N. (1965). Diel activity and social interaction of the lizard Uta stansburiana steinegeri. Copeia, 1965, 99–101.CrossRefGoogle Scholar
Isaac, J. and Johnson, C. (2004). Sexual dimorphism and synchrony of breeding: variation in polygyny potential among populations in the common brushtail possum, Trichosurus vulpecula. Behavioral Ecology, 14, 818–22.CrossRefGoogle Scholar
Isbell, L. A., Cheney, D. L. and Seyfarth, R. M. (2002). Why vervet monkeys (Cercopithecus aethiops) live in multimale groups. In The Guenons: Diversity and Adaptation in African Monkeys, eds. Glenn, M. E. and Cords, M.. New York: Kluwer, pp. 173–88.Google Scholar
Jack, K. M. and Pavelka, M. S. M. (1997). The behavior of peripheral males during the mating season in Macaca fuscata. Primates, 38, 369–77.CrossRefGoogle Scholar
Jackes, A. D. (1973). The use of wintering ground by red deer in Ross-shire, Scotland. M.Phil. thesis, University of Edinburgh.Google Scholar
Jacklin, C. N. and Maccoby, E. E. (1978). Social behaviour at thirty-three months in same-sex and mid-sex dyads. Child Development, 49, 557–69.CrossRefGoogle Scholar
Jakimchuk, R. D., Ferguson, S. H. and Sopuck, L. G. (1987). Differential habitat use and sexual segregation in the Central Arctic caribou herd. Canadian Journal of Zoology, 65, 534–41.CrossRefGoogle Scholar
Janson, C. H. (1990). Ecological consequences of individual spatial choice in foraging brown capuchin monkeys (Cebus apella). Animal Behaviour, 38, 922–34.CrossRefGoogle Scholar
Janson, C. H. (1992). Evolutionary ecology of primate social structure. In Evolutionary Ecology and Human Behavior, eds. Smith, E. A. and Winterhalder, B.. Hawthorne, NY: Aldine.Google Scholar
Janson, C. H. (1998). Testing the predation hypothesis for vertebrate sociality: prospects and pitfalls. Behaviour, 135, 389–410.CrossRefGoogle Scholar
Jantschke, F. (1973). On the breeding and rearing of bush dogs, Speothos venaticus, at the Frankfurt Zoo. International Zoo Yearbook, 13, 141–3.CrossRefGoogle Scholar
Jaremovic, R. V. and Croft, D. B. (1991a). Social organisation of eastern grey kangaroos in southeastern New South Wales. II. Associations within mixed groups. Mammalia, 55, 543–54.Google Scholar
Jaremovic, R. V. and Croft, D. B. (1991b). Social organization of the eastern grey kangaroo (Macropodidae, Marsupialia) in southeastern New South Wales. I. Groups and group home ranges. Mammalia, 55, 169–85.Google Scholar
Jarman, P. J. (1968). The effect of the creation of Lake Kariba upon the terrestrial ecology of the middle Zambezi Valley, with particular references to the large mammals. Ph. D. thesis, University of Manchester.Google Scholar
Jarman, P. J. (1974). The social organisation of antelope in relation to their ecology. Behaviour, 48, 215–67.CrossRefGoogle Scholar
Jarman, P. J. (1983). Mating system and sexual dimorphism in large, terrestrial, mammalian herbivores. Biological Review, 58, 485–520.CrossRefGoogle Scholar
Jarman, P. J. (1989). Sexual dimorphism in Macropodoidea. In Kangaroos, Wallabies and Rat-kangaroos, eds. Grigg, G., Jarman, P. and Hume, I.. New South Wales, Australia: Surrey Beatty and Sons, pp. 433–47.Google Scholar
Jarman, P. J. (1991). Social behaviour and organization in the Macropodoidea. Advances in the Study of Behavior, 20, 1–50.CrossRefGoogle Scholar
Jarman, P. J. and Coulson, G. (1989). Dynamics and adaptiveness of grouping in macropods. In Kangaroos, Wallabies and Rat-kangaroos, eds. Grigg, G., Jarman, P. and Hume, I.. New South Wales, Australia: Surrey Beatty and Sons, pp. 527–47.Google Scholar
Jarman, P. J. and Southwell, C. J. (1986). Grouping, associations, and reproductive strategies in eastern grey kangaroos. In Ecological Aspects of Social Evolution, eds. Rubenstein, D. I. and Wrangham, R. W.. Princeton: Princeton University Press, pp. 399–428.Google Scholar
Jefferson, T. A., Stacey, P. J. and Baird, R. W. (1991). A review of killer whale interactions with other marine mammals: predation to co-existence. Mammal Review, 21, 151–80.CrossRefGoogle Scholar
Jenni, L. (1993). Structure of a brambling Fringilla montifringilla roost according to sex, age and body-mass. Ibis, 135, 85–90.CrossRefGoogle Scholar
Jensen, K. H., Jakobsen, P. J. and Kleiven, O. T. (1998). Fish kairomone regulation of internal swarm structure in Daphnia pulex (Cladocera: Crustacea). Hydrobiologia, 368, 123–7.CrossRefGoogle Scholar
Jenssen, T. A. (1970). The ethoecology of Anolis nebulosusi (Sauria, Iguanidae). Journal of Herpetology, 4, 1–38.CrossRefGoogle Scholar
Jenssen, T. A. and Nunez, S. C. (1998). Spatial and breeding relationships of the lizard, Anolis carolinensis: evidence of intrasexual selection. Behaviour, 135, 981–1003.CrossRefGoogle Scholar
Jessopp, M. J., Forcada, J., Reid, K., Trathan, P. N., Murphy, E. J. (2004). Winter dispersal of Leopard seals (Hydrurga leptonyx): environmental factors influencing demographics and seasonal abundance. Journal of Zoology, 263, 251–8.CrossRefGoogle Scholar
Jewell, P. A. (1986). Survival in a feral population of primitive sheep on St. Kilda, Outer Hebrides, Scotland. National Geographic Research, 2, 402–6.Google Scholar
Jewell, P. A. (1997). Survival and behaviour of castrated Soay sheep (Ovis aries) in a feral island population on Hirta, St. Kilda, Scotland. Journal of Zoology, London, 243, 623–36.CrossRefGoogle Scholar
Jiang, Z., Liu, B., Zeng, Y., Han, G. and Hu, H. (2000). Attracted by the same sex, or repelled by the opposite sex? – Sexual segregation in Pere David's deer. Chinese Science Bulletin, 45, 485–91.CrossRefGoogle Scholar
Johannes, R. E., Squire, L. and Graham, T. (1994). Developing a protocol for monitoring spawning aggregations of Palauan serranids to facilitate the formulation and evaluation of strategies for their management. South Pacific Forum Fisheries Agency Report 94/28. Honiara, Solomon Islands.Google Scholar
Johnson, C. N. (1983). Variations in group size and composition in red and western grey kangaroos, Macropus rufus (Desmarest) and M. fuliginosus (Desmarest). Australian Wildlife Research, 10, 25–31.CrossRefGoogle Scholar
Johnson, C. N. (1989). Grouping and the structure of association in the red-necked wallaby. Journal of Mammalogy, 70, 18–26.CrossRefGoogle Scholar
Johnson, C. N. and Bayliss, P. G. (1981). Habitat selection by sex, age and reproductive class in the red kangaroo, Macropus rufus, in western New South Wales. Australian Wildlife Research, 8, 465–74.CrossRefGoogle Scholar
, Johnson F. A. and Moore, C. T. (1996). Harvesting multiple stocks of ducks. Journal of Wildlife Management, 60, 551–9.Google Scholar
Johnson, P. M. (1980). Field observations on group compositions in the agile wallaby, Macropus agilis (Gould) (Marsupialia: Macropodidae). Australian Wildlife Research, 7, 327–31.CrossRefGoogle Scholar
Johnson, P. M. and Strahan, R. (1982). A further description of the Musky Rat-kangaroo, Hypsiprymnodon moschatus (Ramsay, 1876) (Marsupialia, Potoroidae), with notes on its biology. Australian Zoologist, 21, 27–46.CrossRefGoogle Scholar
Johnstone, G. W. (1977). Comparative feeding ecology of the giant petrels Macronectes giganteus (Gmelin) and M. halli (Mathews). In Adaptations within Antarctic Ecosystems, ed. Llano, G.. Houston: Gulf Publishing Company, pp. 647–68.Google Scholar
Jones, G. and Rydell, J. (1994). Foraging strategy and predation risk as factors influencing emergence time in echolocating bats. Philosophical Transactions of the Royal Society London, B, 346, 445–55.CrossRefGoogle Scholar
Jones, M. (1995). Tasmanian devil. In The Mammals of Australia, ed. Strahan, R.. Sydney: Reed New Holland, pp. 82–4.Google Scholar
Jones, M. and Barmuta, L. (1998). Diet overlap and relative abundance of sympatric dasyurid carnivores: a hypothesis of competition. Journal of Animal Ecology, 67, 410–21.CrossRefGoogle Scholar
Jones, M. and Barmuta, L. (2000). Niche differentiation among sympatric Australian dasyurid carnivores. Journal of Mammalogy, 81, 434–47.2.0.CO;2>CrossRefGoogle Scholar
Jones, M., Grigg, G. and Beard, L. (1997). Body temperatures and activity patterns of Tasmanian devils (Sarcophilus harrisii) and eastern quolls (Dasyurus viverrinus) through a subalpine winter. Physiological Zoology, 70, 53–60.CrossRefGoogle ScholarPubMed
Jones, R. L. (1996). Home range and seasonal movements of the turtle Graptemys flavimaculata. Journal of Herpetology, 30, 376–85.CrossRefGoogle Scholar
Jonsson, K. I. (1997). Capital and income breeding as alternative tactics of resource use in reproduction. Oikos, 78, 57–66.CrossRefGoogle Scholar
Joung, S.-J., Chen, S.-T., Clark, E., Uchida, S. and Huang, W. Y. P. (1996). The whale shark, Rhincodon typus, is a livebearer: 300 embryos found in one ‘megamamma’ supreme. Environmental Biology of Fishes, 46, 219–23.CrossRefGoogle Scholar
Jouventin, P. and Weimerskirch, H. (1990a). Long-term changes in seabird and seal populations in the Southern Ocean. In Antarctic Ecosystems: Ecological Change and Conservation, eds. Kerry, K. R. and Hempel, G.. Berlin: Springer-Verlag, pp. 208–13.CrossRefGoogle Scholar
Jouventin, P. and Weimerskirch, H. (1990b). Satellite tracking of wandering albatrosses. Nature, 343, 746–8.CrossRefGoogle Scholar
Jouventin, P., Lequette, B. and Dobson, F. S. (1999). Age-related mate choice in the wandering albatross. Animal Behaviour, 57, 1099–106.CrossRefGoogle ScholarPubMed
Kajimura, H. (1985). Opportunistic feeding by the northern fur seal (Callorhinus ursinus). In Marine Mammals and Fisheries, eds. Beddington, J. R., Beverton, R. J. H. and Lavigne, D. M.. London: George Allen and Unwin Ltd, pp. 300–18.Google Scholar
Kastelein, R. A. (2002). Walrus. In Encyclopedia of Marine Mammals, eds. Perrin, W. F., Wursig, B. and Thewissen, J. G. M.. San Diego: Academic Press, pp. 1294–300.Google Scholar
Kato, A., Watanuki, Y., Mishiumi, I., Kuroki, M., Shaughnessy, P. and Naito, Y. (2000). Variation in foraging and parental behaviour of king cormorants. Auk, 117, 718–30.CrossRefGoogle Scholar
Katsikaros, K. and Shine, R. (1997). Sexual dimorphism in the tusked frog, Adelotus brevis (Anura: Myobatrachidae): the roles of natural and sexual selection. Biological Journal of the Linnean Society, 60, 39–51.Google Scholar
Kaufman, G. W., Siniff, D. B. and Reichle, R. (1975). Colony behaviour of Weddell seals, Leptonychotes weddellii, at Hutton Cliffs, Antarctica. Rapport et Procès-verbeax des Réunions du Conseil international pour l'Exploration de la mer, 11, 228–46.Google Scholar
Kaufman, J. H. (1992). Habitat use by wood turtles in central Pennsylvania. Journal of Herpetology, 26, 315–21.CrossRefGoogle Scholar
Kaufmann, J. H. (1974). Social ethology of the whiptail wallaby, Macropus parryi, in northeastern New South Wales. Animal Behaviour, 22, 281–369.CrossRefGoogle Scholar
Keenlyne, K. D. (1972). Sexual differences in feeding habits of Crotalus horridus horridus. Journal of Herpetology, 6, 234–7.CrossRefGoogle Scholar
Kendrick, K. M., Atkins, K., Hinton, M. R.et al. (1995). Facial and vocal discrimination in sheep. Animal Behaviour, 49, 1665–76.CrossRefGoogle Scholar
Kerle, J. A. (1983). The population biology of the Northern brushtail possum. Unpublished Ph.D. thesis, Macquarie University.Google Scholar
Kerle, J. A. (1984). Variation in the ecology of Trichosurus: its adaptive significance. In Possums and Gliders, eds. Smith, A. P. and Hume, I. D.. Chipping Norton, NSW: Surrey Beatty and Sons, pp. 115–28.Google Scholar
Kerle, J. A., McKay, G. M. and Sharman, G. B. (1991). A systematic analysis of the brushtail possum, Trichosurus vulpecula (Kerr, 1792) (Marsupialia: Phalangeridae). Australian Journal of Zoology, 39, 313–31.CrossRefGoogle Scholar
Kerth, G. and Reckardt, K. (2003). Information transfer about roosts in female Bechstein's bats: an experimental field study. Proceedings of the Royal Society London, B, 270, 511–15.CrossRefGoogle Scholar
Kerth, G., Mayer, F. and Konig, B. (2000). Mitochondrial DNA (mtDNA) reveals that female Bechstein's bats live in closed societies. Molecular Ecology, 9, 793–800.CrossRefGoogle ScholarPubMed
Kerth, G., Wagner, M. and Konig, B. (2001a). Roosting together, foraging apart: information transfer about food is unlikely to explain sociality in female Bechstein's bats (Myotis bechsteinii). Behavioural Ecology and Sociobiology, 50, 283–91.CrossRefGoogle Scholar
Kerth, G., Weissman, K. and König, B. (2001b). Day roost selection in female Bechstein's bats (Myotis bechsteinii): a field experiment to determine the influence of roost temperature. Oecologia, 126, 1–9.CrossRefGoogle Scholar
Ketterson, E. D. and Nolan, V. Jr (1983). The evolution of differential bird migration. In Current Ornithology 1, ed. Johnston, R. F.. New York: Plenum Press.CrossRefGoogle Scholar
Kie, J. G. and Bowyer, R. T. (1999) Sexual segregation in white-tailed deer: density-dependent changes in use of space, habitat selection, and dietary niche. Journal of Mammalogy, 80, 1004–20.CrossRefGoogle Scholar
King, J. E. (1983). Seals of the World, 2nd edn. Oxford: Oxford University Press.Google Scholar
King, R. B. (1986). Population ecology of the Lake Erie water snake, Nerodia sipedon insularum. Copeia, 1986, 757–72.CrossRefGoogle Scholar
King, R. B. (1989). Sexual dimorphism in snake tail length: sexual selection, natural selection, or morphological constraint?Biological Journal of the Linnean Society, 38, 133–54.CrossRefGoogle Scholar
King, R. B. (1993). Microgeographic, historical, and size-correlated variation in water snake diet composition. Journal of Herpetology, 27, 90–4.CrossRefGoogle Scholar
Kinzey, W. G. (1977). Diet and feeding behavior of Callicebus torquatus. In Primate Ecology, ed. Clutton-Brock, T. H.. Cambridge: Cambridge University Press, pp. 127–51.Google Scholar
Kirsch, J. A. W., Lapointe, F. J. and Springer, M. S. (1997). DNA-hybridisation studies of marsupials and their implications for metatherian classification. Australian Journal of Zoology, 45, 211–80.CrossRefGoogle Scholar
Kissner, K. J., Forbes, M. R. L. and Secoy, D. M. (1997). Rattling behavior of prairie rattlesnakes (Crotalus viridis viridis, Viperidae) in relation to sex, reproductive status, body size, and body temperature. Ethology, 103, 1042–50.CrossRefGoogle Scholar
Kissner, K. J., Weatherhead, P. J. and Francis, C. M. (2003). Sexual size dimorphism and timing of spring migration in birds. Journal of Evolutionary Biology, 16, 154–62.CrossRefGoogle ScholarPubMed
Kitchen, D. W. (1974). Social behavior and ecology of the pronghorn. Wildlife Monographs, 38, 1–96.Google Scholar
Kitching, J. A. and Ebling, F. J. (1967). Ecological studies in Lough Ine. Advances in Ecological Research, 4, 197–291.CrossRefGoogle Scholar
Kleiber, M. (1975). An Introduction to Animal Energetics. Huntington, NY: R. E. Kreiger Publishing Co.Google Scholar
Kleiman, D. G. (1980). The sociobiology of captive propagation. In Conservation Biology: An Evolutionary-Ecological Perspective, eds. Soule, M. and Wilcox, B. A.. Sunderland, Massachusetts: Sinauer Associates, Inc., pp. 243–61.Google Scholar
Kleinenberg, S. E., Yablokov, A. V., Bel'kovich, B. M. and Tarasevich, M. N. (1964). Beluga (Delphinapterus leucas): investigation of the species. Translated by Israel Progr. Sci. Translat., Jerusalem 1969. Moscow: Academy Nauk USSR, p. 376.Google Scholar
Klimley, A. P. (1985). Schooling in Sphyrna lewini, a species with low risk of predation: a non-egalitarian state. Zeitschrift für Tierpsychologie, 70, 297–319.CrossRefGoogle Scholar
Klimley, A. P. (1987). The determinants of sexual segregation in the scalloped hammerhead shark, Sphyrna lewini. Environmental Biology of Fishes, 18, 27–40.CrossRefGoogle Scholar
Kodric-Brown, A. (1990). Mechanisms of sexual selection – insights from fishes. Annales Zoologici Fennici, 27, 87–100.Google Scholar
Kohlmann, S. G., Müller, D. M. and Alkon, P. U. (1996). Antipredator constraints on lactating nubian ibexes. Journal of Mammalogy, 77, 1122–31.CrossRefGoogle Scholar
Komdeur, J. and Deerenberg, C. (1997). The importance of social behavior studies for conservation. In Behavioral Approaches to Conservation in the Wild, eds. Clemmons, J. R. and Buchholz, R.. Cambridge, United Kingdom: Cambridge University Press, pp. 262–76.Google Scholar
Komers, P. E., Messier, F. and , Gates C. C. (1993). Group structure in wood bison: nutritional and reproductive determinants. Canadian Journal of Zoology, 71, 1367–71.CrossRefGoogle Scholar
Koopman, K. F. (1993). Bats. In Mammal Species of the World: Taxonomic and Geographic Reference, eds. Wilson, D. E. and Reeder, D. M.. Washington, DC: Smithsonian Institution Press, pp. 137–241.Google Scholar
Koskimies, J. (1957). Flocking behaviour in capercaillie Tetrao urogallus and blackgame Lyrurus tetrix. Papers on Game Research Published by the Finish Game Foundation, 18, 1–31.Google Scholar
Kovacs, K. M., Lydersen, C., Hammill, M. and Lavigne, D. M. (1996). Reproductive effort of male hooded seals (Cystophora cristata): estimates from mass loss. Canadian Journal of Zoology, 74, 1521–30.CrossRefGoogle Scholar
Krause, J. (1994). The influence of food competition and predation risk on size-assortative shoaling in juvenile chub (Leuciscus cephalus). Ethology, 96, 105–16.CrossRefGoogle Scholar
Krause, J. and Godin, J.-G. J. (1994). Shoal choice in banded killifish (Fundulus diaphanus, Teleostei, Cyprinodontidae) – Effects of predation risk, fish size, species composition and size of shoals. Ethology, 98, 128–36.CrossRefGoogle Scholar
Krause, J. and Godin, J.-G. J. (1996a). Influence of parasitism on shoal choice in the banded killifish (Fundulus diaphanus, Teleostei, Cyprinodontidae). Ethology, 102, 40–9.CrossRefGoogle Scholar
Krause, J. and Godin, J.-G. J. (1996b). Phenotypic variability within and between fish shoals. Ecology, 77, 1586–91.CrossRefGoogle Scholar
Krause, J. and Ruxton, G. D. (2002). Living in Groups. Oxford: Oxford University Press.Google Scholar
Krause, J., Butlin, R., Peuhkuri, N. and Pritchard, V. L. (2000a). The social organisation of fish shoals: a test of the predictive power of laboratory experiments for the field. Biological Reviews, 75, 477–501.CrossRefGoogle Scholar
Krause, J., Godin, J.-G. J. and Brown, D. (1996). Size-assortativeness in multi-species fish shoals. Journal of Fish Biology, 49, 221–5.CrossRefGoogle Scholar
Krause, J., Hoare, D. J., Croft, D., Lawrence, J., Ward, A., Ruxton, G. D., Godin, J. G. J. and James, R. (2000b). Fish shoal composition: mechanisms and constraints. Proceedings of the Royal Society of London Series B, Biological Sciences, 267, 2011–17.CrossRefGoogle Scholar
Krause, J., Loader, S. P., McDermott, J. and Ruxton, G. D. (1998). Refuge use by fish as a function of body length-related metabolic expenditure and predation risks. Proceedings of the Royal Society of London Series B, Biological Sciences, 265, 2373–9.CrossRefGoogle Scholar
Krebs, C. J. (1989). Ecological Methodology. New York: Harper Collins Publishers.Google Scholar
Krohmer, R. W. and Aldridge, R. D. (1985). Female reproductive cycle of the lined snake (Tropidoclonium lineatum). Herpetologica, 41, 39–44.Google Scholar
Krützen, M., Sherwin, W. B., Berggren, P. and Gales, N. (2004). Population structure in an inshore cetacean revealed by microsatellite and mtDNA analysis: bottlenose dolphins (Tursiops sp.) in Shark Bay Western Australia. Marine Mammal Science, 20, 28–47.CrossRefGoogle Scholar
Kruuk, H. (1972) The Spotted Hyena. Chicago: University of Chicago Press.Google Scholar
Kunz, T. H. (1974). Feeding ecology of a temperate insectivorous bat (Myotis velifer). Ecology, 55, 693–711.CrossRefGoogle Scholar
Kunz, T. H. (1982). Roosting ecology of bats. In Ecology of Bats, ed. Kunz, T. H.. New York: Plenum Press, pp. 1–55.CrossRefGoogle Scholar
Kunz, T. H. and Lumsden, L. F. (2003). Ecology of cavity and foliage roosting bats. In Bat Ecology, eds. Kunz, T. H. and Fenton, M. B.. Chicago: University of Chicago Press, pp. 3–89.Google Scholar
Lack, D. (1968). Ecological Adaptations for Breeding in Birds. London: Methuen and Co.Google Scholar
, Lagarde F., Bonnet, X., Corbin, J., Henen, B. and Nagy, K. (2002). A short spring before a long jump: the ecological challenge to the steppe tortoise (Testudo horsfieldi). Canadian Journal of Zoology, 80, 493–502.Google Scholar
Lagarde, F., Bonnet, X., Henen, B.et al. (2003). Sex divergence in space utilisation in the steppe tortoise (Testudo horsfieldi). Canadian Journal of Zoology, 81, 380–7.CrossRefGoogle Scholar
LaGory, K. E., Bagshaw, C. III and Brisbin, I. L. Jr (1991). Niche differences between male and female white-tailed deer on Ossabaw Island, Georgia. Applied Animal Science, 29, 205–14.CrossRefGoogle Scholar
Lahanas, P. N. (1982). Aspects of the life history of the southern black-knobbed sawback, Graptemys nigrinoda delticola Folkerts and Mount. M.Sc. thesis, Auburn, AL: Auburn University.Google Scholar
Laidlaw, W. S., Hutchings, S. and Newell, G. R. (1996). Home range and movement patterns of Sminthopsis leucopus (Marsupialia: Dasyuridae) in coastal dry Heathland, Anglesea, Victoria. Australian Mammalogy, 19, 1–9.Google Scholar
Lailvaux, S. P., Alexander, G. J. and Whiting, M. J. (2003). Sex-based differences and similarities in locomotor performance, thermal preferences, and escape behaviour in the lizard Platysaurus intermedius wilhelmi. Physiological and Biochemical Zoology, 76, 511–21.CrossRefGoogle ScholarPubMed
Lamb, T. (1984). The influence of sex and breeding condition on microhabitat and diet in the pig frog Rana grylio. American Midland Naturalist, 111, 311–18.CrossRefGoogle Scholar
Lande, R. and Barrowclough, G. F. (1987). Effective population size, genetic variation, and their use in population management. In Viable Populations for Conservation, ed. Soule, M. E.. Cambridge, United Kingdom: Cambridge University Press, pp. 87–123.CrossRefGoogle Scholar
Landeau, L. and Terborgh, J. (1986). Oddity and the confusion effect in predation. Animal Behaviour, 34, 1372–80.CrossRefGoogle Scholar
Landys-Ciannelli, M. M., Piersma, T. and Jukema, J. (2003). Strategic size changes of internal organs and muscle tissue in the bar-tailed godwit during fat storage on a spring stopover site. Functional Ecology, 17, 151–9.CrossRefGoogle Scholar
Langman, V. A. (1977). Cow-calf relationships in giraffe (Giraffa camelopardalis giraffa). Zeitschrift für Tierpsychology, 43, 264–86.Google Scholar
Laska, A. L. (1970). The structural niche of Anolis scripta on Inagua. Breviora, 349, 1–6.Google Scholar
Latta, S. C. and Faaborg, J. (2002). Demographic and population responses of Cape May warbler wintering in multiple habitats. Ecology, 83, 2502–15.CrossRefGoogle Scholar
Launchbaugh, K. L., Provenza, F. D. and Pfister, J. A. (2001). Herbivore response to anti-quality factors in forages. Journal of Range Management, 54, 431–40.CrossRefGoogle Scholar
Lausen, C. L. and Barclay, R. M. R. (2002). Roosting behaviour and roost selection of female big brown bats (Eptesicus fuscus) roosting in rock crevices in south-eastern Alberta. Canadian Journal of Zoology, 80, 1069–76.CrossRefGoogle Scholar
Lausen, C. L. and Barclay, R. M. R. (2003). Thermoregulation and roost selection by reproductive female big brown bats (Eptesicus fuscus) roosting in rock crevices. Journal of Zoology, London, 260, 235–44.CrossRefGoogle Scholar
Lawrence, A. L. (1990). Mother-daughter and peer relationships of Scottish hill sheep. Animal Behaviour, 39, 481–6.CrossRefGoogle Scholar
Laws, R. M. (1993). Antarctic Seals: Research Methods and Techniques, 1st edn., Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Lazenby-Cohen, K. A. and Cockburn, A. (1988). Lek promiscuity in a semelparous mammal, Antechinus stuartii (Marsupialia: Dasyuridae)?Behavioral Ecology and Sociobiology, 22, 195–202.CrossRefGoogle Scholar
Lazenby-Cohen, K. A. and Cockburn, A. (1991). Social and foraging components of the home range in Antechinus stuartii (Dasyuridae: Marsupialia). Australian Journal of Ecology, 16, 301–8.CrossRefGoogle Scholar
Boeuf, B. J. (1971). The aggression of the breeding bulls. Natural History, 70, 83–94.Google Scholar
Le Boeuf, B. J. (1991). Pinniped mating systems on land, ice and in the water: emphasis on the Phocidae. In Behaviour of Pinnipeds, ed. Renouf, D.. London: Chapman and Hall, pp. 45–65.CrossRefGoogle Scholar
Le Boeuf, B. J. and Crocker, D. E. (1996). Diving behaviour of elephant seals: implications for predator avoidance. In Great White Sharks: The Biology of Carcharodon carcharias, eds. Klimley, A. P. and Ainley, D. G.. Berkley: University of California Press, pp. 193–206.Google Scholar
Le Boeuf, B. J. and Laws, R. M. (1994). Elephant seals: An introduction to the genus. In Elephant Seals: Population Ecology, Behaviour, and Physiology, eds. Boeuf, B. J. and Laws, R. M.. Berkley: University of California Press, pp. 1–26.Google Scholar
Le Boeuf, B. J. and Reiter, J. (1988). Lifetime reproductive success in northern elephant seals. In Reproductive Success: Studies of Individual Variation in Contrasting Breeding Systems, ed. Clutton-Brock, T. H.. Chicago: University of Chicago Press, pp. 344–62.Google Scholar
Le Boeuf, B. J., Crocker, D. E., Blackwell, S. B., Morris, P. A. and Thorson, P. H. (1993). Sex differences in foraging in northern elephant seals. In Marine Mammals: Advances in Behavioural and Population Biology, ed. Boyd, I. L.. London: Oxford University Press.Google Scholar
Boeuf, B. J., Crocker, D. E., Costa, D. P.et al. (2000). Foraging ecology of northern elephant seals. Ecological Monographs, 70, 353–82.CrossRefGoogle Scholar
Boeuf, B. J., Morris, P. A., Blackwell, S. B., Crocker, D. E. and Costa, D. P. (1996). Diving behavior of juvenile northern elephant seals. Canadian Journal of Zoology, 74, 1632–44.CrossRefGoogle Scholar
Pendu, Y., Guilhem, C., Briedermann, L., Maublanc, M.-L. and Gerard, J.-F. (2000). Interactions and associations between age and sex classes in mouflon sheep (Ovis gmelini) during winter. Behavioural Processes, 52, 97–107.CrossRefGoogle ScholarPubMed
Leaper, C. (1994). Exploring the consequences of gender segregation on social relationships. In Childhood Gender Segregation: Causes and Consequences, ed. Leaper, C.. San Francisco: Jossey-Bass, pp. 67–86.Google Scholar
Lee, A. and Cockburn, A. (1985). Evolutionary Ecology of Marsupials. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Lee, A., Woolley, P. and Braithwaite, R. (1982). Life history strategies of dasyurid marsupials. In Carnivorous Marsupials, ed. Archer, M.. Sydney: Royal Zoological Society of New South Wales, pp. 1–11.Google Scholar
Lee, D. S., Franz, R. and Sanderson, R. A. (1975). A note on the feeding habits of male Barbour's map turtles. Florida Field Naturalist, 3, 45–6.Google Scholar
Lee, Y. F. and McCracken, G. F. (2002). Foraging activity and food resource use of Brasilian free-tailed bats, Tadarida brasiliensis (Molossidae). Ecoscience, 9, 306–13.CrossRefGoogle Scholar
Legault, F. and Strayer, F. F. (1991). Genèse de la ségrégation sexuelle et différences comportementales chez des enfants d'âge préscolaire. Behaviour, 119, 285–301.CrossRefGoogle Scholar
Legler, J. M. L. (1985). Australian chelid turtles: reproductive patterns in wide-ranging taxa. In The Biology of Australasian Frogs and Reptiles, eds. Grigg, G. C., Shine, R. and Ehmann, H.. Sydney: Royal Zoological Society of New South Wales, pp. 117–23.Google Scholar
Lemos-Espinal, J. A., Smith, G. R. and Ballinger, R. E. (1997). Thermal ecology of the lizard, Sceloporus gadoviae, in an arid tropical scrub forest. Journal of Arid Environments, 35, 311–19.CrossRefGoogle Scholar
Lesage, V., Hammill, M. O. and Kovacs, K. M. (2001). Marine mammals and the community structure of the Estuary and Gulf of St Lawrence, Canada: evidence from stable isotope analysis. Marine Ecology Progress Series, 210, 203–21.CrossRefGoogle Scholar
Lettevall, E., Richter, C., Jaquet, N.et al. (2002). Social structure and residency in aggregations of male sperm whales. Canadian Journal of Zoology, 80, 1189–96.CrossRefGoogle Scholar
Leuzinger, Y. and Brossard, C. (1994). Répartition de M. daubentonii en fonction du sexes et de la période de l'année dans le Jura Bernois. Résultats préliminaires. Mitteilunngen der Naturforschenden Gesellschaft Schaffhausen, 39, 135–43.Google Scholar
Lewis, S., Benvenuti, S., Dall'Antonia, L.et al. (2002). Sex-specific foraging behaviour in a monomorphic seabird. Proceeding of the Royal Society of London, 269, 1687–93.CrossRefGoogle Scholar
Lewis, S. E. (1992). Behaviour of Peter's tent-making bat, Uroderma bilobatum, at maternity roosts in Costa Rica. Journal of Mammalogy, 73, 541–6.CrossRefGoogle Scholar
Lewis, S. E. (1996). Low roost-site fidelity in pallid bats: associated factors and effect on group stability. Behavioural Ecology and Sociobiology, 39, 335–44.CrossRefGoogle Scholar
Li, J., Luan, Y., Sun, I., Zhao, D. and Diao, Y. (1990). Studies on some problems of Agkistrodon shedaoensis population due to seasonal changes (in Chinese). In From Water Onto Land, ed. Zhao, E.. China Forestry Press, pp. 273–6.Google Scholar
Lillegraven, J. A., Kielan-Jaworowska, Z. and Clemens, W. A. (1979). Mesozoic Mammals: The First Two-Thirds of Mammalian History. Berkeley: University of California Press.Google Scholar
Lillywhite, H. B. and Henderson, R. W. (1993). Behavioral and functional ecology of arboreal snakes. In Snakes: Ecology and Behavior, eds. Seigel, R. A. and Collins, J. T.. New York: McGraw-Hill, pp. 1–48.Google Scholar
Lindeman, P. V. (2003). Sexual difference in habitat use of Texas map turtles (Emydidae: Graptemys versa) and its relationship to size dimorphism and diet. Canadian Journal of Zoology, 81, 1185–91.CrossRefGoogle Scholar
Ling, J. K. (2002). Australian sea lion. In Encyclopedia of Marine Mammals, eds. Perrin, W. F., Wursig, B. and Thewissen, J. G. M.. San Diego: Academic Press.Google Scholar
Lingle, S. (2000). Seasonal variation in coyote feeding behaviour and mortality of white-tailed deer and mule deer. Canadian Journal of Zoology, 78, 85–99.CrossRefGoogle Scholar
Lingle, S. (2001). Anti-predator strategies and grouping patterns in white-tailed deer and mule deer. Ethology, 107, 295–314.CrossRefGoogle Scholar
Lingle, S. (2002). Coyote predation and habitat segregation of white-tailed deer and mule deer. Ecology, 83, 2037–48.CrossRefGoogle Scholar
Lingle, S. and Pellis, S. M. (2002). Fight or flight? Antipredator behavior and the escalation of coyote encounters with deer. Oecologia, 131, 154–64.CrossRefGoogle ScholarPubMed
Lingle, S. and Wilson, W. F. (2001). Detection and avoidance of predators in white-tailed deer (Odocoileus virginianus) and mule deer (O. hemionus). Ethology, 107, 125–47.CrossRefGoogle Scholar
Lister, B. C. (1976). The nature of niche expansion in West Indian Anolis lizards I. Ecological consequences of reduced competition. Evolution, 30, 659–76.CrossRefGoogle ScholarPubMed
Lister, B. C. and Aguayo, A. G. (1992). Seasonality, predation, and the behaviour of a tropical mainland anole. Journal of Animal Ecology, 61, 717–33.CrossRefGoogle Scholar
Loison, A., Gaillard, J.-M., Pelabon, C. and Yoccoz, N. G. (1999). What factors shape sexual size dimorphism in ungulates?Evolutionary Ecology Research, 1, 611–33.Google Scholar
Long, J. A. and Pellegrini, A. D. (2003). Studying change in bullying and dominance with structural equation modeling. School Psychology Review, 32, 401–17.Google Scholar
Ornat, Lopez A. and Greenberg, R. (1990). Sexual segregation by habitat in migratory warblers in Quintana Roo, Mexico. Auk, 107, 539–43.Google Scholar
Lott, D. F. and Minta, S. C. (1983). Random individual association and social group instability in American bison (Bison bison). Zeitschrift für Tierpsychology, 61, 153–72.CrossRefGoogle Scholar
Lovern, M. B. (2000). Behavioral ontogeny in free-ranging juvenile male and female green anoles, Anolis carolinensis, in relation to sexual selection. Journal of Herpetology, 34, 274–81.CrossRefGoogle Scholar
Low, B. S. (2000). Why Sex Matters: A Darwinian Look at Human Behavior. Princeton: Princeton University Press.Google Scholar
Lowry, L. F., Burns, J. J. and Nelson, R. R. (1987). Polar bear, Ursus maritimus, predation on belugas, Delphinapterus leucas, in the Bering and Chukchi Seas. Canadian Field-Naturalist, 101, 141–6.Google Scholar
Lowry, L. F., Frost, K. J. and Seaman, G. A. (1985). Investigations of Beluga Whales in Coastal Waters of Western and Northern Alaska – III: Food Habits. Fairbanks (Alaska): Alaska Department of Fish and Game, p. 24.Google Scholar
Lue, K. Y. and Chen, T. H. (1999). Activity, movement patterns, and home range of the yellow-margined box turtle (Cuora flavomarginata) in northern Taiwan. Journal of Herpetology, 33, 590–600.CrossRefGoogle Scholar
Luiselli, L. and Agrimi, U. (1991). Composition and variation of the diet of Vipera aspis francisciredi in relation to age and reproductive stage. Amphibia-Reptilia, 12, 137–44.CrossRefGoogle Scholar
Luiselli, L. and Angelici, F. M. (1998). Sexual size dimorphism and natural history traits are correlated with intersexual dietary divergence in royal pythons (Python regius) from the rainforests of southeastern Nigeria. Italian Journal of Zoology, 65, 183–5.CrossRefGoogle Scholar
Luiselli, L., Akani, G. C. and Capizzi, D. (1999). Is there any interspecific competition between dwarf crocodiles (Osteolaemus tetraspis) and Nile monitors (Varanus niloticus ornatus) in the swamps of central Africa? A study from southeastern Nigeria. Journal of Zoology, London, 247, 127–31.CrossRefGoogle Scholar
Luiselli, L., Capula, M. and Shine, R. (1996). Reproductive output, costs of reproduction, and ecology of the smooth snake, Coronella austriaca, in the eastern Italian Alps. Oecologia, 106, 100–10.CrossRefGoogle ScholarPubMed
Luiselli, L., Capula, M. and Shine, R. (1997). Food habits, growth rates, and reproductive biology of grass snakes, Natrix natrix (Colubridae) in the Italian Alps. Journal of Zoology, London, 241, 371–80.CrossRefGoogle Scholar
Lunn, N. J. and Arnould, J. P. Y. (1997). Maternal investment in Antarctic fur seals: evidence for equality in the sexes. Behavioral Ecology and Sociobiology, 40, 351–62.CrossRefGoogle Scholar
Lunney, D. (1995). White-footed Dunnart. In The Mammals of Australia, ed. Strahan, R.. Sydney: Reed New Holland, pp. 143–5.Google Scholar
Lunney, D. and Leary, T. (1989). Movement patterns of the white-footed dunnart, Sminthopsis leucopus (Marsupialia: Dasyuridae), in a logged, burnt forest on the south coast of New South Wales. Australian Wildlife Research, 16, 207–16.CrossRefGoogle Scholar
Lunney, D., Matthews, A. and Grigg, J. (2001). The diet of Antechinus agilis and A. swainsonii in unlogged and regenerating sites in Mumbulla State Forest, south-eastern New South Wales. Wildlife Research, 28, 459–64.CrossRefGoogle Scholar
Lunney, D., O'Connell, M. and Sanders, J. (1989). Habitat of the white-footed dunnart, Sminthopsis leucopus (Gray) (Dasyuridae: Marupialia), in a logged, burnt forest near Bega, New South Wales. Australian Journal of Ecology, 14, 335–44.CrossRefGoogle Scholar
Lyderson, C., Hammill, M. O. and Kovacs, K. M. (1994). Activity of lactating ice-breeding grey seals, Halichoerus grypus, from the Gulf of St. Lawrence, Canada. Animal Behaviour, 48, 1417–25.CrossRefGoogle Scholar
Lyle, J. M. (1983). Food and feeding habits of the lesser spotted dogfish, Scyliorhinus canicula (L.) in Isle of Man waters. Journal of Fish Biology, 23, 725–38.CrossRefGoogle Scholar
Lynch, J. F., Morton, E. S. and Voort, M. E. (1985). Habitat segregation between the sexes of wintering hooded warblers, Wilsonia citrina. Auk, 102, 714–21.Google Scholar
MacArthur, R. A., , Geist V. and Johnston, R. H. 1982. Cardiac and behavioral responses of mountain sheep to human disturbance. Journal of Wildlife Management, 46, 351–8.CrossRefGoogle Scholar
Maccoby, E. E. (1998). The Two Sexes: Growing Up Apart, Coming Together. Cambridge, MA, Harvard University Press.Google Scholar
MacFarlane, A. M. and Coulson G. (in press). Synchrony and timing of breeding influences sexual segregation in western grey and red Kangaroos (Macropus Fuliginosus and M. rufus). Journal of Zoology, London.
Macdonald, D. (2001). The New Encyclopedia of Mammals. Oxford: Oxford University Press.Google Scholar
Madsen, T. (1983). Growth rates, maturation and sexual size dimorphism in a population of grass snakes, Natrix natrix, in southern Sweden. Oikos, 40, 277–82.CrossRefGoogle Scholar
Madsen, T. (1984). Movements, home range size and habitat use of radio-tracked grass snakes (Natrix natrix) in southern Sweden. Copeia, 1984, 707–13.CrossRefGoogle Scholar
Madsen, T. and Shine, R. (1993a). Costs of reproduction in a population of European adders. Oecologia, 94, 488–95.CrossRefGoogle Scholar
Madsen, T. and Shine, R. (1993b). Phenotypic plasticity in body sizes and sexual size dimorphism in European grass snakes. Evolution, 47, 321–5.CrossRefGoogle Scholar
Madsen, T. and Shine, R. (1996). Seasonal migration of predators and prey: pythons and rats in tropical Australia. Ecology, 77, 149–56.CrossRefGoogle Scholar
Madsen, T. and Shine, R. (2000). Energy versus risk: costs of reproduction in free-ranging pythons in tropical Australia. Austral Ecology, 25, 670–5.CrossRefGoogle Scholar
Madsen, T., Shine, R., Loman, J. and Håkansson, T. (1993). Determinants of mating success in male adders, Vipera berus. Animal Behaviour, 45, 491–9.CrossRefGoogle Scholar
Magnusson, W. E. (1993). Body temperatures of field-active Amazonian savanna lizards. Journal of Herpetology, 27, 53–8.CrossRefGoogle Scholar
Magurran, A. E. (1990). The adaptive significance of schooling as an antipredator defense in fish. Annales Zoologici Fennici, 27, 51–66.Google Scholar
Magurran, A. E. and Garcia, M. (2000). Sex differences in behaviour as an indirect consequence of mating system. Journal of Fish Biology, 57, 839–57.CrossRefGoogle Scholar
Magurran, A. E., Seghers, B. H., Shaw, P. W. and Carvalho, G. R. (1994). Schooling preferences for familiar fish in the Guppy, Poecilia reticulata. Journal of Fish Biology, 45, 401–6.CrossRefGoogle Scholar
Mahmoud, I. Y. (1969). Comparative ecology of the kinosternid turtles of Oklahoma. Southwestern Naturalist, 14, 31–66.CrossRefGoogle Scholar
Main, M. B. (1994). Advantages of habitat selection and sexual segregation in mule and white-tailed deer. Ph.D. thesis, Corvallis, OR, USA: Oregon State University.Google Scholar
Main, M. B. (1998). Sexual segregation in ungulates: a reply. Journal of Mammalogy, 79, 1410–15.CrossRefGoogle Scholar
Main, M. B. and Coblentz, B. E. (1990). Sexual segregation among ungulates: a critique. Wildlife Society Bulletin, 18, 204–10.Google Scholar
Main, M. B. and Coblentz, B. E. (1996). Sexual segregation in Rocky Mountain mule deer. Journal of Wildlife Management, 60, 97–507.CrossRefGoogle Scholar
Main, M. B., Weckerly, F. W. and Bleich, V. C. (1996). Sexual segregation in ungulates: new directions for research. Journal of Mammalogy, 77, 449–61.CrossRefGoogle Scholar
Mann, J. and Barnett, H. (1999). Lethal tiger shark (Galeocerdo cuvier) attack on bottlenose dolphin (Tursiops sp.) calf: defense and reactions by the mother. Marine Mammal Science, 15, 568–75.CrossRefGoogle Scholar
Mann, J., Connor, R. C., Barre, L. M. and Heithaus, M. R. (2000a). Female reproductive success in wild bottlenose dolphins (Tursiops sp.): Life history, habitat, provisioning, and group size effects. Behavioral Ecology, 11, 210–19.CrossRefGoogle Scholar
Mann, J., Connor, R. C., Tyack, P. L. and Whitehead, H. (2000b). Cetacean Societies: Field Studies of Dolphins and Whales. Chicago: University of Chicago Press.Google Scholar
Mansergh, I. and Broome, L. (1994). The Mountain Pygmy-Possum of the Australian Alps. Kensington: New South Wales University Press.Google Scholar
Mansergh, I. and Scotts, D. (1989). Habitat continuity and social organizations of the Mountain Pygmy-possum restored by tunnel. Journal of Wildlife Management, 53, 701–7.CrossRefGoogle Scholar
Mansergh, I. and Scotts, D. (1990). Aspects of the life history and breeding biology of the Mountain Pygmy-possum, Burramys parvus, (Marsupialia: Burramyidae) in alpine Victoria. Australian Mammalogy, 13, 179–91.Google Scholar
Mansergh, I., Baxter, B., Scotts, D., Brady, T. and Jolley, D. (1990). Diet of the mountain pygmy-possum, Burramys parvus (Marsupialia: Burramyidae) and other small mammals in the alpine environment at Mt. Higginbotham, Victoria. Australian Mammalogy, 13, 167–77.Google Scholar
Marchal, C., Gerard, J.-F., Boisaubert, B. and Bideau, E. (1998). Instability and diurnal variation in size of winter groupings of field roe deer. Revue d'Ecologie (Terre Vie), 53, 59–68.Google Scholar
Marchant, S. and Higgins, P. J. (1993). Handbook of Australian, New Zealand and Antarctic Birds. vol. 2, Melbourne, Oxford University Press.Google Scholar
Marquiss, M. (1980). Habitat and diet of male and female hen harriers in Scotland in winter. British Birds, 73, 555–60.Google Scholar
Marra, P. P. (2000). The role of behavioral dominance in structuring patterns of habitat occupancy in a migrant bird during the nonbreeding season. Behavioural Ecology, 11, 299–308.CrossRefGoogle Scholar
Marra, P. P. and Holberton, R. L. (1998). Corticosterone levels as indicators of habitat quality: effects of habitat segregation in a migratory bird during the non-breeding season. Oecologia, 116, 284–92.CrossRefGoogle Scholar
Marra, P. P. and Holmes, R. T. (1997). Avian removal experiments: do they test for habitat saturation or female availability?Ecology, 78, 947–52.CrossRefGoogle Scholar
Marra, P. P. and Holmes, R. T. (2001). Consequences of dominance-mediated habitat segregation in American redstarts during the nonbreeding season. Auk, 118, 92–104.CrossRefGoogle Scholar
Marra, P. P., Hobson, K. A. and Holmes, R. T. (1998). Linking winter and summer events in a migratory bird by using stable carbon isotopes. Science, 282, 1884–6.CrossRefGoogle Scholar
Marra, P. P., Sherry, T. W. and Holmes, R. T. (1993). Territorial exclusion by a long-distance migrant warbler in Jamaica: a removal experiment with American redstarts, Setophaga ruticilla. Auk, 110, 565–72.CrossRefGoogle Scholar
Marsden, S. J. and Sullivan, M. S. (2000). Intersexual differences in feeding ecology in a male-dominated wintering pochard Aythya ferina population. Ardea, 88, 1–7.Google Scholar
Marsh, C. (1981). Time budget of Tana River red colobus. Folia Primatologica, 35, 30–50.CrossRefGoogle ScholarPubMed
Martin, A. R. and da Silva, V. M. F. (2004). River dolphins and flooded forest: seasonal habitat use and sexual segregation of botos (Inia geoffrensis) in an extreme cetacean environment. Journal of Zoology, London, 263, 295–305.CrossRefGoogle Scholar
Martin, J. and Lopez, P. (1999). Nuptial coloration and mate guarding affect escape decisions of male lizards (Psammodromus algirus). Ethology, 105, 439–47.CrossRefGoogle Scholar
Martin, R. and Handasyde, K. (1990). Population dynamics of the koala (Phascolarctos cinereus) in southeastern Australia. In Biology of the Koala, eds. Lee, A. K., Handasyde, K. A. and Sanson, G. D.. Chipping Norton, NSW: Surrey Beatty and Sons, pp. 75–84.Google Scholar
Martin, R. D. (1972). A preliminary field study of the lesser mouse lemur (Microcebus murinus J. F. Miller 1777). Zietschrift für Tierpsycholgie, 9, 43–89.Google Scholar
Maruhashi, T. (1981). Activity patterns of a troop of Japanese macaques (Macaca fuscata yakuii) on Yakushima Island, Japan. Primates, 22, 1–14.CrossRefGoogle Scholar
Matthysen, E., Grubb, T. C. and Cimprich, D. (1991). Social control of sex-specific foraging behaviour in downy woodpeckers, Picoides pubescens. Animal Behaviour, 42, 515–17.CrossRefGoogle Scholar
Mautz, W. W. (1978). Sledding on a brushy hillside: the fat cycle in deer. Wildlife Society Bulletin, 6, 88–90.Google Scholar
Smith, Maynard J. and Brown, R. L. W. (1986). Competition and body size. Theoretical Population Biology, 30, 166–79.CrossRefGoogle ScholarPubMed
McBride, A. F. and Kritzler, H. (1951). Observations on pregnancy, parturition, and post-natal behavior in the bottlenose dolphin. Journal of Mammalogy, 32, 251–66.CrossRefGoogle Scholar
McCann, T. S. (1980). Territoriality and breeding behaviour of adult male Antarctic Fur seal, Arctocephalus gazella. Journal of Zoology London, 192, 295–310.CrossRefGoogle Scholar
McCracken, G. F. (1984). Communal nursing in Mexican free-tailed bat maternity colonies. Science, 223, 1090–1.CrossRefGoogle ScholarPubMed
McCracken, G. F. and Bradbury, J. W. (1981). Social organisation and kinship in the polygynous bat Phyllostomus hastatus. Behavioural Ecology and Sociobiology, 8, 11–34.CrossRefGoogle Scholar
McCracken, G. F. and Wilkinson, G. S. (2000). Bat mating systems. In Reproductive Biology of Bats, eds. Crichton, E. G. and Krutzsch, P. H., San Diego: Academic Press, pp. 321–62.Google Scholar
McCullough, D. R. (1979). The George Reserve Deer Herd. Ann Arbor, University of Michigan Press.Google Scholar
McCullough, D. R. and McCullough, Y. (2000). Kangaroos in Outback Australia. New York: Columbia University Press.Google Scholar
McCullough, D. R., Hirth, D. R. and Newhouse, S. J. (1989). Resource partitioning between sexes in white-tailed deer. Journal of Wildlife Management, 53, 277–83.CrossRefGoogle Scholar
McFarland-Symington, M. (1988). Demography, activity budgets, and ranging patterns of black spider monkeys (Ateles belzebuth chamek) in the Manu National Park. American Journal of Primatology, 15, 45–67.CrossRefGoogle Scholar
McFarland-Symington, M. (1990). Fission-fusion social organization in Ateles and Pan. International Journal of Primatology, 11, 47–61.CrossRefGoogle Scholar
McIlhenny, E. A. (1937). Life history of the boat-tailed grackle in Louisiana. Auk, 54, 274–95.CrossRefGoogle Scholar
McKibben, J. N. and Nelson, D. R. (1986). Patterns of movement and grouping of gray reef sharks, Carcharhinus amblyrhynchos, at Enewetak, Marshall Islands. Bulletin of Marine Science, 38, 89–110.Google Scholar
McKinnon, J. (1994). Feeding habits of the dusky dolphin, Lagenorhynchus obscurus, in the coastal waters of central Peru. Fishery Bulletin, 92, 569–78.Google Scholar
McLaughlin, R. H. and O'Gower, A. K. (1971). Life history and underwater studies of a heterodont shark. Ecological Monographs, 41, 271–89.CrossRefGoogle Scholar
McLoyd, V. (1980). Verbally expressed modes of transformation in the fantasy and play of black preschool children. Child Development, 51, 1133–9.CrossRefGoogle Scholar
McNaughton, S. J. and Georgiadis, N. J. (1986) Ecology of African grazing and browsing mammals. Annual Review of Ecology and Systematics, 17, 39–65.CrossRefGoogle Scholar
McRobert, S. P. and Bradner, J. (1998). The influence of body coloration on shoaling preferences in fish. Animal Behaviour, 56, 611–15.CrossRefGoogle Scholar
Meaney, M. J. (1988). The sexual differentiation of social play. Trends in Neurosciences, 11, 54–8.CrossRefGoogle ScholarPubMed
Meaney, M. J., Stewart, J. and Beatty, W. W. (1985). Sex differences in social play: the socialization of sex roles. Advances in the Study of Behaviour, 15, 1–58.CrossRefGoogle Scholar
Menchen, F. C. and Winfield, I. (2000). Job search and sex segregation: Does sex of social contact matter?Sex Roles, 42(9–10), 847–64.CrossRefGoogle Scholar
Merchant, M. E., Shukla, S. S. and Akers, H. A. (1991). Lead concentrations in wing bones of the mottled duck. Environmental Toxicology and Chemistry, 10, 1503–7.CrossRefGoogle Scholar
Mesnick, S. L. (2001). Genetic relateness in sperm whales: evidence and cultural implications. Behavior and Brain Science, 26, 346–7.CrossRefGoogle Scholar
Metten, H. (1939a). Reproduction of the dogfish. Nature, 143, 121–2.CrossRefGoogle Scholar
Metten, H. (1939b). Studies on the reproduction of the dogfish. Philosophical Transactions of the Royal Society of London B, 230, 217–38.CrossRefGoogle Scholar
Michaud, R. (1993). Distribution estivale du béluga du Saint-Laurent; synthèse 1986 à 1992. Rapport technique canadien des sciences halieutiques et aquatiques, 1906, vi–28.Google Scholar
Michaud, R. (1999). Social organization of the St. Lawrence beluga, Delphinapterus leucas. 13th Conference on the Biology of Marine Mammals, Maui, Hawaii, Society for marine mammalogy.Google Scholar
Michelena, P., Bouquet, P. M., Dissac, A.et al. (2004). An experimental test of hypotheses explaining social segregation in dimorphic ungulates. Animal Behaviour, 68, 1371–80.CrossRefGoogle Scholar
Miles, D. B., Snell, H. L. and Snell, H. M. (2001). Intrapopulation variation in endurance of Galapagos lava lizards (Microlophus albemarlensis): evidence for an interaction between natural and sexual selection. Evolutionary Ecology Research, 3, 795–804.Google Scholar
Milinski, M. (1993). Predation risk and feeding behaviour. In Behaviour of Teleost Fishes, ed. Pitcher, T. J.. London: Chapman and Hall, pp. 285–305.CrossRefGoogle Scholar
Millar, J. S. and Hickling, G. J. (1992). The fasting endurance hypothesis revisited. Functional Ecology, 6, 496–8.Google Scholar
Minchin, D. (1987). Fishes of the Lough Hyne marine reserve. Journal of Fish Biology, 31, 343–52.CrossRefGoogle Scholar
Miquelle, D. G., Peek, J. M. and Ballenberghe, V. (1992). Sexual segregation in Alaskan moose. Wildlife Monographs, 122, 1–57.Google Scholar
Miranda, J. P. and Andrade, G. V. (2003). Seasonality in diet, perch use, and reproduction of the gecko Gonatodes humeralis from Eastern Brazilian Amazon. Journal of Herpetology, 37, 433–8.CrossRefGoogle Scholar
Mitani, J. C. and Amsler, S. (2003). Social and spatial aspects of male subgrouping in a community of wild chimpanzees. Behaviour, 140, 869–84.CrossRefGoogle Scholar
Mitani, J. C. and Watts, D. P. (2002). Why do male chimpanzees hunt and share meat?Animal Behaviour, 61, 915–24.CrossRefGoogle Scholar
Mitani, J. C., Gros-Louis, J. and Richards, A. F. (1996). Sexual dimorphism, the operational sex ratio, and the intensity of male competition in poygynous primates. American Naturalist, 147, 966–80.CrossRefGoogle Scholar
Mitani, J. C., Watts, D. P. and Lwanga, J. (2002). Ecological and social correlates of chimpanzee party size and composition. In Behavioural Diversity in Chimpanzees and Bonobos, eds., Boesch, C., Hohmann, G. and Marchant, L. F.. Cambridge: Cambridge University Press, pp. 102–11.CrossRefGoogle Scholar
Mitchell, C. (1994). Migration alliances and coalitions among adult male South American squirrel monkeys. Behaviour, 130, 169–89.CrossRefGoogle Scholar
Mitchell, P. (1990). The home ranges and social activity of koalas – a quantitative analysis. In Biology of the Koala, eds. Lee, A. K., Handasyde, K. A. and Sanson, G. D.. Chipping Norton, NSW: Surrey Beatty and Sons, pp. 171–87.Google Scholar
Miyazaki, N. and Nishiwaki, M. (1978). School structure of the striped dolphins off the Pacific coast of Japan. Scientific Report of the Whale Research Institute, 30, 65–115.Google Scholar
Moll, D. (1990). Population sizes and foraging ecology in a tropical freshwater stream turtle community. Journal of Herpetology, 24, 48–53.CrossRefGoogle Scholar
Moll, E. O. and Legler, J. M. (1971). The life history of a neotropical slider turtle, Pseudemys scripta (Schoepff) in Panama. Bulletin of the Los Angeles County Museum of Natural History, Science, 11, 1–102.Google Scholar
Moore, C. L. (1985). Development of mammalian sexual behavior. In The Comparative Development of Adaptative Skills: Evolutionary Implications, ed. Gollin, E. S.. Hillsdale: Lawrence Erlbaum Associate, pp. 19–56.Google Scholar
Morgantini, L. E. and Hudson, R. J. (1981). Sex differential in use of the physical environment by bighorn sheep (Ovis canadensis). Canadian Field-Naturalist, 95, 69–74.Google Scholar
Mori, A. and Watanabe, K. (2003). Life history of male Japanese monkeys living on Koshima Islet. Primates, 44, 119–26.Google ScholarPubMed
Moritz, C. (1993). The origin and evolution of parthenogenesis in the Heteronotia binoei complex: synthesis. Genetica, 90, 269–80.CrossRefGoogle Scholar
Morreale, S. J., Gibbons, J. W. and Congdon, J. D. (1984). Significance of activity and movement in the yellow-bellied slider turtle (Pseudemys scripta). Canadian Journal of Zoology, 62, 1038–2.CrossRefGoogle Scholar
Morrison, M. L. and With, K. A. (1987). Interseasonal and intersexual resource partitioning in hairy and white-headed woodpeckers. Auk, 104, 225–33.Google Scholar
Morrissey, J. F. and Gruber, S. H. (1993). Habitat selection by juvenile lemon sharks, Negaprion brevirostris. Environmental Biology of Fishes, 38, 311–19.CrossRefGoogle Scholar
Morton, E. S. (1990). Habitat segregation by sex in the hooded warbler: experiments on proximate causation and discussion of its evolution. American Naturalist, 135, 319–33.CrossRefGoogle Scholar
Morton, E. S., Voort, M. E. and Greenberg, R. (1993). How a warbler chooses its habitat: field support for laboratory experiments. Animal Behaviour, 46, 47–53.CrossRefGoogle Scholar
Morton, S. (1978). An ecological study of Sminthopsis crassicaudata (Marsupialia: Dasyuridae) III. Reproduction and life history. Australian Wildlife Research, 5, 183–211.CrossRefGoogle Scholar
Morton, E. S., Lynch, J. F., Young, K. and Mehlhop, P. (1987). Do male hooded warblers exclude females from non-breeding territories in tropical forest?Auk, 104, 133–5.CrossRefGoogle Scholar
Moseby, K. E. and O'Donnell, E. (2003). Reintroduction of the greater bilby, Macrotis lagotis (Reid) (Marsupialia: Thylacomyidae), to northern South Australia: survival, ecology and notes on reintroduction protocols. Wildlife Research, 30, 15–27.CrossRefGoogle Scholar
Moskowitz, D. S., Suh, E. J. and Desaulniers, J. (1994). Situational influences on gender differences in agency and communion. Journal of Personality and Social Psychology, 66(4), 753–61.CrossRefGoogle ScholarPubMed
Moss, R. (1983). Gut size, body weight, and digestion of winter foods by grouse and ptarmigan. Condor, 85, 185–93.CrossRefGoogle Scholar
Mougin, J. L. (1968). Etude écologique de quatre espèces de pétrels antarctique. Oiseau Revue Francophone Ornithology, 38, 1–52.Google Scholar
Mougin, J. L. (1975). Ecologie comparée des Procellariidae Antarctiques et sub-Antarctiques. CNFRA, 36, 1–195.Google Scholar
Mueller, H. C. (1990). The evolution of reversed sexual dimorphism in size in monogamous species of birds. Biological Reviews of the Cambridge Philosophical Society, 65, 553–85.CrossRefGoogle Scholar
Müller, A. E. and Thalmann, U. (2000). Origin and evolution of primate social organisation: a reconstruction. Biological Reviews of the Cambridge Philosophical Society, 75, 405–35.CrossRefGoogle ScholarPubMed
Murphy, M. T., Pierce, A., Shoen, J.et al. (2001). Population structure and habitat use by overwintering neotropical migrants on a remote oceanic island. Biological Conservation, 102, 333–45.CrossRefGoogle Scholar
Murray, M. G. and Illius, A. W. (2000). Vegetation modification and resource competition in grazing ungulates. Oikos, 89, 501–8.CrossRefGoogle Scholar
Murray, T. E., Bartle, J. A., Kalish, S. R. and Taylor, P. R. (1993). Incidental capture of seabirds by Japanese southern bluefin tuna longline vessels in New Zealand waters, 1988–1992. Bird Conservation International, 3, 181–210.CrossRefGoogle Scholar
Mushinsky, H. R., Hebrard, J. J. and Vodopich, D. S. (1982). Ontogeny of water snake foraging ecology. Ecology, 63, 1624–9.CrossRefGoogle Scholar
Myers, J. P. (1981). A test of three hypotheses for latitudinal segregation of the sexes in wintering birds. Canadian Journal of Zoology, 59, 1527–34.CrossRefGoogle Scholar
Myrberg, A. A. and Gruber, S. H. (1974). The behaviour of the bonnethead, Sphyrna tiburo. Copeia, 1974, 358–74.CrossRefGoogle Scholar
Myres, B. C. and Eells, M. M. (1968). Thermal aggregation in Boa constrictor. Herpetologica, 24, 61–6.Google Scholar
Mysterud, A. (2000). The relationship between ecological segregation and sexual body size dimorphism in large herbivores. Oecologia, 124, 40–54.CrossRefGoogle ScholarPubMed
Nakagawa, N. (1989). Activity budget and diet of patas monkeys in Kala Maloue National Park, Cameroon. Primates, 30, 27–34.CrossRefGoogle Scholar
Naulleau, G. (1966). Etude complementaire de l'activitie de Vipera aspis dans la nature. Vie et Milieu, 17, 461–509.Google Scholar
Nel, D. C., Ryan, P. G., Nel, J. L.et al. (2002). Foraging interactions between wandering albatrosses Diomedea exulans breeding on Marion Island and long-line fisheries in the southern Indian Ocean. Ibis, 144 (on-line), E141–E154.CrossRefGoogle Scholar
Nelson, J. E. (1965). Behaviour of Australian Pteropodidae (Megachiroptera). Animal Behaviour, 13, 544–57.CrossRefGoogle Scholar
Nelson, M. E. and Mech, L. D. (1981). Deer social organization and wolf predation in northeastern Minnesota. Wildlife Monographs, 77, 1–53.Google Scholar
Nelson, M. E. and Mech, L. D. (1984). Home-range formation and dispersal of deer in Northeastern Minnesota. Journal of Mammalogy, 65, 567–75.CrossRefGoogle Scholar
Neuhaus, P. and Ruckstuhl, K. E. (2002a). The link between sexual dimorphism, activity budgets, and group cohesion: the case of the plains zebra (Equus burchelli). Canadian Journal of Zoology, 80, 1437–41.CrossRefGoogle Scholar
Neuhaus, P. and Ruckstuhl, K. E. (2002b). Foraging behaviour in Alpine ibex (Capra ibex): consequences of reproductive status, body size, age and sex. Ecology, Ethology and Evolution, 14, 373–81.CrossRefGoogle Scholar
Newsome, A. E. (1980). Differences in the diets of male and female red kangaroos in central Australia. African Journal of Ecology, 18, 27–31.CrossRefGoogle Scholar
Newton, I. (1986). The Sparrowhawk. Calton: T and AD Poyser.Google Scholar
Newton, I. (1998). Population Limitation in Birds. San Diego: Academic Press.Google Scholar
Newton-Fisher, N., Reynolds, V. and Plumtre, A. J. (2000). Food supply and chimpanzee (Pan troglodytes schweinfurthii) in the Budongo Forest Reserve, Uganda. International Journal of Primatology, 21, 613–28.CrossRefGoogle Scholar
Nicholls, D. G., Murray, M. D., Butcher, E. and Moors, P. (1997). Weather systems determine the non-breeding distribution of wandering albatrosses over southern oceans. Emu, 95, 240–4.Google Scholar
Nichols, J. D. and Aramis, G. M. (1980). Sex-specific differences in winter distribution patterns of canvasbacks. Condor, 82, 406–16.CrossRefGoogle Scholar
Nievergelt, B. (1967). Die zusammensetzung der gruppen beim alpensteinbock. Zeitschrift für Säugetierkunde, 32, 129–44.Google Scholar
Nievergelt, B. (1981). Ibexes in an African Environment. Berlin: Springer Verlag.CrossRefGoogle Scholar
Nikaido, M., Rooney, A. P. and Okada, N. (1999). Phylogenetic relationships among cetartiodactyls based on insertions of short and long interspersed elements: hippopotamuses are the closest extant relatives of whales. Proceedings of the National Academy of Science, 96, 10261–6.CrossRefGoogle Scholar
Nilsson, L. (1970). Food seeking activity of south Swedish diving ducks in the non-breeding season. Oikos, 21, 145–54.CrossRefGoogle Scholar
Nisbet, I. C. T. and Medway, L. (1972). Dispersion, population ecology and migration of eastern great reed warblers Acrocephalus orientalis wintering in Malaysia. Ibis, 114, 451–94.CrossRefGoogle Scholar
Nisbet, I. C. T., Montoya, J. P., Burger, J. and Hatch, J. J. (2002). Use of stable isotopes to investigate individual differences in diets and mercury exposures among common terns Sterna hirundo in breeding and wintering grounds. Marine Ecology Progress Series, 242, 267–74.CrossRefGoogle Scholar
Nishida, T. (1968). The social group of wild chimpanzees in the Mahale Mountains. Primates, 9, 167–224.CrossRefGoogle Scholar
Nishida, T. and Hosaka, K. (1996). Coalition strategies among adult male chimpanzees of the Mahale Mountains, Tanzania. In Great Ape Societies, eds. McGrew, W. A., Marchant, L. A., and Nishida, T.. Cambridge: Cambridge University Press, pp. 114–34.CrossRefGoogle Scholar
Nogueira, C., Sawaya, R. J. and Martins, M. (2003). Ecology of the pitviper, Bothrops moojeni, in the Brazilian Cerrado. Journal of Herpetology, 37, 653–9.CrossRefGoogle Scholar
Norbury, G. L., Coulson, G. M. and Walters, B. L. (1988). Aspects of the demography of the western grey kangaroo, Macropus fuliginosus melanops, in semiarid north-west Victoria. Australian Wildlife Research, 15, 257–66.CrossRefGoogle Scholar
Norris, K. S. (1994). Comparative view of cetacean social ecology, culture, and evolution. In The Hawaiian Spinner Dolphin, eds. Norris, K. S., Würsig, B., Well, R. S. and Würsig, M.. Berkeley: University of California Press, pp. 301–44.Google Scholar
Norris, K. S. and Dohl, T. P. (1980). The structure and functions of cetacean schools. In Cetacean Behavior: Mechanisms and Functions, ed. Herman, L. M.. New York: John Wiley and Sons, pp. 211–59.Google Scholar
Norris, K. S. and Schilt, C. R. (1988). Cooperative societies in three-dimensional space: on the origins of aggregations, flocks, and schools, with special reference to dolphins and fish. Ethology and Sociobiology, 9, 149–79.CrossRefGoogle Scholar
Northcutt, R. G. (1977). Elasmobranch central nervous system organization and its possible evolutionary significance. American Zoologist, 17, 411–29.CrossRefGoogle Scholar
Novacek, M. J. (1992). Mammalian phylogeny: shaking the tree. Nature, London, 356, 121–5.CrossRefGoogle ScholarPubMed
Nowak, R. M. (1991). Walker's Mammals of the World, 5th edn. Baltimore: John Hopkins University Press.Google Scholar
Nunn, C. (1999). The number of males in primate social groups: a comparative test of the socioecological model. Behavioral Ecology and Sociobiology, 46, 1–13.CrossRefGoogle Scholar
O'Donnell, C. F. J. and Sedgeley, J. A. (1999). Use of roosts by the long-tailed bat, Chalinolobus tuberculatus, in temperate rain forest in New Zealand. Journal of Mammalogy, 80, 913–23.CrossRefGoogle Scholar
Oakwood, M. (2000). Reproduction and demography of the northern quoll, Dasyurus hallucatus, in the lowland savanna of northern Australia. Australian Journal of Zoology, 48, 519–39.CrossRefGoogle Scholar
Oakwood, M. (2002). Spatial and social organization of a carnivorous marsupial Dasyurus hallucatus (Marsupialia: Dasyuridae). Journal of Zoology, 257, 237–48.CrossRefGoogle Scholar
Ohguchi, O. (1978). Experiments on the selection against colour oddity of water fleas by three-spined stickelbacks. Zeitschrift fur Tierpsychologie, 47, 254–67.CrossRefGoogle Scholar
Olesiuk, P. F., Bigg, M. A. and Ellis, G. M. (1990). Life history and population dynamics of resident killer whales (Orcinus orca) in the coastal waters of British Columbia and Washington state. Report of the International Whaling Commission, Special Issue, 12, 209–405.
Oli, M. K. (1996). Seasonal patterns in habitat use of blue sheep Pseudois nayaur (Artiodactyla, Bovidae) in Nepal. Mammalia, 60, 187–93.CrossRefGoogle Scholar
Oliver, W. R. B. (1955). New Zealand Birds. Wellington: A. H. and A. W. Reed.Google Scholar
Olsen, A. M. (1954). The biology, migration, and growth rate of the school shark, Galeorhinus australis (Macleay) (Carcharhinidae) in southeastern Australian waters. Australian Journal of Marine and Freshwater Research, 5, 353–410.Google Scholar
Ong, P. S. (1994). The Social Organization of the Common Ringtail Possum, Pseudocheirus peregrinus. Unpublished Ph.D. thesis, Monash University.Google Scholar
Ordway, L. L. and Krausman, P. R. (1986). Habitat use by mule deer. Journal of Wildlife Management, 50, 677–83.CrossRefGoogle Scholar
Ortega, J. and Arita, H. T. (2000). Defence of females by dominant males of Artibeus jamaicensis (Chiroptera: Phyllostomidae). Ethology, 106, 395–407.CrossRefGoogle Scholar
Ortega, J. and Arita, H. T. (2002). Subordinate males in harem groups of Jamaican fruit-eating bats (Artibeus jamaicensis): Satellites or sneaks?Ethology, 108, 1077–91.CrossRefGoogle Scholar
Oswald, C., Fonken, P., Atkinson, D. and Palladino, M. (1993). Lactational water-balance and recycling in white-footed mice, red-backed voles, and gerbils. Journal of Mammalogy, 74, 963–70.CrossRefGoogle Scholar
Owen, M. and Black, J. M. (1990). Waterfowl Ecology. Glasgow and London: Blackie.Google Scholar
Owen, M. and Black, J. M. (1991). The importance of migration mortality in non-passerine birds. In Bird Population Studies. Relevance to Conservation and Management, eds. Perrins, C. M., Lebreton, J.-D. and Hirons, G. J. M.. Oxford University Press.Google Scholar
Owen, M. and Dix, M. (1986). Sex ratios in some common British wintering ducks. Wildfowl, 37, 104–12.Google Scholar
Owens, I. P. F. and Bennet, P. M. (1994). Mortality costs of parental care and sexual dimorphism in birds. Proceedings of the Royal Society of London B, 257, 1–8.CrossRefGoogle Scholar
Owen-Smith, N. (1993). Comparative mortality rates of male and female kudus: the costs of sexual size dimorphism. Journal of Animal Ecology, 62, 428–40.CrossRefGoogle Scholar
Owen-Smith, R. N. (1988). Megaherbivores. The Influence of Very Large Body Size on Ecology. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Ozoga, J. J. (1968). Variations in microclimate in a conifer swamp deeryard in Northern Michigan. Journal of Wildlife Management, 32, 574–81.CrossRefGoogle Scholar
Ozoga, J. J. and , Verme L. J. (1970). Winter feeding patterns of penned white-tailed deer. Journal of Wildlife Management, 34, 431–9.CrossRefGoogle Scholar
Palombit, R. A. (1999). Infanticide and the evolution of pair bonds in nonhuman primates. Evolutionary Anthropology, 7, 117–28.3.0.CO;2-O>CrossRefGoogle Scholar
Palombit, R. A., Cheney, D. L., Fischer, J. et al. (2000). Male infanticide and defense of infants in chacma baboons. In Infanticide by Males and Its Implications, eds. Schaik, C. P. and Janson, C. H.. Cambridge: Cambridge University Press, pp. 123–52.
Papastavrou, V., Smith, S. C. and Whitehead, H. (1989). Diving behaviour of the sperm whale, Physeter macrocephalus, off the Galapagos Islands. Canadian Journal of Zoology, 67, 839–46.CrossRefGoogle Scholar
Parke, R. D. and Suomi, S. J. (1981). Adult male infant relationships: human and nonhuman primate evidence. In Behavioral Development, eds. Immelman, K., Barlow, G. W., Petronovitch, L., and Main, M.. New York: Cambridge University Press, pp. 700–25.Google Scholar
Parker, K. L., Gillingham, M. P., Hanley, T. A. and Robbins, C. T. (1999). Energy and protein balance of free-ranging black-tailed deer in a natural forest environment. Wildlife Monographs, 143, 1–48.Google Scholar
Parker, W. S. and Brown, W. S. (1980). Comparative ecology of two colubrid snakes, Masticophis t. taeniatus and Pituophis melanoleucus deserticola, in northern Utah. Milwaukee Public Museum Publications in Biology and Geology, 7, 1–104.Google Scholar
Parmelee, D. F., Parmelee, J. M. and Fuller, M. R. (1985). Ornithological investigations at Palmer Station: the first long-distance tracking of seabirds by satellites. Antarctic Journal of the United States, 162–3.Google Scholar
Parmelee, J. R. and Guyer, C. (1995). Sexual differences in foraging behavior of an anoline lizard, Norops humilis. Journal of Herpetology, 29, 619–21.CrossRefGoogle Scholar
Parr, A. E. (1931). Sex dimorphism and schooling behavior among fishes. American Naturalist, 65, 173–80.CrossRefGoogle Scholar
Parrish, J. D. and Sherry, T. W. (1994). Sexual habitat segregation by American redstarts wintering in Jamaica: importance of resource seasonality. Auk, 111, 38–49.CrossRefGoogle Scholar
Parsons, K. N., Jones, G., Davidson-Watts, I. and Greenaway, F. (2003). Swarming of bats at underground sites in Britain – implications for conservation. Biological Conservation, 111, 63–70.CrossRefGoogle Scholar
Pasinelli, G. (2000). Sexual dimorphism and foraging niche partitioning in the middle spotted woodpecker Dendrocopos medius. Ibis, 2000, 635–44.Google Scholar
Patterson, D. L. and Fraser, W. R. (2003). Satellite tracking southern giant petrels at Palmer station, Antarctica. Feature Articles, Microwave Telemetry, Inc., 8, 3–4.Google Scholar
Patterson, I. A. P., Reid, R. J., Wilson, B.et al. (1998). Evidence for infanticide in bottlenose dolphins: an explanation for violent interactions with harbour porpoises?Proceedings of the Royal Society of London, Series B, 265, 1167–70.CrossRefGoogle ScholarPubMed
Pearson, D., Shine, R. and How, R. (2002a). Sex-specific niche partitioning and sexual size dimorphism in Australian pythons (Morelia spilota imbricata). Biological Journal of the Linnean Society, 77, 113–25.CrossRefGoogle Scholar
Pearson, D., Shine, R. and Williams, A. (2002b). Geographic variation in sexual size dimorphism within a single snake species (Morelia spilota, Pythonidae). Oecologia, 131, 418–26.CrossRefGoogle Scholar
Pearson, D., Shine, R. and Williams, A. (2003). Thermal biology of large snakes in cool climates: a radio-telemetric study of carpet pythons (Morelia spilota imbricata) in south-western Australia. Journal of Thermal Biology, 28, 117–31.CrossRefGoogle Scholar
Pellegrini, A. D. (2003). Perceptions and possible functions of play and real fighting in early adolescence. Child Development, 74, 1459–70.CrossRefGoogle Scholar
Pellegrini, A. D. and Bartini, M. (2001). Dominance in early adolescent boys: affiliative and aggressive dimensions and possible functions. Merrill-Palmer Quarterly, 47, 142–63.CrossRefGoogle Scholar
Pellegrini, A. D. and Long, J. D. (2003). A sexual selection theory longitudinal analysis of sexual segregation and integration in early adolescence. Journal of Experimental Child Psychology, 85, 257–78.CrossRefGoogle ScholarPubMed
Pellegrini, A. D. and Perlmutter, J. C. (1989). Classroom contextual effects on children's play. Developmental Psychology, 25, 289–96.CrossRefGoogle Scholar
Pellegrini, A. D. and Smith, P. K. (1998). Physical activity play: the nature and function of a neglected aspect of play. Child Development, 69, 577–98.CrossRefGoogle Scholar
Pellegrini, A. D., Blatchford, P., Kato, K. and Baines, E. (2003). A short-time longitudinal study of children's playground games in primary school: implications for adjustment to school and social adjustment in the USA and the UK. Social Development, 13, 107–23.CrossRefGoogle Scholar
Pellegrini, A. D., Horvat, M. and Huberty, P. D. (1998). The relative cost of children's physical activity play. Animal Behaviour, 55, 1053–106.CrossRefGoogle ScholarPubMed
Pellew, R. A. (1983). The impacts of elephant, giraffe and fire upon the Acacia tortilis woodlands of the Serengeti. African Journal of Ecology, 21, 41–74.CrossRefGoogle Scholar
Pennycuick, C. J. (1987). Flight of seabirds. In Seabirds: Feeding Ecology and Role in Marine Ecosystems, ed. Croxall, J. P.. Cambridge: Cambridge University Press, pp. 43–62.Google Scholar
Pepper, J. M., Mitani, J. C. and Watts, D. P. (1999). General gregariousness and specific partner preference among wild chimpanzees. International Journal of Primatology, 20, 613–32.CrossRefGoogle Scholar
Pereira, M. E. (1988). Effects of age and sex on intra-group spacing in juvenile savanna baboons, Papio cynocephalus cynocephalus. Animal Behaviour, 36, 184–204.CrossRefGoogle Scholar
Pereira, M. E. (1989). Agonistic interactions of juvenile savanna baboons. II. Agonistic support and rank acquisition. Ethology, 80, 152–71.CrossRefGoogle Scholar
Pereira, M. E. and Fairbanks, L. A. (2002). Foreword 2002: family, friends, and the evolution of childood. In Juvenile Primates, eds. Pereira, M. E. and Fairbanks, L. A., 2nd edn. Chicago: University of Chicago Press, pp. vii–xxiv.Google Scholar
Pereira, M. E. and Leigh, S. R. (2003). Modes of primate development. In Primate Life Histories and Socioecology, eds. Kappeler, P. M. and Pereira, M. E.. Chicago: University of Chicago Press, pp. 149–76.Google Scholar
Pérez-Barbería, F. J. and Gordon, I. J. (1998a). The influence of sexual dimorphism in body size and mouth morphology on diet selection and sexual segregation in cervids. Acta Veterinaria Hungarica, 46, 357–67.Google Scholar
Pérez-Barbería, F. J. and Gordon, I. J. (1998b). Factors affecting food comminution during chewing in ruminants: a review. Biological Journal of the Linnean Society, 63, 233–56.CrossRefGoogle Scholar
Pérez-Barbería, F. J. and Gordon, I. J. (1999). Body size dimorphism and sexual segregation in polygynous ungulates: an experimental test with Soay sheep. Oecologia, 120, 258–67.Google ScholarPubMed
Pérez-Barbería, F. J., Gordon, I. J. and Pagel, M. (2002). The origins of sexual dimorphism in body size in ungulates. Evolution, 56, 1276–85.CrossRefGoogle ScholarPubMed
Perez-Mellado, V. and Riva, I. (1993). Sexual size dimorphism and ecology: the case of a tropical lizard, Tropidurus melanopleurus (Sauria, Tropiduridae). Copeia, 1993, 969–76.CrossRefGoogle Scholar
Pernetta, J. C. (1977). Observations on the habits and morphology of the snake Laticauda colubrina in Fiji. Canadian Journal of Zoology, 55, 1612–19.CrossRefGoogle Scholar
Perret, M. (1998). Energetic advantage of nest sharing in a solitary primate, the lesser mouse lemur (Microcebus murinus). Folia Primatologica, 59, 1–25.Google Scholar
Perrin, W. F. and Reilly, S. B. (1984). Reproductive parameters of dolphins and small whales of the family Delphinidae. Report of the International Whaling Commission, Special Issue, 6, 97–134.Google Scholar
Perrin, W. F., Wilson, C. E. and Archer II, F. I. (1994). Striped dolphin – Stenella coeruleoalba (Meyen, 1833). In Handbook of Marine Mammals: The First Book of Dolphins, vol. 5, eds. Ridgway, S. H. and Harrison, R.. London: Academic Press, pp. 129–60.Google Scholar
Perry, G. (1996). The evolution of sexual dimorphism in the lizard Anolis polylepis (Iguania): evidence from intraspecific variation in foraging behavior and diet. Canadian Journal of Zoology, 74, 1238–45.CrossRefGoogle Scholar
Perry, G. and Brandeis, M. (1992). Variation in stomach contents of the gecko Ptyodactylus hasselquistii guttatus in relation to sex, season, and locality. Amphibia-Reptilia, 13, 275–82.CrossRefGoogle Scholar
Perry, G. and Garland, T. (2002). Lizard home ranges revisited: effects of sex, body size, diet, habitat, and phylogeny. Ecology, 83, 1870–85.CrossRefGoogle Scholar
Peter, H.-U., Kaiser, M. and Gebauer, A. (1988). Investigations on birds and seals at King George Island. Geodätische und geophysikalische Veröffentlichungen, 1, 5–80.Google Scholar
Peters, W. D. and Grubb, T. C. (1983). An experimental analysis of sex-specific foraging in the downy woodpecker Picoides pubescens. Ecology, 64, 1437–43.CrossRefGoogle Scholar
Peterson, C. R., Gibson, A. R. and Dorcas, M. E. (1993). Snake thermal ecology: the causes and consequences of body-temperature variation. In Snakes: Ecology and Behavior, eds. Seigel, R. A. and Collins, J. T.. New York: McGraw-Hill, pp. 241–314.Google Scholar
Peterson, R. L. (1965). A review of the bats of the genus Ametrida, family Phyllostomidae. Life Science Contribution, Royal Ontario Museum, 65, 1–13.Google Scholar
Pettit, K. E., Bishop, C. A. and Brooks, R. J. (1995). Home range and movements of the common snapping turtle, Chelydra serpentina serpentina, in a coastal wetland of Hamilton Harbor, Lake Ontario, Canada. Canadian Field-Naturalist, 109, 192–200.Google Scholar
Peuhkuri, N. (1999). Size-assorted fish shoals and the majority's choice. Behavioural Ecology and Sociobiology, 46, 307–12.CrossRefGoogle Scholar
Pfeffer, P. (1967). Le mouflon de Corse (Ovis ammon musimon Schreber, 1782); position systématique, écologie, et éthologie comparées. Mammalia, 31, 1–262.Google Scholar
Phelps, T. W. (1978). Seasonal movement of the snakes Coronella austriaca, Vipera berus, and Natrix natrix in southern England. British Journal of Herpetology, 5, 775–81.Google Scholar
Phillips, J. A. (1995). Movement patterns and density of Varanus albigularis. Journal of Herpetology, 29, 407–16.CrossRefGoogle Scholar
Phillips, R. A., Silk, J. R. D., Croxall, J. P., Afanasyev, V. and Briggs, D. R. (2003). Accuracy of geolocation estimates for flying seabirds. Marine Ecology Progress Series, 266, 265–72.CrossRefGoogle Scholar
Phillips, R. A., Silk, J. R. D., Phalan, B., Catry, P. and Croxall, J. P. (2004). Seasonal sexual segregation in the two Thalassarche albatross species: competitive exclusion, reproductive role specialization or foraging niche divergence?Proceeding of the Royal Society of London, 271, 1283–91.CrossRefGoogle ScholarPubMed
Pianka, E. R. and Vitt, L. J. (2003). Lizards: Windows to the Evolution of Diversity. Berkeley, CA: University of California Press.Google Scholar
Pickering, S. P. C. (1989). Attendance patterns and behaviour in relation to experience and pair-bond formation in wandering albatrosses Diomedea exulans at South Georgia. Ibis, 131, 183–95.CrossRefGoogle Scholar
Pickering, S. P. C. and Berrow, S. D. (2002). Courtship behavior of the wandering albatross Diomedea exulans at Bird Island, South Georgia. Marine Ornithology, 29, 29–37.Google Scholar
Pierce, B. M., Bleich, V. C. and Bowyer, R. T. (2000). Prey selection by mountain lions and coyotes: effects of hunting style, body size, and reproductive status. Journal of Mammalogy, 81, 462–72.2.0.CO;2>CrossRefGoogle Scholar
Pierce, G. J. and Boyle, P. R. (1991). A review of methods for diet analysis in piscivorous marine mammals. Oceanography and Marine Biology, 29, 409–86.Google Scholar
Pierotti, R. and Annett, C. A. (1991). Diet choice in the herring gull: constraints imposed by reproductive and ecological factors. Ecology, 72, 319–28.CrossRefGoogle Scholar
Pippard, L. (1985). Status of the St. Lawrence River population of beluga, Delphinapterus leucas. Canadian Field-Naturalist, 99, 438–50.Google Scholar
Pitcher, E. G. and Schultz, L. H. (1983). Boys and Girls at Play: The Development of Sex Roles. South Hadley, MA: Bergin and Garvey.Google Scholar
Pitcher, T. J. and Parrish, J. K. (1993). Functions of shoaling behaviour in teleosts. In Behaviour of Teleost Fishes, ed. Pitcher, T. J.. Chapman and Hall, London, pp. 363–439.CrossRefGoogle Scholar
Pitcher, T. J., Magurran, A. E. and Edwards, J. I. (1985). Schooling mackerel and herring choose neighbours of similar size. Marine Biology, 86, 319–22.CrossRefGoogle Scholar
Pitman, R. L., Ballance, L. T., Mesnick, S. I. and Chivers, S. J. (2001). Killer whale predation on sperm whales: observations and implications. Marine Mammal Science, 17, 494–507CrossRefGoogle Scholar
Plummer, M. V. (1977). Activity, habitat, and population structure in the turtle Trionyx muticus. Copeia, 1977, 431–440.CrossRefGoogle Scholar
Plummer, M. V. (1981a). Communal nesting of Opheodrys aestivus in the laboratory. Copeia, 1981, 243–6.CrossRefGoogle Scholar
Plummer, M. V. (1981b). Habitat utilization, diet and movements of a temperate arboreal snake (Opheodrys aestivus). Journal of Herpetology, 15, 425–32.CrossRefGoogle Scholar
Plummer, M. V. and Farrar, D. B. (1981). Sexual dietary differences in a population of Trionyx muticus. Journal of Herpetology, 15, 175–9.CrossRefGoogle Scholar
Pluto, T. G. and Bellis, E. D. (1986). Habitat utilization by the turtle, Graptemys geographica, along a river. Journal of Herpetology, 20, 22–31.CrossRefGoogle Scholar
Pluto, T. G. and Bellis, E. D. (1988). Seasonal and annual movements of riverine map turtles, Graptemys geographica. Journal of Herpetology, 22, 152–8.CrossRefGoogle Scholar
Poindron, P., Lévy, F. and Krehbiel, D. (1988). Genital, olfactory, and endocrine interactions in the development of maternal behaviour in the parturient ewe. Psychoneuroendocrinology, 13, 99–125.CrossRefGoogle ScholarPubMed
Polis, G. A. (1984). Age structure component of niche width and intra-specific resource partitioning: can age groups function as ecological species?American Naturalist, 123, 541–64.CrossRefGoogle Scholar
Pope, T. (1990). The reproductive consequences of male cooperation in the red howler monkey: paternity exclusion in multi-male and single-male troops using genetic markers. Behavioral Ecology and Sociobiology, 27, 439–46.CrossRefGoogle Scholar
Pope, T. (2000a). The evolution of male philopatry in neotropical monkeys. In Primate Males, ed. Kappeler, P.. Cambridge: Cambridge University Press, pp. 219–35.Google Scholar
Pope, T. (2000b). Reproductive success increases with degree of kinship in cooperative coalitions of female red howler monkeys. Behavioral Ecology and Sociobiology, 48, 253–67.CrossRefGoogle Scholar
Porter, D. A. (1990). Feeding ecology of Graptemys caglei Haynes and McKown in the Guadalupe River, Dewitt County, Texas. M.Sc. thesis, Canyon, TX: West Texas State University.Google Scholar
Porter, R. H., Désiré, L., Bon, R., Orgeur, P. (2001). The role of familiarity in the development of social recognition by lambs. Behaviour, 134, 207–19.CrossRefGoogle Scholar
Porter, W. F. (1994). Family Meleagridae (Turkeys). In Handbook of the Birds of the World, vol. 2, eds. Hoyo, J. del, Elliot, A. and Sargatal, J.. Barcelona: Lynx Edicions.Google Scholar
Post, D., Hausfater, G. and McKusky, S. (1981). Feeding behavior of yellow baboons (Papio cynocephalus): relationship to age, gender, and dominance rank. Folia Primatologica, 34, 170–95.CrossRefGoogle Scholar
Post, D. M., Armbrust, T. S., Horne, E. A. and Goheen, J. R. (2001). Sexual segregation results in differences in content and quality of bison (Bos bison) diets. Journal of Mammalogy, 82, 407–13.2.0.CO;2>CrossRefGoogle Scholar
Post, E., Langvatn, R., Forchhammer, M. C. and Stenseth, N. C. (1999). Environmental variation shapes sexual dimorphism in red deer. Proceedings of the National Academy of Sciences, 96, 4467–71.CrossRefGoogle ScholarPubMed
Pough, F. H. (1973). Lizard energetics and diet. Ecology, 54, 837–44.CrossRefGoogle Scholar
Pough, F. H. (1980). The advantages of ectothermy for tetrapods. American Naturalist, 115, 92–112.CrossRefGoogle Scholar
Pough, F. H. and Groves, J. D. (1983). Specialization in the body form and food habits of snakes. American Zoologist, 23, 443–54.CrossRefGoogle Scholar
Pounds, J. A. and Jackson, J. F. (1983). Utilization of perch sites by sex and size classes of Sceloporus undulatus undulatus. Journal of Herpetology, 17, 287–9.CrossRefGoogle Scholar
Powell, G. L. and Russell, A. P. (1984). The diet of the eastern short-horned lizard (Phrynosoma douglassi brevirostre) in Alberta and its relationship to sexual size dimorphism. Canadian Journal of Zoology, 62, 428–40.CrossRefGoogle Scholar
Power, T. G. (2000). Play and Exploration in Children and Animals. Mahwah, NJ: Erlbaum.Google Scholar
Pratt, H. L. (1979). Reproduction in the blue shark, Prionace glauca. U.S. Fishery Bulletin, 77, 445–70.Google Scholar
Pratt, H. L. and Carrier, J. C. (2001). A review of elasmobranch reproductive behaviour with a case study on the nurse shark, Ginglymostoma cirratum. Environmental Biology of Fishes, 60, 157–88.CrossRefGoogle Scholar
Preest, M. R. (1994). Sexual size dimorphism and feeding energetics in Anolis carolinensis: why do females take smaller prey than males?Journal of Herpetology, 28, 292–8.CrossRefGoogle Scholar
Preston, C. R. (1990). Distribution of raptor foraging in relation to prey biomass and habitat structure. Condor, 92, 107–12.CrossRefGoogle Scholar
Prestt, I. (1971). An ecological study of the viper Vipera berus in southern Britain. Journal of Zoology, London, 164, 373–418.CrossRefGoogle Scholar
Prince, P. A. and Francis, M. D. (1984). Activity budgets of foraging grey-headed albatrosses. Condor, 86, 297–300.CrossRefGoogle Scholar
Prince, P. A. and Morgan, R. A. (1987). Diet and feeding ecology of Procellariiformes. In Seabirds: Feeding Ecology and Role in Marine Ecosystems, ed. Croxall, J. P.. Cambridge: Cambridge University Press, pp. 135–71.Google Scholar
Prince, P. A. and Walton, D. W. H. (1984). Automated measurement of feed size and feeding frequency in albatrosses. Journal of Applied Ecology, 21, 789–94.CrossRefGoogle Scholar
Prince, P. A., Croxall, J. P., Trathan, P. N. and Wood, A. G. (1998). The pelagic distribution of South Georgia albatrosses and their relationships with fisheries. In Albatross Biology and Conservation, eds. Robertson, G. and Gales, R.. Chipping Norton: Surrey Beatty and Sons, pp. 69–83.Google Scholar
Prince, P. A., Weimerskirch, H., Wood, A. G. and Croxall, J. P. (1999). Areas and scales of interactions between albatrosses and the marine environment: species, populations and sexes. In Proceedings of the 22 International Ornitholological Congress, Durban, eds. Adams, N. J. and Slotow, R. H.. South Africa: Birdlife, pp. 2001–20.Google Scholar
Prins, H. H. T. (1987). The buffalo of Manyara. Ph.D. thesis, Rijksuniversiteit te Groningen.
Prins, H. H. T. (1989). Condition changes and choice of social environment in African buffalo bulls. Behaviour, 108, 297–324.CrossRefGoogle Scholar
Prins, H. H. T. (1996). Ecology and Behaviour of the African Buffalo: Social Inequality and Decision Making. London: Chapman & Hall.CrossRefGoogle Scholar
Prins, H. H. T. and Iason, G. R. (1989). Dangerous lions and nonchalant buffalo. Behaviour, 108, 262–95.CrossRefGoogle Scholar
Promislow, D. E. L., Montgomerie, R. and Martin, T. E. (1992). Mortality costs of sexual dimorphism in birds. Proceedings of the Royal Society of London B, 250, 143–50.CrossRefGoogle Scholar
Provenza, F. D. and Balph, D. F. (1988). Development of dietary choice in livestock on rangelands and its implications for management. Journal of Animal Science, 66, 2356–68.CrossRefGoogle Scholar
Pusey, A. E. (1990). Behavioural changes at adolescence in chimpanzees. Behaviour, 115, 203–46.CrossRefGoogle Scholar
Pusey, A. E., Williams, J. and Goodall, J. (1997). The influence of dominance rank on the reproductive success of female chimpanzees. Science, 277, 828–31.CrossRefGoogle ScholarPubMed
Puttick, G. M. (1981). Sex-related differences in foraging behaviour of curlew sandpipers. Ornis Scandinavica, 12, 13–17.CrossRefGoogle Scholar
Qualls, R., Shine, R., Donnellan, S. and Hutchinson, M. (1996). The evolution of viviparity within the Australian scincid lizard Lerista bougainvillii. Journal of Zoology, London, 237, 13–26.CrossRefGoogle Scholar
Quinn, J. S. (1990). Sexual size dimorphism and parental care patterns in a monomorphic and a dimorphic larid. Auk, 107, 260–74.CrossRefGoogle Scholar
Quintana, F. and Dell'Arciprete, O. P. (2002). Foraging grounds of southern giant petrels (Macronectes giganteus) on the Patagonian shelf. Polar Biology, 25, 159–61.CrossRefGoogle Scholar
Radespiel, U. (2000). Sociality in the grey mouse lemur (Microcebus murinus). American Journal of Primatology, 51, 21–40.3.0.CO;2-C>CrossRefGoogle Scholar
Radespiel, U., Cepok, S., Zietemann, V. and Zimmermann, E. (1998). Sex specific usage patterns of sleeping sites in grey mouse lemurs (Microcebus murinus) in northwest Madagascar. American Journal of Primatology, 46, 77–84.3.0.CO;2-S>CrossRefGoogle Scholar
Radespiel, U., Ehresmann, P. and Zimmermann, E. (2001a). Contest versus scramble competition for mates: the composition and spatial structure of a population of grey mouse lemurs (Microcebus murinus) in Madagascar. Primates, 42, 207–20.CrossRefGoogle Scholar
Radespiel, U., Sarikaya, Z., Zimmermann, E. and Bruford, M. (2001b). Sociogenetic structure in a free-living nocturnal primate population: Sex specific differences in the grey mouse lemur (Microcebus murinus). Behavioral Ecology and Sociobiology, 50, 493–502.CrossRefGoogle Scholar
Radford, A. N. and Plessis, M. A. (2003). Bill dimorphism and foraging niche partitioning in the green woodhoopoe. Journal of Animal Ecology, 72, 258–69.CrossRefGoogle Scholar
Rahmani, A. R. (1989). The Great Indian Bustard. Bombay: Bombay Natural History Society.Google Scholar
Rajporohit, L. S., Sommer, V. and Mohnot, S. M. (1995). Wanderers between harems and bachelor bands. Male hanuman langurs (Presbytis entellus) at Jodhpur in Rajasthan. Behaviour, 132, 255–99.CrossRefGoogle Scholar
Ralls, K. and Mesnick, S. L. (2002). Sexual dimorphism. In Encyclopedia of Marine Mammals, eds. Perrin, W. F., Wursig, B. and Thewissen, J. G. M.. San Diego: Academic Press.Google Scholar
Raman, T. R. S. (1997). Factors influencing seasonal and monthly changes in the group size of chital or axis deer in southern India. Journal of Biosciences, 22, 203–18.CrossRefGoogle Scholar
Ramírez Ávila, G. M., Guisset, J. L. and Deneubourg, J. L. (2003). Synchronization in light-controlled oscillators. Physica, 182, 254–73.Google Scholar
Ramirez-Bautista, A. and Benabib, M. (2001). Perch height of the arboreal lizard Anolis nebulosus (Sauria: Polychrotidae) from a tropical dry forest of Mexico: effect of the reproductive season. Copeia, 2001, 187–93.CrossRefGoogle Scholar
Rands, S. A., Cowlishaw, G., Pettifor, R. A., Rowcliffe, J. M. and Johnstone, R. A. (2003). Spontaneous emergence of leaders and followers in foraging pairs. Nature, 423, 432–4.CrossRefGoogle ScholarPubMed
Ransome, R. D. (1990). The Natural History of Hibernating Bats. London: Christopher Helm.Google Scholar
Ranta, E. and Lindström, K. (1990). Assortative schooling in 3-spined sticklebacks. Annales Zoologici Fennici, 27, 67–75.Google Scholar
Ranta, E., Peuhkuri, N. and Laurila, A. (1994). A theoretical exploration of antipredatory and foraging factors promoting phenotype-assorted fish schools. Ecoscience, 1, 99–106.CrossRefGoogle Scholar
Rau, G. H., Ainley, D. G., Bengtson, J. L., Torre, J. J. and Hopkins, T. L. (1992). 15N/14N and 13C/12C in Weddell Sea birds, seals, and fish: implications for diet and trophic structure. Marine Ecology Progress Series, 84, 1–8.CrossRefGoogle Scholar
Read, A. J. (1990). Reproductive seasonality in harbour porpoises, Phocoena phocoena, from the Bay of Fundy. Canadian Journal of Zoology, 68, 284–8.CrossRefGoogle Scholar
Read, A. J. (1999). Harbour porpoise – Phocoena phocoena (Linnaeus, 1758). In Handbook of Marine Mammals: The Second Book of Dolphins and the Porpoises, vol. 6, eds. Ridgway, S. H. and Harrison, R.. London: Academic Press, pp. 323–56.Google Scholar
Read, A. J. and Hohn, A. A. (1995). Life in the fast lane: the life history of harbor porpoises from the Gulf of Maine. Marine Mammal Science, 11, 423–40.CrossRefGoogle Scholar
Read, A. J. and Tolley, K. A. (1997). Postnatal growth and allometry of harbour porpoises from the Bay of Fundy. Canadian Journal of Zoology, 75, 122–30.CrossRefGoogle Scholar
Read, A. J., Wells, R. S., Hohn, A. A. and Tolley, K. A. (1993). Patterns of growth in wild bottlenose dolphins, Tursiops truncatus. Journal of Zoology, London, 231, 107–23.CrossRefGoogle Scholar
Read, D. (1982). Observations of the movements of two arid zone planigales (Dasyuridae, Marsupialia). In Carnivorous Marsupials, ed. Archer, M.. Sydney: Royal Zoological Society of New South Wales, pp. 227–31.Google Scholar
Read, D. G. (1984a). Movements and home ranges of three sympatric dasyurids, Sminthopsis crassicaudata, Planigale gilesi and P. tenuirostris (Marsupialia), in semiarid western New South Wales. Australian Wildlife Research, 11, 223–34.CrossRefGoogle Scholar
Read, D. G. (1984b). Reproduction and breeding season of Planigale gilesi and P. tenuirostris (Marsupialia: Dasyuridae). Australian Mammalogy, 8, 161–74.Google Scholar
Read, D. G. (1989). Microhabitat separation and diel activity patterns of Planigale gilesi and P. tenuirostris (Marsupialia: Dasyuridae). Australian Mammalogy, 12, 45–54.Google Scholar
Read, D. G. (1995). Giles' Planigale. In The Mammals of Australia, ed. Strahan, R.. Sydney: Reed New Holland, pp. 107–9.Google Scholar
Reaney, L. T. and Whiting, M. J. (2002). Life on a limb: ecology of the tree agama (Acanthocercus a. atricollis) in southern Africa. Journal of Zoology, London, 257, 439–48.CrossRefGoogle Scholar
Recchia, C. A. and Read, A. J. (1989). Stomach content of harbour porpoises, Phoceona phoceona, (L.), from the Bay of Fundy. Canadian Journal of Zoology, 67, 2140–6.CrossRefGoogle Scholar
Recher, H. F. and Holmes, R. T. (2000). The foraging ecology of birds of eucalypt forest and woodland. I. Differences between males and females. Emu, 100, 205–15.CrossRefGoogle Scholar
Reddy, E. and Fenton, M. B. (2003). Exploiting vulnerable prey: moths and red bats (Lasiurus borealis; Vespertilionidae). Canadian Journal of Zoology, 18, 1553–60.CrossRefGoogle Scholar
Reebs, S. G. and Saulnier, N. (1997). The effect of hunger on shoal choice in golden shiners (Pisces: Cyprinidae, Notemigonus crysoleucas). Ethology, 103, 642–52.CrossRefGoogle Scholar
Reed, R. N. and Shine, R. (2002). Lying in wait for extinction: ecological correlates of conservation status among Australian elapid snakes. Conservation Biology, 16, 451–61.CrossRefGoogle Scholar
Reid, K. (1995). The diet of Antarctic fur seals Arctocephalus gazella Peters 1875 during winter at South Georgia. Antarctic Science, 7(3), 241–9.CrossRefGoogle Scholar
Reid, K. and Arnould, J. P. Y. (1996). The diet of Antarctic fur seals Arctocephalus gazella during the breeding season at South Georgia. Polar Biology, 16, 105–14.CrossRefGoogle Scholar
Reijinders, P., Brasseur, S., Toorn, J.et al. (1993). Seals, Fur Seals, Sea Lions, and Walrus. Status Survey and Conservation Action Plan. Gland, Switzerland: IUCN.Google Scholar
Reilly, J. J. and Fedak, M. A. (1991). Rates of water turnover and energy-expenditure of free-living male common seals (Phoca vitulina). Journal of Zoology, 223, 461–8.CrossRefGoogle Scholar
Reinert, H. K. (1984). Habitat variation within sympatric snake populations. Ecology, 65, 1673–82.CrossRefGoogle Scholar
Reinert, H. K. and Kodrich, W. R. (1982). Movements and habitat utilization by the massasauga, Sistrurus catenatus catenatus. Journal of Herpetology, 16, 162–71.CrossRefGoogle Scholar
Reinert, H. K. and Zappalorti, R. T. (1988). Timber rattlesnakes (Crotalus horridus) of the Pine Barrens (New Jersey, USA): their movement patterns and habitat preference. Copeia, 1988, 964–78.CrossRefGoogle Scholar
Remis, M. J. (1999). Tree structure and sex differences in arboreality among western lowland gorillas (Gorilla gorilla gorilla) at Bai Hokou, Central African Republic. Primates, 40, 383–96.CrossRefGoogle Scholar
Rendell, L. and Whitehead, H. (2001). Culture in whales and dolphins. Behavioral and Brain Sciences, 24, 309–82.CrossRefGoogle ScholarPubMed
Renfree, M. B., Russell, E. M. and Wooller, R. D. (1984). Reproduction and life history of the Honey Possum, Tarsipes rostratus. In Possums and Gliders, eds. Smith, A. P. and Hume, I. D.. Chipping Norton, NSW: Surrey Beatty and Sons, pp. 427–37.Google Scholar
Rexstad, E. A. and Anderson, D. R. (1988). Effect of the point system on redistributing hunting pressure on mallards. Journal of Wildlife Management, 52, 89–94.CrossRefGoogle Scholar
Reynolds, J. D. and Jennings, S. (2000). The role of animal behavior in marine conservation. In Behavior and Conservation, eds. Gosling, L. M. and Sutherland, W. J.. Cambridge, United Kingdom: Cambridge University Press, pp. 238–57.Google Scholar
Reznick, D. (1996). Life history evolution in guppies: A model system for the empirical study of adaptation. Netherlands Journal of Zoology, 46, 172–90.CrossRefGoogle Scholar
Rhine, R., Bloland, P. and Lodwick, L. (1985). Progression of adult male chacma baboons (Papio ursinus) in the Moremi Wildlife Reserve. International Journal of Primatology, 6, 115–22.CrossRefGoogle Scholar
Rice, D. W. (1998). Marine Mammals of the World, Systematics and Distribution. Lawrence, KA, USA: Society of Marine Mammalogy.Google Scholar
Rice, D. W. (1989). Sperm whale – Physeter macrocephalus (Linnaeus, 1758). In Handbook of Marine Mammals: River Dolphins and the Larger Toothed Whales, vol. 4, eds. Ridgway, S. H. and Harrison, R.. London: Academic Press, pp. 177–233.Google Scholar
Richard, A., Rakotomanaga, P. and Schwartz, M. (1993). Dispersal by Propithecus verrauxi at Beza Mahafaly Reserve. American Journal of Primatology, 30, 1–20.CrossRefGoogle Scholar
Richard, C. and Pépin, D. (1990). Seasonal variation in intragroup-spacing behavior of foraging isards (Rupicapra pyrenaica). Journal of Mammalogy, 71, 145–50.CrossRefGoogle Scholar
Richard, K. R., Dillon, M. C., Whitehead, H. and Wright, J. M. (1996). Patterns of kinship in groups of free-living sperm whales (Physeter macrocephalus) revealed by multiple molecular analyses. Proceedings of the Academy of Science of United States, 93, 8792–5.CrossRefGoogle Scholar
Richard-Hansen, C. (1992). Association between individually marked isards (Rupicapra pyrenaica): seasonal and inter-annual variations. In Ongulés/Ungulates 91, eds. Spitz, F., Janeau, G., Gonzalez, G., and Aulagnier, S.. Paris: S.F.E.P.M.-I.R.G.M., pp. 299–304.Google Scholar
Richards, A. F. (1996). Life history and behavior of female dolphins (Tursiops sp.) in Shark Bay, Western Australia. Ph.D thesis, Ann Arbor, Michigan: University of Michigan.Google Scholar
Richardson, A. J., Maharaj, G., Compagno, L. J. V.et al. (2000). Abundance, distribution, morphometrics, reproduction and diet of the Izak catshark. Journal of Fish Biology, 56, 552–76.CrossRefGoogle Scholar
Ridgway, S. H. and Harrison, R. (1989). Handbook of Marine Mammals: River Dolphins and the Larger Toothed Whales. London: Academic Press.Google Scholar
Ridgway, S. H. and Harrison, R. (1994). Handbook of Marine Mammals: The First Book of Dolphins. London: Academic Press.Google Scholar
Ridgway, S. H. and Harrison, R. (1999). Handbook of Marine Mammals: The Second Book of Dolphins and the Porpoises. London: Academic Press.Google Scholar
Riedman, M. (1990). The Pinnipeds: Seals, Sea Lions, and Walruses. Berkeley: University of California Press.Google Scholar
Rintamäki, P. T., Karvonen, E., Alatalo, R. V. and Lundberg, A. (1999). Why do black grouse males perform on lek sites outside the breeding season?Journal of Avian Biology, 30, 359–66.CrossRefGoogle Scholar
Ripley, W. E. (1946). The soupfin shark and the fishery. California Fish Bulletin, 64, 7–37.Google Scholar
Rivas, J. and Burghardt, G. M. (2001). Understanding sexual size dimorphism in snakes: wearing the snake's shoes. Animal Behaviour, 62, F1–F6.CrossRefGoogle Scholar
Robert, K. A. and Thompson, M. B. (2001). Viviparous lizard selects sex of embryos. Nature, 412, 698–9.CrossRefGoogle ScholarPubMed
Roberts, C., Brotherton, P., Luschekina, A. A., Kholodova, M. V. and Milner-Gulland, E. J. (2001). Gazelles, dwarf antelopes and saigas. In The New Encyclopedia of Mammals, ed. Macdonald, D.. Oxford: Oxford University Press, pp. 560–6.Google Scholar
Robichaud, D. and Rose, G. A. (2003). Sex differences in cod residency on a spawning ground. Fisheries Research, 60, 33–43.CrossRefGoogle Scholar
Rock, J. and Cree, A. (2003). Intraspecific variation in the effect of temperature on pregnancy in the viviparous gecko Hoplodactylus maculatus. Herpetologica, 59, 8–22.CrossRefGoogle Scholar
Rock, J., Andrews, R. M. and Cree, A. (2000). Effects of reproductive condition, season, and site on selected temperatures of a viviparous gecko. Physiological and Biochemical Zoology, 73, 344–55.CrossRefGoogle ScholarPubMed
Rodenhouse, N. L., Sherry, T. W. and Holmes, R. T. (1997). Site-dependent regulation of population size: a new synthesis. Ecology, 78, 2025–42.Google Scholar
Rodhouse, P. G., Elvidge, C. D. and Trathan, P. N. (2001). Remote sensing of the global light-fishing fleet: an analysis of interactions with oceanography, other fisheries and predators. Advances in Marine Biology, 39, 261–303.CrossRefGoogle Scholar
Rodman, P. (1988). Diversity and consistency in ecology and behavior. In Orang-Utan Biology, ed. Schwartz, J. H.. Oxford: Oxford University Press, pp. 31–51.Google Scholar
Rodman, P. S. (1979). Individual activity patterns and the solitary nature of orangutans. In The Great Apes, eds. Hamburg, D. A. and McKown, E. R.. Menlo Park: Benjamin Cummings, pp. 235–56.Google Scholar
Rodman, P. S. and Mitani, J. C. (1987). Orangutans: sexual dimorphism in a solitary species. In Primate Societies, eds. Smuts, B. B., Cheney, D. L., Seyfarth, R. M., Wrangham, R. W. and Struhsaker, T. T.. Chicago: University of Chicago Press, pp. 146–54.Google Scholar
Rodway, M. S., Regehr, H. M. and Cook, F. (2003). Sex and age differences in distribution, abundance, and habitat preferences of wintering harlequin ducks: implications for conservation and estimating recruitment rates. Canadian Journal of Zoology, 81, 492–503.CrossRefGoogle Scholar
Romeo, G., Lovari, S., Festa-Bianchet, M. (1997). Group leaving in mountain goats: are males ousted by adult females?Behavioural Processes, 40, 243–6.CrossRefGoogle ScholarPubMed
Rook, A. J. and Penning, P. D. (1991). Synchronization of eating, ruminating and idling activity by grazing sheep. Applied Animal Behaviour Science, 32, 157–66.CrossRefGoogle Scholar
Roosenburg, W. M., Halcy, K. L. and McGuire, S. (1999). Habitat selection and movements of diamondback terrapins, Malaclemys terrapin, in a Maryland estuary. Chelonian Conservation Biology, 3, 425–9.Google Scholar
Rootes, W. L. and Chabreck, R. H. (1993). Cannibalism in the American alligator. Herpetologica, 49, 99–107.Google Scholar
Rose, B. R. (1981). Factors affecting activity in Sceloporus virgatus. Ecology, 62, 706–16.CrossRefGoogle Scholar
Rose, F. L. and Judd, F. W. (1975). Activity and home range size of the Texas tortoise, Gopherus berlandieri, in South Texas. Herpetologica, 31, 448–56.Google Scholar
Rose, L. (1994). Sex differences in diet and foraging in Cebus capucinus. International Journal of Primatology, 15, 95–114.CrossRefGoogle Scholar
Ross, C. (1992). Basal metabolic rate, body weight, and diet in primates: an evaluation of the evidence. Folia Primatologica, 58, 7–23.CrossRefGoogle Scholar
Ross, D. A. (1989). Population ecology of painted and Blanding's turtles (Chrysemys picta and Emydoidea blandingii) in central Wisconsin. Transactions of the Wisconsin Academy of Science, Arts and Letters, 77, 77–84.Google Scholar
Ross, D. A., Brewster, K. N., Anderson, R. K., Ratner, N. and Brewster, C. M. (1991). Aspects of the ecology of wood turtles, Clemmys insculpta, in Wisconsin. Canadian Field-Naturalist, 105, 363–7.Google Scholar
Ross, P. I., Jalkotzy, M. G. and Festa-Bianchet, M. (1997). Cougar predation on bighorn sheep in southwestern Alberta during winter. Canadian Journal of Zoology, 75, 771–5.CrossRefGoogle Scholar
Rothstein, A. and Griswold, J. G. (1991). Age and sex preferences for social partners by juvenile bison bulls, Bison bison. Animal Behaviour, 41, 227–37.CrossRefGoogle Scholar
Rowe, J. W. and Moll, E. O. (1991). A radiotelemetric study of activity and movements of the Blanding's turtle (Emydoidea blandingii) in northeastern Illinois. Journal of Herpetology, 25, 178–85.CrossRefGoogle Scholar
Rowell, T. E. (1973). Social organization of wild talapoin monkeys. American Journal of Physical Anthropology, 38, 593–8.CrossRefGoogle ScholarPubMed
Rubin, E. S., Boyce, W. M. and Bleich, V. C. (2000). Reproductive strategies of desert bighorn sheep. Journal of Mammalogy, 81, 769–86.2.3.CO;2>CrossRefGoogle Scholar
Rubin, E. S., Boyce, W. M. and Caswell-Chen, E. P. (2002). Modeling demographic processes in an endangered population of bighorn sheep. Journal of Wildlife Management, 66, 796–810.CrossRefGoogle Scholar
Rubin, E. S., Boyce, W. M., Jorgensen, M. C.et al. (1998). Distribution and abundance of bighorn sheep in the Peninsular Ranges, California. Wildlife Society Bulletin, 26, 539–51.Google Scholar
Ruby, D. E. and Baird, D. I. (1994). Intraspecific variation in behavior: comparisons between populations at different altitudes of the lizard Sceloporus jarrovi. Journal of Herpetology, 28, 70–8.CrossRefGoogle Scholar
Ruckstuhl, K. E. (1998). Foraging behaviour and sexual segregation in bighorn sheep. Animal Behaviour, 56, 99–106.CrossRefGoogle ScholarPubMed
Ruckstuhl, K. E. (1999). To synchronise or not to synchronise: a dilemma for young bighorn males?Behaviour, 136, 805–18.CrossRefGoogle Scholar
Ruckstuhl, K. E. and Festa-Bianchet, M. (1998). Do reproductive status and lamb gender affect the foraging behavior of bighorn ewes?Ethology, 104, 941–54.CrossRefGoogle Scholar
Ruckstuhl, K. E. and Festa-Bianchet, M. (2001). Group choice by subadult male bighorn sheep: trade-offs between foraging efficiency and predator avoidance. Ethology, 107, 161–72.CrossRefGoogle Scholar
Ruckstuhl, K. E. and Kokko, H. (2002). Modelling sexual segregation in ungulates: effects of group size, activity budgets and synchrony. Animal Behaviour, 64, 909–14.CrossRefGoogle Scholar
Ruckstuhl, K. E. and Neuhaus, P. (2000). Sexual segregation in ungulates: a new approach. Behaviour, 137, 361–77.CrossRefGoogle Scholar
Ruckstuhl, K. E. and Neuhaus, P. (2001). Behavioral synchrony in ibex groups: effects of age, sex, and habitat. Behaviour, 138, 1033–46.CrossRefGoogle Scholar
Ruckstuhl, K. E. and Neuhaus, P. (2002). Sexual segregation in ungulates: a comparative test of three hypotheses. Biological Review, 77, 77–96.CrossRefGoogle ScholarPubMed
Ruckstuhl, K. E., Manica, A., MacColl, A. D. C., Pilkington, J. G., Clutton-Brock, T. H. (2005). The effects of castration, sex ratio and population density on social segregation and habitat use in Soay sheep. Behavioral Ecology and Sociobiology (in press).
Ruedas, L. A., Demboski, J. R. and Sison, R. V. (1994). Morphological and ecological variation in Otopteropus cartilagonodus Koch, 1969 (Mammalia: Chiroptera: Pteropodidae) from Luzon, Philippines. Proceedings of the Biological Society of Washington, 107, 1–16.Google Scholar
Ruibal, R. and Philibosian, R. (1974). The population ecology of the lizard Anolis acutus. Ecology, 55, 525–37.CrossRefGoogle Scholar
Russell, E. M. (1979). The size and composition of groups in the red kangaroo, Macropus rufus. Australian Wildlife Research, 6, 237–44.CrossRefGoogle Scholar
Russell, E. M. (1986). Observations on the behaviour of the honey possum, Tarsipes rostratus (Marsupialia: Tarsipedidae) in captivity. Australian Journal of Zoology, 121, 1–63.Google Scholar
Russo, D. (2002). Elevation affects the distribution of the two sexes in Daubenton's bats Myotis daubentonii (Chiroptera: Vespertilionidae) from Italy. Mammalia, 66, 543–51.CrossRefGoogle Scholar
Rutherford, P. L. and Gregory, P. T. (2003). How age, sex, and reproductive condition affect retreat-site selection and emergence, patterns in a temperate-zone lizard, Elgaria coerulea. Ecoscience, 10, 24–32.CrossRefGoogle Scholar
Ryan, P. G. and Boix-Hinzen, C. (1999). Consistent male-biased seabird mortality in the Patagonian toothfish longline fishery. Auk, 116, 851–4.CrossRefGoogle Scholar
Sabo, J. L. (2003). Hot rocks or no hot rocks: overnight retreat availability and selection by a diurnal lizard. Oecologia, 136, 329–35.CrossRefGoogle ScholarPubMed
Sachs, B. D. and Harris, V. S. (1978). Sex differences and developmental changes in selected juvenile activities (play) of domestic lambs. Animal Behaviour, 26, 678–84.CrossRefGoogle Scholar
Sailer, L. D., Gaulin, S. J. C., Boster, J. S. and Kurland, J. A. (1985). Measuring the relationship between dietary quality and body size in primates. Primates, 26, 14–27.CrossRefGoogle Scholar
Salamolard, M. and Weimerskirch, H. (1993). Relationship between foraging effort and energy requirement throughout the breeding season in the wandering albatross. Functional Ecology, 7, 643–52.CrossRefGoogle Scholar
Saltz, E., Dixon, D. and Johnson, J. (1977). Training disadvantaged preschoolers on various fantasy activities: Effects on cognitive functioning and impulse control. Child Development, 48, 367–80.CrossRefGoogle ScholarPubMed
Sanderson, R. A. (1974). Sexual dimorphism in the Barbour's map turtle, Malaclemys barbouri (Carr and Marchand). M.A. thesis, Tampa: University of South Florida.Google Scholar
Santos, X. and Llorente, G. A. (1998). Sexual and size-related differences in the diet of the snake Natrix maura from the Ebro Delta, Spain. Herpetological Journal, 8, 161–5.Google Scholar
Santos, X., Gonzalez-Solis, J. and Llorente, G. A. (2000). Variation in the diet of the viperine snake Natrix maura in relation to prey availability. Ecography, 23, 185–92.CrossRefGoogle Scholar
Sato, K., Mitani, Y., Cameron, M. F.et al. (2002). Deep foraging dives in relation to the energy depletion of Weddell seal (Leptonychotes weddellii) mothers during lactation. Polar Biology, 25, 696–702.Google Scholar
Savidge, J. A. (1988). Food habits of Boiga irregularis, an introduced preditor on Guam. Journal of Herpetology, 22, 275–82.CrossRefGoogle Scholar
Sayler, R. D. and Afton, A. D. (1981). Ecological aspects of common goldeneyes Bucephala clangula wintering on the Mississippi River, USA. Ornis Scandinavica, 12, 99–108.CrossRefGoogle Scholar
Schaefer, J. A. and Messier, F. (1995). Habitat selection as a hierarchy: the spatial scales of winter foraging by musk oxen. Ecography, 18, 333–44.CrossRefGoogle Scholar
Schaefer, R. J., Torres, S. G. and Bleich, V. C. (2000). Survivorship and cause-specific mortality in sympatric populations of mountain sheep and mule deer. California Fish and Game, 86, 127–35.Google Scholar
Schmid, J. and Kappeler, P. M. (1998). Fluctuating sexual dimorphism and differential hibernation by sex in a primate, the grey mouse lemur. Behavioral Ecology and Sociobiology, 43, 125–32.CrossRefGoogle Scholar
Schmid-Nielson, K. (1989). Scaling. Why is Animal Size so Important?Cambridge: Cambridge University Press.Google Scholar
Schmidt-Nielsen, K. (1972). Locomotion: energy cost of swimming, flying, and running. Science, 177, 222–8.CrossRefGoogle ScholarPubMed
Schoener, T. W. (1967). The ecological significance of sexual dimorphism in size in the lizard Anolis conspersus. Science, 155, 474–6.CrossRefGoogle ScholarPubMed
Schoener, T. W. (1968). The Anolis lizards of Bimini: resource partitioning in a complex fauna. Ecology, 49, 704–26.CrossRefGoogle Scholar
Schoener, T. W. and Gorman, G. C. (1968). Some niche differences in three lesser Antillean lizards of the genus Anolis. Ecology, 49, 819–30.CrossRefGoogle Scholar
Schoener, T. W., Slade, J. B. and Stinson, C. H. (1982). Diet and sexual dimorphism in the very catholic lizard genus, Leiocephalus of the Bahamas. Oecologia, 53, 160–9.CrossRefGoogle ScholarPubMed
Schreer, J. F. and Kovacs, K. M. (1997). Allometry of diving capacity in air-breathing vertebrates. Canadian Journal of Zoology, 75, 339–58.CrossRefGoogle Scholar
Schwaner, T. D. and Sarre, S. D. (1988). Body size of tiger snakes in southern Australia, with particular reference to Notechis ater serventyi (Elapidae) on Chappell Island. Journal of Herpetology, 22, 24–33.CrossRefGoogle Scholar
Schwartz, O. A., Bleich, V. C. and Holl, S. A. (1986). Genetics and the conservation of mountain sheep Ovis canadensis nelsoni. Biological Conservation, 37, 179–90.CrossRefGoogle Scholar
Schwarzkopf, L. (1994). Measuring trade-offs: a review of studies of costs of reproduction in lizards. In Lizard Ecology: Historical and Experimental Perspectives, eds. Vitt, L. J. and Pianka, E. R.. Princeton, NJ: Princeton University Press, pp. 7–30.CrossRefGoogle Scholar
Schwarzkopf, L. and Shine, R. (1991). Thermal biology of reproduction in viviparous skinks, Eulamprus tympanum: why do gravid females bask more?Oecologia, 88, 562–9.CrossRefGoogle ScholarPubMed
Schwarzkopf, L. and Shine, R. (1992). Costs of reproduction in lizards: escape tactics and susceptibility to predation. Behavioral Ecology and Sociobiology, 31, 17–25.CrossRefGoogle Scholar
Scott, M. D., Wells, R. S. and Irvine, A. B. (1990). A long-term study of bottlenose dolphin on the west coast of Florida. In The Bottlenose Dolphin, eds. Leatherwood, S. and Reeves, R. R.. San Diego: Academic Press, pp. 235–44.Google Scholar
Scott, N. J. Jr, Wilson, D. E., Jones, C. and Andrews, R. M. (1976). The choice of perch dimensions by lizards of the genus Anolis (Reptilia, Lacertilia, Iguanidae). Journal of Herpetology, 10, 75–84.CrossRefGoogle Scholar
Scudder, R. M. and Burghardt, G. M. (1983). A comparative study of defensive behavior in three sympatric species of water snakes (Nerodia). Zeitschrift für Tierpsychologie, 63, 17–26.CrossRefGoogle Scholar
Searcy, W. A. and Yasukawa, K. (1981). Sexual size dimorphism and survival of male and female blackbirds (Icteridae). Auk, 98, 457–65.Google Scholar
Secor, S. M. (1994). Ecological significance of movements and activity range for the sidewinder, Crotalus cerastes. Copeia, 1994, 631–45.CrossRefGoogle Scholar
Seebacher, F., Grigg, G. C. and Beard, L. A. (1999). Crocodiles as dinosaurs: behavioural thermoregulation in very large ectotherms leads to high and stable body temperatures. Journal of Experimental Biology, 202, 77–86.Google ScholarPubMed
Seghers, B. H. (1973). Zoology. The University of British Columbia.Google Scholar
Seghers, B. H. (1974). Schooling behaviour in the guppy (Poecilia reticulata): An evolutionary response to predation. Evolution, 28, 486–9.Google ScholarPubMed
Seib, R. L. (1981). Size and shape in a neotropical burrowing colubrid snake, Geophis nasalis, and its prey. American Zoologist, 21, 933.Google Scholar
Selander, R. K. (1966). Sexual dimorphism and differential niche utilization in birds. Condor, 68, 113–51.CrossRefGoogle Scholar
Selander, R. K. (1972). Sexual selection and dimorphism in birds. In Sexual Selection and the Descent of Man 1971–1971, ed. Campbell, B.. Chicago: Heinemann, pp. 180–229.Google Scholar
Senft, R. L., Coughenour, M. B., Bailey, D. W.et al. (1987). Large herbivore foraging and ecological hierarchies. Bioscience, 37, 789–99.CrossRefGoogle Scholar
Sergeant, D. E. (1973). Biology of white whales (Delphinapterus leucas) in Western Hudson Bay. Fisheries Research Board of Canada, 30, 1065–90.CrossRefGoogle Scholar
Sergeant, D. E. and Brodie, P. F. (1969). Body size in white whales, Delphinapterus leucas. Journal of the Fisheries Research Board of Canada, 26, 2561–80.CrossRefGoogle Scholar
Sexton, O. J. (1964). Differential predation by the lizard, Anolis carolinensis, upon unicolored and polycolored insects after an interval of no contact. Animal Behaviour, 12, 101–10.CrossRefGoogle Scholar
Sexton, O. J., Bauman, J. and Ortleb, E. (1972). Seasonal food habits of Anolis limifrons. Ecology, 53, 182–6.CrossRefGoogle Scholar
Shackleton, D. M. (1991). Social maturation and productivity in bighorn sheep: are young males incompetent. Applied Animal Behaviour Science, 29, 173–84.CrossRefGoogle Scholar
Shaffer, S. A., Weimerskirch, H. and Costa, D. P. (2001). Functional significance of sexual dimorphism in wandering albatrosses, Diomedea exulans. Functional Ecology, 15, 203–10.CrossRefGoogle Scholar
Shah, B., Shine, R., Hudson, S. and Kearney, M. (2003). Sociality in lizards: why do thick-tailed geckos (Nephrurus milii) aggregate?Behaviour, 140, 1039–52.CrossRefGoogle Scholar
Shank, C. C. (1982). Age-sex differences in the diets of wintering Rocky Mountain bighorn sheep. Ecology, 63, 627–33.CrossRefGoogle Scholar
Shank, C. C. (1985). Inter- and intra-sexual segregation of chamois (Rupicapra rupicapra) by altitude and habitat during summer. Zeitschrift für Säugetierkunde, 50, 117–25.Google Scholar
Shealer, D. A. (2002). Foraging behaviour and food of seabirds. In Biology of Marine Birds, eds. Schreiber, E. A. and Burger, J.. Boca Raton, Florida, USA: CRC Press, pp. 137–77.Google Scholar
Shealy, R. M. (1976). The natural history of the Alabama map turtle, Graptemys pulchra Baur, in Alabama. Bulletin of the Florida State Museum, Biological Sciences Series, 21, 47–111.Google Scholar
Sherry, T. W. and Holmes, R. T. 1996. Winter habitat quality, population limitation and conservation of Neotropical-Nearctic migrant birds. Ecology, 77, 36–48.CrossRefGoogle Scholar
Shetty, S. and Shine, R. (2002a). Activity patterns of yellow-lipped sea kraits (Laticauda colubrina) on a Fijian island. Copeia, 2002, 77–85.CrossRefGoogle Scholar
Shetty, S. and Shine, R. (2002b). Sexual divergence in diets and morphology in Fijian sea snakes Laticauda colubrina (Laticaudinae). Austral Ecology, 27, 77–84.CrossRefGoogle Scholar
Shimoka, Y. (2003). Seasonal variation in association patterns of wild spider monkeys (Ateles belzebuth belzebuth) at La Macarena, Columbia. Primates, 44, 83–90.Google Scholar
Shine, R. (1978). Sexual size dimorphism and male combat in snakes. Oecologia, 33, 269–78.CrossRefGoogle ScholarPubMed
Shine, R. (1979). Activity patterns in Australian elapid snakes (Squamata: Serpentes: Elapidae). Herpetologica, 35, 1–11.Google Scholar
Shine, R. (1980a). ‘Costs’ of reproduction in reptiles. Oecologia, 46, 92–100.CrossRefGoogle Scholar
Shine, R. (1980b). Ecology of the Australian death adder, Acanthophis antarcticus (Elapidae): evidence for convergence with the Viperidae. Herpetologica, 36, 281–9.Google Scholar
Shine, R. (1985). The evolution of viviparity in reptiles: an ecological analysis. In Biology of the Reptilia, vol. 15, eds. Gans, C. and Billett, F.. New York: John Wiley and Sons, pp. 605–94.Google Scholar
Shine, R. (1986). Sexual differences in morphology and niche utilization in an aquatic snake, Acrochordus arafurae. Oecologia, 69, 260–7.CrossRefGoogle Scholar
Shine, R. (1989a). Ecological causes for the evolution of sexual dimorphism: a review of the evidence. Quarterly Review of Biology, 64, 419–60.CrossRefGoogle Scholar
Shine, R. (1989b). Constraints, allometry and adaptation: food habits and reproductive biology of Australian brownsnakes (Pseudonaja, Elapidae). Herpetologica, 45, 195–207.Google Scholar
Shine, R. (1991a). Intersexual dietary divergence and the evolution of sexual dimorphism in snakes. American Naturalist, 138, 103–22.CrossRefGoogle Scholar
Shine, R. (1991b). Why do larger snakes eat larger prey?Functional Ecology, 5, 493–502.CrossRefGoogle Scholar
Shine, R. (1994a). Allometric patterns in the ecology of Australian snakes. Copeia, 1994, 851–67.CrossRefGoogle Scholar
Shine, R. (1994b). Sexual size dimorphism in snakes revisited. Copeia, 1994, 326–46.CrossRefGoogle Scholar
Shine, R. (1999). Why is sex determined by nest temperature in many reptiles?Trends in Ecology and Evolution, 14, 186–9.CrossRefGoogle ScholarPubMed
Shine, R. and Crews, D. (1988). Why male garter snakes have small heads: the evolution and endocrine control of sexual dimorphism. Evolution, 42, 1105–10.CrossRefGoogle ScholarPubMed
Shine, R. and Fitzgerald, M. (1995). Variation in mating systems and sexual size dimorphism between populations of the Australian python Morelia spilota (Serpentes: Pythonidae). Oecologia, 103, 490–8.CrossRefGoogle Scholar
Shine, R. and Fitzgerald, M. (1996). Large snakes in a mosaic rural landscape: the ecology of carpet pythons Morelia spilota (Serpentes: Pythonidae) in coastal eastern Australia. Biological Conservation, 76, 113–22.CrossRefGoogle Scholar
Shine, R. and Harlow, P. S. (1993). Maternal thermoregulation influences offspring viability in a viviparous lizard. Oecologia, 96, 122–7.CrossRefGoogle Scholar
Shine, R. and Wall, M. (2004). Why is intraspecific niche partitioning more common in snakes than in lizards? In Foraging Modes in Lizards, eds. Miles, D. B., Reilly, S. M. and McBrayer, L. D.. Cambridge University Press.Google Scholar
Shine, R., Barrott, E. G. and Elphick, M. J. (2002a). Some like it hot: effects of forest clearing on nest temperatures of montane reptiles. Ecology, 83, 2808–15.CrossRefGoogle Scholar
Shine, R., Branch, W. R., Harlow, P. S. and Webb, J. K. (1998a). Reproductive biology and food habits of horned adders, Bitis caudalis (Viperidae), from southern Africa. Copeia, 1998, 391–401.CrossRefGoogle Scholar
Shine, R., Cogger, H. G., Reed, R. N. S., Shetty, S. and Bonnet, X. (2002b). Aquatic and terrestrial locomotor speeds of amphibious sea-snakes (Serpentes, Laticaudidae). Journal of Zoology, London, 259, 261–8.CrossRefGoogle Scholar
Shine, R., Elphick, M. and Donnellan, S. (2002c). Co-occurrence of multiple, supposedly incompatible modes of sex determination in a lizard population. Ecology Letters, 5, 486–9.CrossRefGoogle Scholar
Shine, R., Haagner, G. V., Branch, W. R., Harlow, P. S. and Webb, J. K. (1996). Natural history of the African shieldnose snake Aspidelaps scutatus (Serpentes, Elapidae). Journal of Herpetology, 30, 361–6.CrossRefGoogle Scholar
Shine, R., Harlow, P. S., Keogh, J. S. and Boeadi, (1998b). The influence of sex and body size on food habits of a giant tropical snake, Python reticulatus. Functional Ecology, 12, 248–258.CrossRefGoogle Scholar
Shine, R., Langkilde, T. and Mason, R. T. (2003b). Cryptic forcible insemination: male snakes exploit female physiology, anatomy and behavior to obtain coercive matings. American Naturalist, 162, 653–67.CrossRefGoogle Scholar
Shine, R., O'Connor, D., LeMaster, M. P. and Mason, R. T. (2001). Pick on someone your own size: ontogenetic shifts in mate choice by male garter snakes result in size-assortative mating. Animal Behaviour, 61, 1–9.CrossRefGoogle Scholar
Shine, R., Olsson, M. M., Lemaster, M. P., Moore, I. T. and Mason, R. T. (2000). Effects of sex, body size, temperature, and location on the antipredator tactics of free-ranging gartersnakes (Thamnophis sirtalis, Colubridae). Behavioral Ecology, 11, 239–45.CrossRefGoogle Scholar
Shine, R., Reed, R. N., Shetty, S. and Cogger, H. G. (2002d). Relationships between sexual dimorphism and niche partitioning within a clade of sea-snakes (Laticaudinae). Oecologia, 133, 45–53.CrossRefGoogle Scholar
Shine, R., Shine, T. and Shine, B. G. (2003a). Intraspecific habitat partitioning by the sea snake Emydocephalus annulatus (Serpentes, Hydrophiidae): the effects of sex, body size, and colour pattern. Biological Journal of the Linnean Society, 80, 1–10.CrossRefGoogle Scholar
Shine, R., Sun, L. X., Fitzgerald, M. and Kearney, M. (2003c). A radiotelemetric study of movements and thermal biology of insular Chinese pit-vipers (Gloydius shedaoensis, Viperidae). Oikos, 100, 342–52.CrossRefGoogle Scholar
Shine, R., Phillips, B., Langkilde, T., Lutterschmidt, D. and Mason, R. T. (2004). Mechanisms and consequences of sexual conflict in garter snakes (Thamnophis sirtalis, Colubridae). Behavioral Ecology, 15, 654–60.CrossRefGoogle Scholar
Shively, S. H. (1982). Factors limiting the upstream distribution of the Sabine map turtle. M.Sc. thesis, Lafayette, LA: University of Southwestern Louisiana.Google Scholar
Shively, S. H. and Jackson, J. F. (1985). Factors limiting the upstream distribution of the Sabine map turtle. American Midland Naturalist, 114, 292–303.CrossRefGoogle Scholar
Short, H. L. (1963). Rumen fermentation and energy relationships in white-tailed deer. Journal of Wildlife Management, 27, 184–95.CrossRefGoogle Scholar
Shumway, C. A. (1999). A neglected science: applying behavior to aquatic conservation. Environmental Biology of Fishes, 55, 183–210.CrossRefGoogle Scholar
Silk, J. (1987). Activities and feeding behavior of free-ranging pregnant female baboons. International Journal of Primatology, 8, 596–613.CrossRefGoogle Scholar
Sillet, T. S. and Holmes, R. T. (2002). Variation in survivorship of a migratory songbird throughout its annual cycle. Journal of Animal Ecology, 71, 296–308.CrossRefGoogle Scholar
Simon, C. A. (1975). Size selection of prey by the lizard Sceloporus jarrovi. American Midland Naturalist, 96, 246–51.Google Scholar
Sims, D. W. (1996). The effect of body size on the standard metabolic rate of lesser spotted dogfish, Scyliorhinus canicula. Journal of Fish Biology, 48, 542–4.CrossRefGoogle Scholar
Sims, D. W. (2003). Tractable models for testing theories about natural strategies: foraging behaviour and habitat selection of free-ranging sharks. Journal of Fish Biology, 63 (Supplement A), 53–73.CrossRefGoogle Scholar
Sims, D. W. and Davies, S. J. (1994). Does specific dynamic action (SDA) regulate return of appetite in the lesser spotted dogfish, Scyliorhinus canicula?Journal of Fish Biology, 45, 341–8.Google Scholar
Sims, D. W. and Quayle, V. A. (1998). Selective foraging behaviour of basking sharks on zooplankton in a small-scale front. Nature, 393, 460–4.CrossRefGoogle Scholar
Sims, D. W., Davies, S. J. and Bone, Q. (1993). On the diel rhythms in metabolism and activity of post-hatching lesser spotted dogfish, Scyliorhinus canicula. Journal of Fish Biology, 43, 749–54.CrossRefGoogle Scholar
Sims, D. W., Genner, M. J., Southward, A. J. and Hawkins, S. J. (2001b). Timing of squid migration reflects North Atlantic climate variability. Proceedings of the Royal Society of London B, 268, 2607–11.CrossRefGoogle Scholar
Sims, D. W., Nash, J. P. and Morritt, D. (2001a). Movements and activity of male and female dogfish in a tidal sea lough: alternative behavioural strategies and apparent sexual segregation. Marine Biology, 139, 1165–75.Google Scholar
Sinclair, A. R. E. (1977). The African Buffalo. A Study of Resource Limitation of Populations. Chicago: University of Chicago Press.Google Scholar
Singleton, I. and Schaik, C. P. (2001). Orangutan home range size and its determinants. International Journal of Primatology, 22, 877–912.CrossRefGoogle Scholar
Siniff, D. B., DeMaster, D. P., Hofman, R. J. and Eberhardt, L. L. (1977). An analysis of the dynamics of a Weddell seal population. Ecological Monographs, 47, 319–35.CrossRefGoogle Scholar
Slip, D. J. and Shine, R. (1988a). Feeding habits of the diamond python, Morelia s. spilota: ambush predation by a boid snake. Journal of Herpetology, 22, 323–30.CrossRefGoogle Scholar
Slip, D. J. and Shine, R. (1988b). Habitat use, movements, and activity patterns of free-ranging diamond pythons, Morelia spilota spilota (Serpentes, Boidae): a radiotelemetric study. Australian Wildlife Research, 15, 515–31.CrossRefGoogle Scholar
Slip, D. J., Hindell, M. A. and Burton, H. R. (1994). Diving behaviour of southern elephant seals from Macquarie Island: An overview. In Elephant Seals: Population Ecology, Behaviour and Physiology, eds. Boeuf, B. J. and Laws, R. M.. Berkley: University of California Press, pp. 253–70.Google Scholar
Slotow, R., Dyk, G., Poole, J., Page, B. and Klocke, A. (2000). Older bull elephants control young males. Nature, 408, 425–6.CrossRefGoogle ScholarPubMed
Smith, A. P. and Broome, L. (1992). The effects of season, sex and habitat on the diet of the mountain pygmy-possum (Burramys parvus). Wildlife Research, 19, 755–68.CrossRefGoogle Scholar
Smith, C. (1977). Feeding and ranging behavior of mantled howler monkeys (Alouatta palliata). In Primate Ecology, ed. Clutton-Brock, T. H.. Cambridge: Cambridge University Press, pp. 183–222.Google Scholar
Smith, G. J. D. and Gaskin, D. E. (1983). An environmental index for habitat utilization by female harbour porpoises with calves near Deer Island, Bay of Fundy. Ophelia, 22, 1–13.CrossRefGoogle Scholar
Smith, G. R. (1996). Habitat use and fidelity in the striped plateau lizard Sceloporus virgatus. American Midland Naturalist, 135, 68–80.CrossRefGoogle Scholar
Smith, G. R. and Ballinger, R. E. (1994). Thermal ecology of Sceloporus virgatus from southeastern Arizona, with comparison to Urosaurus ornatus. Journal of Herpetology, 28, 65–9.CrossRefGoogle Scholar
Smith, G. R., Ballinger, R. E. and Congdon, J. D. (1993). Thermal ecology of the high-altitude bunch grass lizard, Sceloporus scalaris. Canadian Journal of Zoology, 71, 2152–5.CrossRefGoogle Scholar
Smith, M. S. R. (1966). Injuries as an indication of social behaviour in the Weddell seal (Leptonychotes wedellii). Mammalia, 30, 241–6.CrossRefGoogle Scholar
Smith, P. C. and Evans, P. R. (1973). Studies of shorebirds at Lindisfarne, Northumberland. 1. Feeding ecology and behaviour of the bar-tailed godwit. Wildfowl, 24, 135–9.Google Scholar
Smith, R. J. and Jungers, W. L. (1997). Body mass in comparative primatology. Journal of Human Evolution, 32, 523–59.CrossRefGoogle ScholarPubMed
Smith, T. G., Hammill, M. O. and Martin, A. R. (1994). Herd composition and behaviour of white whales (Delphinapterus leucas) in two Canadian arctic estuaries. Bioscience, 39, 175–84.Google Scholar
Snell, H. L., Jennings, R. D., Snell, H. M. and Harcourt, S. (1988). Intrapopulation variation in predator avoidance performance of Galapagos lava lizards: the interaction of sexual and natural selection. Evolutionary Ecology, 2, 353–69.CrossRefGoogle Scholar
Snelson, F. F., Mulligan, T. J. and Williams, S. E. (1984). Food habits, occurrence, and population structure of the bull shark, Carcharhinus leucas, in Florida coastal lagoons. Bulletin of Marine Science, 34, 71–80.Google Scholar
Southwell, C. J. (1984). Variability in grouping in the eastern grey kangaroo, Macropus giganteus I. Group density and group size. Australian Wildlife Research, 11, 423–35.CrossRefGoogle Scholar
Spaeth, D. F., Bowyer, R. T., Stephenson, T. R. and Barboza, P. S. (2004). Sexual segregation in moose Alces alces: an experimental manipulation of foraging behaviour. Wildlife Biology, 10, 59–72.CrossRefGoogle Scholar
Speakman, J. R. and Thomas, D. W. (2003). Physiological ecology and energetics of bats. In Bat Ecology, eds. Kunz, T. H. and Fenton, M. B.. Chicago: University of Chicago Press, pp. 430–90.Google Scholar
Speakman, J. R., Irwin, N., Tallach, N. and Stone, R. (1999). Effect of roost size on the emergence behaviour of pipistrelle bats. Animal Behaviour, 58, 787–95.CrossRefGoogle ScholarPubMed
Sprague, D. S., Suzuki, S., Takahashi, H. and Sato, S. (1998). Male life histories in natural populations of Japanese macaques: migration, dominance rank, and troop participation of males in two habitats. Primates, 39, 351–63.CrossRefGoogle Scholar
Springer, M. S., Kirsch, J. A. W. and Case, J. A. (1997). The chronicle of marsupial evolution. In Molecular Evolution and Adaptive Radiation, eds. Givnish, T. J. and Sytsma, K. J.. Cambridge: Cambridge University Press, pp. 129–61.Google Scholar
Springer, S. (1967). Social organization of shark populations. In Sharks, Skates and Rays, eds. Gilbert, P. W., Mathewson, R. F. and Rall, D. P.. Baltimore MD: Johns Hopkins University Press, pp. 149–74.Google Scholar
Sroufe, L. A., Bennett, C., Englund, M., Urban, J. and Shulman, S. (1993). The significance of cross-gender boundaries in preadolescence: contemporary correlates and antecedents of boundary violation and maintenance. Child Development, 64, 455–66.CrossRefGoogle Scholar
Aubin, St D. J., Smith, T. G. and Geraci, J. R. (1990). Seasonal epidermal molt in beluga whales, Delphinapterus leucas. Canadian Journal of Zoology, 68, 359–67.CrossRefGoogle Scholar
Stahl, J. C. and Sagar, P. M. (2000). Foraging strategies of southern Buller's albatrosses Diomedea b. bulleri breeding on The Snares, New Zealand. Journal of the Royal Society of New Zealand, 30, 299–318.CrossRefGoogle Scholar
Staines, B. W. (1976). The use of natural shelter by red deer (Cervus elaphus) in relation to weather in North-East Scotland. Journal of Zoology, 180, 1–8.CrossRefGoogle Scholar
Staines, B. W. (1977). Factors affecting the seasonal distribution of red deer (Cervus elaphus, L.) in Glen Dye, North-East Scotland. Annals of Applied Biology, 87, 495–512.CrossRefGoogle Scholar
Staines, B. W., Crisp, J. M. and Parish, T. (1982). Differences in the quality of food eaten by red deer (Cervus elaphus) stags and hinds in winter. Journal of Applied Ecology, 19, 65–77.CrossRefGoogle Scholar
Stammbach, E. (1987). Desert, forest, and montane baboons: multilevel societies. In Primate Societies, eds. Smuts, B. B., Cheney, D. L., Seyfarth, R. M., Wrangham, R. W. and Struhsaker, T. T.. Chicago: University of Chicago Press, pp. 112–20.Google Scholar
Stamps, J. A. (1977). The function of the survey posture in Anolis lizards. Copeia, 1977, 756–8.CrossRefGoogle Scholar
Stamps, J. A. (1983). Sexual selection, sexual dimorphism, and territoriality. In Lizard Ecology: Studies of a Model Organism, eds. Huey, R. B., Pianka, E. R. and Schoener, T. W.. Cambridge, MA: Harvard University Press, pp. 169–204.CrossRefGoogle Scholar
Staniland, I. J. and Boyd, I. L. (2003). Variation in the foraging location of Antarctic fur seals (Arctocephalus gazella), the effects on diving behaviour. Marine Mammal Science, 19, 331–43.CrossRefGoogle Scholar
Steenbeek, R., Sterck, E. H. M., de Vries, H. and van Hooff, J. A. R. A. M. (2000). Costs and benefits of the one-male, age-graded, and all-male phase in wild Thomas' langurs groups. In Primate Males, ed. Kappeler, P.. Cambridge: Cambridge University Press, pp. 130–45.Google Scholar
Sterck, E. H., Watts, D. P. and Schaik, C. P. (1997). The evolution of social relationships in female primates. Behavioral Ecology and Sociobiology, 41, 291–309.CrossRefGoogle Scholar
Stevens, J. D. (1974). The occurrence and significance of tooth cuts on the blue shark (Prionace glauca L.) from British waters. Journal of the Marine Biological Association of the United Kingdom, 54, 373–8.CrossRefGoogle Scholar
Stevens, J. D. (1976). First results of shark tagging in the north-east Atlantic, 1972–1975. Journal of the Marine Biological Association of the United Kingdom, 56, 929–37.CrossRefGoogle Scholar
Stevick, P. T., Allen, J., Berube, M.et al. (2003). Segregation of migration by feeding ground origin in North Atlantic humpback whales (Megaptera novaeangliae). Journal of Zoology, London, 259, 231–7.CrossRefGoogle Scholar
Stewart, B. S. and Delong, R. L. (1994). Postbreeding foraging migrations of northern elephant seals. In Elephant Seals: Population Ecology, Behaviour, and Physiology, eds. Boeuf, B. J. and Laws, R. M.. Berkley: University of California Press, pp. 290–309.Google Scholar
Stewart, B. S., Craig, M. P. and Antonelis, G. A. (1998). Characterization of Hawaiian Monk Seal (Monachus schauinslandi) pelagic habitat, home range and diving behaviour. Honolulu: National Marine Fisheries Service, Southwest Fisheries Science Center Honolulu Laboratory, 2570 Dole Street, Honolulu, Hawaii 96822–2902.
Stirling, I. (1969). Ecology of the Weddell seal in McMurdo Sound, Antarctica. Ecology, 50, 573–86.CrossRefGoogle Scholar
Stirling, I. (1983). The evolution of mating systems in pinnipeds. In Advances in the Study of Mammalian Behaviour, eds. Eisenberg, J. F. and Kleinmann, D. G.. Lawrence, Kansas: Allen Press, pp. 489–527.Google Scholar
Stoddart, D. and Braithwaite, R. (1979). A strategy for utilization of regenerating heathland habitat by the brown bandicoot (Isoodon obesulus; Marsupialia, Peramelidae). Journal of Animal Ecology, 48, 165–79.CrossRefGoogle Scholar
Stokke, S. (1999). Sex differences in feeding-patch choice in a megaherbivore: elephants in Chobe National Park, Botswana. Canadian Journal of Zoology, 77, 1723–32.CrossRefGoogle Scholar
Stokke, S. and du Toit, J. T. (2000). Sex and size related differences in the dry season feeding patterns of elephants in Chobe National Park, Botswana. Ecography, 23, 70–80.CrossRefGoogle Scholar
Stokke, S. and du Toit, J. T. (2002). Sexual segregation in habitat use by elephants in Chobe National Park, Botswana. African Journal of Ecology, 40, 360–71.CrossRefGoogle Scholar
Storz, J. F. and Williams, C. F. (1996). Summer population structure of sub-alpine bats in Colorado. Southwestern Naturalist, 41, 322–4.Google Scholar
Storz, J. F., Bhat, H. R. and Kunz, T. H. (2000). Social structure of a polygynous tent-making bat, Cynopterous sphinx (Megachiroptera). Journal of Zoology, London, 251, 151–65.CrossRefGoogle Scholar
Strahan, R. (1995). The Mammals of Australia, 2nd edn. Sydney: Reed New Holland.Google Scholar
Straits, B. C. (1998). Occupational sex segregation: the role of personal ties. Journal of Vocational Behavior, 52(2), 191–207.CrossRefGoogle Scholar
Strikwerda, T. E., Fuller, M. R., Seegar, W. S., Howey, P. W. and Black, H. D. (1986). Bird-borne satellite transmitter and location program. Johns Hopkins APL Technical Digest, 7, 203–8.Google Scholar
Suhonen, J. and Kuitunen, M. (1991). Intersexual foraging niche differentiation within the breeding pair in the common treecreeper Certhia familiaris. Ornis Scandinavica, 22, 313–18.CrossRefGoogle Scholar
Sukumar, R. and Gadgil, M. (1988). Male-female differences in foraging on crops by Asian elephants. Animal Behaviour, 36, 1233–5.CrossRefGoogle Scholar
Summers, R. W., Westlake, G. E. and Feare, C. J. (1987). Differences in the ages, sexes and physical condition of starlings Sturnus vulgaris at the centre and periphery of roosts. Ibis, 129, 96–102.CrossRefGoogle Scholar
Sun, L. X., Shine, R., Zhao, D. B. and Tang, Z. R. (2002). Low costs, high output: reproduction in an insular pit-viper (Gloydius shedaoensis, Viperidae) from north-eastern China. Journal of Zoology, London, 256, 511–21.Google Scholar
Sund, O. (1943). Et brugdebarsel. Naturen, 67, 285–6.Google Scholar
Swennen, C. (1984). Differences in quality of roosting flocks of oystercatchers. In Coastal Waders and Wildfowl in Winter, eds. Evans, P. R., Goss-Custard, J. D. and Hale, W. G.. Cambridge: Cambridge University Press.Google Scholar
Swennen, C., Bruijn, L. L. M., Duiven, P., Leopold, M. F. and Marteijn, E. C. L. (1983). Differences in bill form of the oystercatcher Haematopus ostralegus; a dynamic adaptation to specific foraging techniques. Netherlands Journal of Sea Research, 17, 57–83.CrossRefGoogle Scholar
Swingland, I. R. and Lessells, C. M. (1979). The natural regulation of giant tortoise populations on Aldabra Atoll: movement polymorphism, reproductive success and mortality. Journal of Animal Ecology, 48, 639–54.CrossRefGoogle Scholar
Sydeman, W. J. and Nur, N. (1994). Life history strategies of female northern elephant seals. In Elephant Seals: Population Ecology, Behaviour and Physiology, eds. Boeuf, B. J. and Laws, R. M.. Berkley: University of California Press, pp. 137–53.Google Scholar
Talbot, J. J. (1979). Time budget, niche overlap, inter- and intraspecific aggression in Anolis humilis and A. limifrons from Costa Rica. Copeia, 1979, 472–81.CrossRefGoogle Scholar
Tamisier, A. (1985). Hunting as a key environmental parameter for the Western Palearctic duck populations. Wildfowl, 36, 95–103.Google Scholar
Taylor, R. J. (1981). The comparative ecology of the Eastern grey kangaroo and wallaroo in the New England tablelands of New South Wales. Ph.D. thesis, University of New England.Google Scholar
Taylor, R. J. (1983). Association of social classes of the wallaroo, Macropus robustus (Marsupialia: Macropodidae). Australian Wildlife Research, 10, 39–45.CrossRefGoogle Scholar
Tedman, R. A. and Bryden, M. M. (1979). Cow-pup behaviour of the Weddell seal, Leptonychotes weddelli (Pinnipedia), in McMurdo Sound, Antarctica. Australian Journal of Wildlife Research, 6, 19–37.CrossRefGoogle Scholar
Terborgh, J. (1990). Where Have All the Birds Gone?Princeton: Princeton University Press.Google Scholar
Terranova, M. L. and Laviola, G. (1995). Individual differences in mouse behavioural development: effects of precious weaning and ongoing sexual segregation. Animal Behaviour, 50, 1261–71.CrossRefGoogle Scholar
Testa, J. W. (1994). Over winter movements and diving behaviour of female Weddell seal (Leptonychotes weddellii) in the Southwestern Ross Sea, Antarctica. Canadian Journal of Zoology, 72, 1700–10.CrossRefGoogle Scholar
Theodorakis, C. W. (1989). Size segregation and the effects of oddity on predation risk in minnow schools. Animal Behaviour, 38, 496–502.CrossRefGoogle Scholar
Thill, R. E., Martin, A. Jr, Morris, H. F. Jr and , McCune E. D. (1987). Grazing and burning impacts on deer diets on Louisiana pine-bluestem range. Journal of Wildlife Management, 51, 873–80.CrossRefGoogle Scholar
Thomas, C. D., Baguette, M. and Lewis, O. T. (2000). Butterfly movement and conservation in patchy landscapes. In Behavior and Conservation, eds. Gosling, L. M. and Sutherland, W. J.. Cambridge, United Kingdom: Cambridge University Press, pp. 85–104.Google Scholar
Thomas, D. W. and Cloutier, D. (1992). Evaporative water loss by hibernating little brown bats, Myotis lucifugus. Physiological Ecology, 65, 433–56.Google Scholar
Thomas, L. N. (1987). The effects of stress on some aspects of the demography and physiology of Isoodon obesulus (Shaw and Nodder). Unpublished Masters thesis, University of Western Australia.Google Scholar
Thompson, P. M., Fedak, M. A., McConnell, B. J. and Nicholas, K. S. (1989). Seasonal and sex-related variation in the activity patterns of common seals (Phoca vitulina). Journal of Applied Ecology, 27, 521–35.CrossRefGoogle Scholar
Thompson, P. M., Mackay, A., Tollit, D. J., Enderby, S. and Hammond, P. S. (1998). The influence of body size and sex on the characteristics of harbour seal foraging trips. Canadian Journal of Zoology – Revue Canadienne De Zoologie, 76, 1044–53.CrossRefGoogle Scholar
Thompson, P. M., Tollit, D. J., Wood, D.et al. (1997). Estimating harbour seal abundance and distribution in an esturine habitat in N.E. Scotland. Journal of Applied Ecology, 34, 43–52.CrossRefGoogle Scholar
Thorhallsdottir, A. G., Provenza, F. D. and Balph, D. F. (1990). Ability of lambs to learn about novel foods while observing or participating with social models. Applied Animal Behaviour Science, 25, 25–33.CrossRefGoogle Scholar
Tickell, W. L. N. (1968). The biology of the great albatrosses Diomedea exulans and Diomedea epomophora. Antarctic Research Series, 12, 1–55.Google Scholar
Tickell, W. L. N. (2000). Albatrosses. East Sussex, England: Pica Press.Google Scholar
Tieszen, L. L. and Imbamba, S. K. (1980). Photosynthetic systems, carbon isotope discrimination and herbivore selectivity in Kenya. African Journal of Ecology, 18, 237–42.CrossRefGoogle Scholar
Tolley, K. A., Read, A. J., Well, R. S.et al. (1995). Sexual dimorphism in wild bottlenose dolphins (Tursiops truncatus) from Sarasota, Florida. Journal of Mammalogy, 76, 1190–8.CrossRefGoogle Scholar
Townshend, D. J. (1981). The importance of field feeding to the survival of wintering male and female curlews Numenius arquata on the Tees estuary. In Feeding and Survival Strategies of Estuarine Organisms, eds. Jones, N. V. and Wolff, W. J.. New York: Plenum Press.CrossRefGoogle Scholar
Tricas, T. C. and Feuvre, E. M. (1985). Mating in the reef white-tip shark Triaenodon obesus. Marine Biology, 84, 233–7.CrossRefGoogle Scholar
Trivers, R. (1971). The evolution of reciprocal altruism. Quarterly Review of Biology, 46, 35–57.CrossRefGoogle Scholar
Trivers, R. L. (1972). Parental investment and sexual selection. In Sexual Selection and the Descent of Man, ed. Campbell, B.. Chicago: Aldine, pp. 1871–971.Google Scholar
Tuck, G. N., Polacheck, T. and Bulman, C. (2003). Spatio-temporal trends of longline fishing effort in the Southern Ocean and implications for seabird by-catch. Biological Conservation, 114, 1–27.CrossRefGoogle Scholar
Tucker, A. D., FitzSimmons, N. N. and Gibbons, J. W. (1995). Resource partitioning by the estuarine turtle Malaclemys terrapin: trophic, spatial, and temporal foraging constraints. Herpetologica, 51, 167–81.Google Scholar
Tucker, A. D., Limpus, C. J., McCallum, H. I. and McDonald, K. R. (1997). Movements and home ranges of Crocodylus johnstoni in the Lynd River, Queensland. Wildlife Research, 24, 379–96.CrossRefGoogle Scholar
Tucker, A. D., McCallum, H. I., Limpus, C. J. and McDonald, K. R. (1998). Sex-biased dispersal in a long-lived polygynous reptile (Crocodylus johnstoni). Behavioral Ecology and Sociobiology, 44, 85–90.CrossRefGoogle Scholar
Tuttle, M. D. (1979). Status, causes of decline, and management of endangered grey bats. Journal of Wildlife Management, 43, 1–17.CrossRefGoogle Scholar
Twente, J. W. (1955). Some aspects of habitat selection and other behaviour of cavern-dwelling bats. Ecology, 36, 706–32.CrossRefGoogle Scholar
Tyndale-Biscoe, H. and Renfree, M. (1987). Reproductive Physiology of Marsupials. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Uetz, P. (2000). How many reptile species?Herpetological Review, 31, 13–15.Google Scholar
Lichtenbelt, Marken W. D., Wesselingh, R. A., Vogel, J. T. and Albers, K. B. M. (1993). Energy budgets in free-living green iguanas in a seasonal environment. Ecology, 74, 1157–72.CrossRefGoogle Scholar
van Noordwijk, A. J. (1994). The interaction of inbreeding depression and environmental stochasticity in the risk of extinction of small populations. In Conservation Genetics, eds. Loeschcke, V., Tomiuk, J. and Jain, S. K.. Basel, Switzerland: Birkhauser Verlag, pp. 131–46.CrossRefGoogle Scholar
Noordwijk, M. and Schaik, C. P. (2002). Career moves: transfers and rank challenge decisions by male long-tailed macaques. Behaviour, 138, 359–95.CrossRefGoogle Scholar
van Noordwijk, M., Hemelrijk, C. K., Herremans, L. A. M. and Sterck, E. H. M. (1993). Spatial position and behavioral sex differences in juvenile long-tailed macaques. In Juvenile Primates, eds. Pereira, M. E. and Fairbanks, L. A.. Oxford: Oxford University Press, pp. 77–85.Google Scholar
Schaik, C. P. (1983). Why are diurnal primates living in groups?Behaviour, 37, 120–44.CrossRefGoogle Scholar
van Schaik, C. P. (1989). The ecology of social relationships amongst female primates. In Comparative Socioecology, eds. Standen, V. and Foley, R.. London: Blackwell, pp. 195–218.Google Scholar
Schaik, C. P. (1999). The fission-fusion social system of orangutans. Primates, 40, 69–86.CrossRefGoogle Scholar
van Schaik, C. P. (2000a). Vulnerability to infanticide by males: patterns among mammals. In Infanticide by Males and Its Implications, eds. Schaik, C. P. and Janson, C. H.. Cambridge: Cambridge University Press, pp. 61–71.CrossRefGoogle Scholar
van Schaik, C. P. (2000b). Infanticide by male primates: the sexual selection hypothesis revisited. In Infanticide by Males and Its Implications, eds. Schaik, C. P. and Janson, C. H.. Cambridge: Cambridge University Press, pp. 27–60.CrossRefGoogle Scholar
Schaik, C. P. and Kappeler, P. M. (1997). Infanticide risk and the evolution of male-female association in primates. Proceedings of the Royal Society, London, B, 264, 1687–94.CrossRefGoogle ScholarPubMed
Schaik, C. P. and Noordwijk, M. A. (1989). The special role of male Cebus monkeys in predation avoidance and its effects on group composition. Behavioral Ecology and Sociobiology, 24, 265–76.CrossRefGoogle Scholar
Soest, P. J. (1982). Nutritional Ecology of the Ruminant. Corvallis: O and B Books.Google Scholar
Soest, P. J. (1994). Nutritional Ecology of the Ruminant, 2nd edn. Ithaca: Cornell University Press.Google Scholar
Soest, P. J. (1996). Allometry and ecology of feeding behavior and digestive capacity in herbivores. Zoo Biology, 15, 455–79.3.0.CO;2-A>CrossRefGoogle Scholar
Vaughan, T. A. (1976). Nocturnal behaviour of the African false vampire bat (Cardioderma cor). Journal of Mammalogy, 57, 227–48.CrossRefGoogle Scholar
Vaughan, T. A. and O'Shea, T. J. (1976). Roosting ecology of the pallid bat, Antrozous pallidus. Journal of Mammalogy, 57, 227–48.CrossRefGoogle Scholar
Vaughan, T. A. and Vaughan, R. P. (1986). Seasonality and the behaviour of the African yellow-winged bat (Lavia frons). Journal of Mammalogy, 67, 91–102.CrossRefGoogle Scholar
Vaughan, T. A. and Vaughan, R. P. (1987). Parental behaviour in the African yellow-winged bat (Lavia frons). Journal of Mammalogy, 68, 217–23.CrossRefGoogle Scholar
Vehrencamp, S. L., Stiles, F. G. and Bradbury, J. W. (1977). Observations on the foraging behaviour and avian prey of the neotropical carnivorous bat, Vampyrum spectrum. Journal of Mammalogy, 58, 469–78.CrossRefGoogle Scholar
Verme, L. J. (1988). Niche selection by male white-tailed deer: an alternative hypothesis. Wildlife Society Bulletin, 16, 448–51.Google Scholar
Vernes, K. and Pope, L. C. (2001). Stability of nest range, home range and movements of the northern bettong (Bettongia tropica) following moderate-intensity fire in a tropical woodland, north-eastern Queensland. Wildlife Research, 28, 141–50.CrossRefGoogle Scholar
Vernes, K. and Pope, L. C. (2002). Fecundity, pouch young survivorship and breeding season of the northern bettong (Bettongia tropica) in the wild. Australian Mammalogy, 23, 95–100.CrossRefGoogle Scholar
Vidal, R. M., Macias-Caballero, C. and Duncan, C. D. (1994). The occurrence and ecology of the golden-cheeked warbler in the highlands of Northern Chiapas, Mexico. Condor, 96, 684–91.Google Scholar
Viitanen, P. (1967). Hibernation and seasonal movements of the viper, Vipera berus berus (L.) in southern Finland. Annales Zoologici Fennici, 4, 472–546.Google Scholar
Villard, M.-A., Martin, P. R. and Drummond, C. G. (1993). Habitat fragmentation and pairing success in the ovenbird, Seiurus aurocapillus. Auk, 110, 759–68.Google Scholar
Villaret, J. C. and Bon, R. (1995). Social and spatial segregation in alpine ibex (Capra ibex) in Bargy, French Alps. Ethology, 101, 291–300.CrossRefGoogle Scholar
Villaret, J. C. and Bon, R. (1998). Sociality and relationships in Alpine ibex (Capra ibex). Revue d'Ecologie (Terre Vie), 53, 153–70.Google Scholar
Villaret, J. C., Rivet, A. and Bon, R. (1997). Sexual segregation of habitat by the alpine ibex in the French Alps. Journal of Mammalogy, 78, 1273–81.CrossRefGoogle Scholar
Vincent, A. and Sadovy, Y. (1998). Reproductive ecology in the conservation and management of fishes. In Behavioral Ecology and Conservation Biology, ed. Caro, T.. New York: Oxford University Press, pp. 209–45.Google Scholar
Vincent, S. E., Herrel, A. and Irschick, D. J. (2004). Ontogeny of intersexual head shape and prey selection in the pitviper Agkistrodon piscivorus. Biological Journal of the Linnean Society, 81, 151–9.CrossRefGoogle Scholar
Visagie, L. (2001). Grouping behaviour in the armadillo lizard, Cordylus cataphractus. M.Sc. thesis, Matieland, South Africa: University of Stellenbosch.Google Scholar
Vitt, L. J. and Cooper, W. E. (1986). Foraging and diet of a diurnal predator (Eumeces laticeps) feeding on hidden prey. Journal of Herpetology, 20, 408–15.CrossRefGoogle Scholar
Vitt, L. J. and Zani, P. A. (1996). Ecology of the elusive tropical lizard Tropidurus [equals Uracentron] flaviceps (Tropiduridae) in lowland rain forest of Ecuador. Herpetologica, 52, 121–32.Google Scholar
Vladykov, V. D. (1946). Nourriture du marsouin blanc ou béluga (Delphinapterus leucas) du fleuve Saint-Laurent. Etudes sur les mammifères aquatiques (IV). Québec: Département des pêcheries, Province de Québec, pp. 129.Google Scholar
Vogt, R. C. (1981). Food partitioning in three sympatric species of map turtle, genus Graptemys (Testudinata, Emydidae). American Midland Naturalist, 105, 102–11.CrossRefGoogle Scholar
Voigt, C. C. and Streich, W. J. (2003). Queuing for harem access in colonies of the greater sac-winged bat. Animal Behaviour, 65, 149–56.CrossRefGoogle Scholar
Voisin, J. (1968). Les pétrels géants (Macronectes halli et Macronectes giganteus) de l'ile de la Possession. L'Oiseau, 38, 7–122.Google Scholar
Voisin, J. (1991). Sur le régime et l'écologie alimentaires des pétrels géants Macronectes halli et M. giganteus de l'archipel Crozet. L'Oiseau, 61, 39–49.Google Scholar
Voisin, J. and Bester, M. N. (1981). The specific status of Giant petrels Macronectes at Gough Island. In Proceedings of the Symposium on Birds on the Sea and Shore, 1979, ed. Cooper, J.. Cape Town: African Seabird Group, pp. 215–22.Google Scholar
Volkman, N. J., Presler, P. and Trivelpiece, W. (1980). Diets of pygoscelid penguins at King George Island, Antarctica. Condor, 82, 373–8.CrossRefGoogle Scholar
Wagenknecht, E. (1986). Rotwild. Malsungen: Neumann-Neudamm Verlag.Google Scholar
Wagner, R. H. (1997). Differences in prey species delivered to nestlings by male and female razorbills Alca torda. Seabird, 19, 58–9.Google Scholar
Walker, B. G. and Bowen, W. D. (1993). Changes in body-mass and feeding-behavior in male harbor seals, Phoca vitulina, in relation to female reproductive status. Journal of Zoology, 231, 423–36.CrossRefGoogle Scholar
Wanless, S., Corfield, T., Harris, M. P., Buckland, S. T. and Morris, J. A. (1993). Diving behaviour of the shag (Aves: Pelecaniformes) in relation to water depth and prey size. Journal of Zoology, London, 231, 11–25.CrossRefGoogle Scholar
Wanless, S., Finney, S. K., Harris, M. P. and McCafferty, D. J. (1999). Effect of the diel light cycle on the diving behaviour of two bottom feeding marine birds: the blue-eyed shag Phalacrocorax atriceps and the European shag P. aristotelis. Marine Ecology Progress Series, 188, 219–24.CrossRefGoogle Scholar
Wanless, S., Harris, M. P., Morris, J. A. (1995). Factors affecting daily activity budgets of South-Georgian shags during chick rearing at Bird Island, South Georgia. Condor, 97, 550–8.CrossRefGoogle Scholar
Wapstra, E., Olsson, M., Shine, R.et al. (2004). Maternal basking behaviour determines offspring sex in a viviparous reptile. Biology Letters, 271, S230–S232.Google Scholar
Ward, A. J. W. and Hart, P. J. B. (2003). The effects of kin and familiarity on interactions between fish. Fish and Fisheries, 4, 348–58.CrossRefGoogle Scholar
Ward, S. J. (1990). Life history of the feathertail glider, Acrobates pygmaeus (Acrobatidae: Marsupialia) in south-eastern Australia. Australian Journal of Zoology, 38, 503–17.CrossRefGoogle Scholar
Ward, S. J. and Renfree, M. B. (1988). Reproduction in females of the feathertail glider, Acrobates pygmaeus (Marsupialia). Journal of Zoology, London, 216, 225–39.CrossRefGoogle Scholar
Wardle, C. S. (1993). Fish behaviour and fishing gear. In Behaviour of Teleost Fishes, ed. Pitcher, T. J.. London: Chapman and Hall, pp. 607–43.CrossRefGoogle Scholar
Warham, J. (1990). The Petrels: Their Ecology and Breeding Systems. London: Academic Press.Google Scholar
Waser, P. M. (1977). Feeding, ranging, and group size in the mangabey, Cercocebus alibigena. In Primate Ecology, ed. Clutton-Brock, T. H.. Cambridge: Cambridge University Press, pp. 183–222.Google Scholar
Waser, P. M. (1985). Spatial structure in mangabey groups. International Journal of Primatology, 6, 569–80.CrossRefGoogle Scholar
Waterman, J. M. (1995). The social organization of the Cape ground squirrel (Xerus inauris; Rodentia: Sciuridae). Ethology, 101, 130–47.CrossRefGoogle Scholar
Waterman, J. M. (1997). Why do male Cape ground squirrels live in groups?Animal Behaviour, 53, 809–17.CrossRefGoogle Scholar
Waters, J. C. (1974). The biological significance of the basking habit in the black-knobbed sawback, Graptemys nigronoda Cagle. M.Sc. thesis, Auburn, AL: Auburn University.Google Scholar
Watkins, J. L., Buchholz, F., Priddle, J., Morris, D. J. and Ricketts, C. (1992). Variation in reproductive status of Antarctic krill swarms evidence for a size-related sorting mechanism?Marine Ecology Progress Series, 82, 163–74.CrossRefGoogle Scholar
Watson, A. and Staines, B. W. (1978). Differences in the quality of wintering areas used by male and female red deer (Cervus elaphus) in Aberdeenshire. Journal of Zoology, London, 186, 544–50.Google Scholar
Watt, E. M. and Fenton, M. B. (1995). DNA fingerprinting provides evidence of discriminant suckling and non-random mating in little brown bats, Myotis lucifugus. Molecular Ecology, 4, 261–4.CrossRefGoogle Scholar
Watts, D. P. (1984). Composition and variability of mountain gorilla diets in the central Virungas. American Journal of Primatology, 7, 325–56.CrossRefGoogle Scholar
Watts, D. P. (1988). Environmental influences on mountain gorilla time budgets. American Journal of Primatology, 15, 295–312.CrossRefGoogle Scholar
Watts, D. P. (1991). Ecology of gorillas and its relationship to female transfer in mountain gorillas. International Journal of Primatology, 11, 21–45.CrossRefGoogle Scholar
Watts, D. P. and Mitani, J. C. (2001). Boundary patrols and intergroup encounters in wild chimpanzees. Behaviour, 138, 299–327.CrossRefGoogle Scholar
Watts, P. S., Hansen, S. and Lavigne, D. M. (1993). Models of heat loss by marine mammals: Thermoregulation below the zone of irrelevance. Journal of Theoretical Biology, 163, 505–25.CrossRefGoogle Scholar
Webb, G. J. W. and Smith, A. M. A. (1984). Sex ratio and survivorship in the Australian freshwater crocodile, Crocodylus johnstoni. Symposia of the Zoological Society of London, 52, 319–55.Google Scholar
Webb, G. J. W., Buckworth, R. and Manolis, S. C. (1983). Crocodylus johnstoni in the McKinlay River area, N. T. III. Growth, movement and the population age structure. Australian Wildlife Research, 10, 383–401.CrossRefGoogle Scholar
Webb, J. K. and Shine, R. (1997). A field study of spatial ecology and movements of a threatened snake species, Hoplocephalus bungaroides. Biological Conservation, 82, 203–17.CrossRefGoogle Scholar
Webb, J. K., Brown, G. P. and Shine, R. (2001). Body size, locomotor speed and antipredator behaviour in a tropical snake (Tropidonophis mairii, Colubridae): the influence of incubation environments and genetic factors. Functional Ecology, 15, 561–8.CrossRefGoogle Scholar
Webb, R. G. (1961). Observations on the life histories of turtles (genus Pseudemys and Graptemys) in Lake Texoma, Oklahoma. American Midland Naturalist, 65, 193–214.CrossRefGoogle Scholar
Weckerly, F. W. (1993). Intersexual resource partitioning in black-tailed deer: a test of the body size hypothesis. Journal of Wildlife Management, 57, 475–94.CrossRefGoogle Scholar
Weckerly, F. W. (1998). Sexual-size dimorphism: influence of mass and mating systems in the most dimorphic mammals. Journal of Mammalogy, 79, 33–52.CrossRefGoogle Scholar
Weckerly, F. W., Ricca, M. A. and Meyer, K. P. (2001). Sexual segregation in Roosevelt elk: cropping rates and aggression in mixed-sex groups. Journal of Mammalogy, 82, 825–35.2.0.CO;2>CrossRefGoogle Scholar
Weeden, R. B. (1964). Spatial separation of sexes in rock and willow ptarmigan in winter. Auk, 81, 534–41.CrossRefGoogle Scholar
Weidt, A., Hagenah, N., Randrianambinina, B., Radespiele, U. and Zimmerman, E. (2004). Social organization of the golden brown mouse lemur (Microcebus ravelobensis). American Journal of Physical Anthropology, 123, 40–51.CrossRefGoogle Scholar
Weihs, D. and Webb, P. W. (1983). Optimisation of locomotion. In Fish Biomechanics, eds. Webb, P. W. and Weihs, D.. New York: Praeger, pp. 339–71.Google Scholar
Weilgart, L. S. and Whitehead, H. (1986). Observations of a sperm whale (Physeter catodon) birth. Journal of Mammalogy, 67, 399–401.CrossRefGoogle Scholar
Weilgart, L., Whitehead, H. and Payne, K. (1996). A colossal convergence. American Scientist, 84, 278–87.Google Scholar
Weimerskirch, H. (1991). Sex-specific differences in molt strategy in relation to breeding in the wandering albatross. Condor, 93, 731–7.CrossRefGoogle Scholar
Weimerskirch, H. (1992). Reproductive effort in long-lived birds: age-specific patterns of condition, reproduction and survival in the wandering albatross. Oikos, 64, 464–73.CrossRefGoogle Scholar
Weimerskirch, H. (1995). Regulation of foraging trips and incubation routine in male and female wandering albatrosses. Oecologia, 102, 37–43.CrossRefGoogle ScholarPubMed
Weimerskirch, H. (1998). Foraging strategies of Indian Ocean albatrosses and their relationships with fisheries. In Albatross Biology and Conservation, eds. Robertson, G. and Gales, R.. Chipping Norton, Australia: Surrey Beatty and Sons, pp. 168–79.Google Scholar
Weimerskirch, H. and Jouventin, P. (1987). Population dynamics of the wandering albatross, Diomedea exulans, of the Crozet Islands: causes and consequences of the population decline. Oikos, 49, 315–22.CrossRefGoogle Scholar
Weimerskirch, H. and Jouventin, P. (1998). Changes in population sizes and demographic parameters of six albatross species breeding on the French sub-Antarctic islands. In Albatross Biology and Conservation, eds. Robertson, G. and Gales, R.. Chipping Norton, Australia: Surrey Beatty and Sons, pp. 84–91.Google Scholar
Weimerskirch, H. and Wilson, R. P. (2000). Oceanic respite for wandering albatrosses. Nature, 406, 955–6.CrossRefGoogle ScholarPubMed
Weimerskirch, H., Barbraud, C. and Lys, P. (2000). Sex differences in parental investment and chick growth in Wandering Albatrosses: Fitness consequences. Ecology, 81, 309–18.CrossRefGoogle Scholar
Weimerskirch, H., Bonadonna, F., Bailleul, F.et al. (2002). GPS tracking of foraging albatrosses. Science, 295, 1259–69.CrossRefGoogle ScholarPubMed
Weimerskirch, H., Brothers, N. and Jouventin, P. (1997a). Population dynamics of wandering albatross Diomedea exulans and the Amsterdam albatross D. amsterdamensis in the Indian Ocean and their relationships with long-line fisheries: conservation implications. Biological Conservation, 79, 257–70.CrossRefGoogle Scholar
Weimerskirch, H., Cherel, Y., Cuenot-chaillet, F. and Ridoux, V. (1997b). Alternative foraging strategies and resource allocation by male and female wandering albatrosses. Ecology, 78, 2051–63.CrossRefGoogle Scholar
Weimerskirch, H., Lequette, B. and Jouventin, P. (1989). Development and maturation of plumage in the wandering albatross Diomedea exulans. Journal of Zoology, London, 219, 411–21.CrossRefGoogle Scholar
Weimerskirch, H., Salamolard, M., Sarrazin, F. and Jouventin, P. (1993). Foraging strategy of wandering albatrosses through the breeding season: a study using satellite telemetry. Auk, 110, 325–42.Google Scholar
Wells, R. S. (2003). Dolphin social complexity: lessons from long-term study and life history. In Animal Social Complexity: Intelligence, Culture, and Individualized Societies, eds. Waal, F. B. M. and Tyack, P. L.. Cambridge, Massachusetts: Harvard University Press, pp. 32–56.CrossRefGoogle Scholar
Wells, R. S. and Scott, M. S. (1999). Bottlenose dolphin – Tursiops truncatus (Montagu, 1821). In Handbook of Marine Mammals: The Second Book of Dolphins and Porpoises, vol. 6, eds. Ridgway, S. H. and Harrison, R.. London: Academic Press, pp. 137–82.Google Scholar
Wells, R. S., Irvine, A. B. and Scott, M. D. (1980). The social ecology of inshore odontocetes. In Cetacean Behavior: Mechanisms and Functions, ed. Herman, L. M.. New York: John Wiley and Sons, pp. 263–317.Google Scholar
Wells, R. S., Scott, M. D. and Irvine, A. B. (1987). The social structure of free-ranging Bottlenose dolphins. In Current Mammalogy, vol. 1, ed. Genoways, H. H.. New York: Plenum Press, pp. 247–305.CrossRefGoogle Scholar
Wells, R. T. (1978). Field observations of the hairy-nosed wombat, Lasiorhinus latifrons (Owen). Australian Wildlife Research, 5, 299–303.CrossRefGoogle Scholar
Wetherbee, B. M., Gruber, S. H. and Cortes, E. (1990). Diet, feeding habits, digestion and consumption in sharks, with special reference to the lemon shark, Negaprion brevirostris. In Elasmobranchs as Living Resources: Advances in the Biology, Ecology, Systematics and Status of the Fisheries, NOAA Technical Report 90, eds. Pratt, H. L., Gruber, S. H. and Taniuchi, T.. Seattle, WA: National Oceanic and Atmospheric Administration, pp. 29–47.Google Scholar
Whitaker, P. B. and Shine, R. (2001). Thermal biology and activity patterns of the eastern brownsnake (Pseudonaja textilis): a radiotelemetric study. Herpetologica, 58, 436–62.CrossRefGoogle Scholar
Whitaker, P. B. and Shine, R. (2003). A radiotelemetric study of movements and shelter-site selection by free-ranging brownsnakes (Pseudonaja textilis, Elapidae). Herpetological Monographs, 17, 130–44.CrossRefGoogle Scholar
White, M. and Kolb, J. A. (1974). A preliminary study of Thamnophis near Sagehagen Creek, California. Copeia, 1974, 126–36.CrossRefGoogle Scholar
Whitehead, H. (1993). The behaviour of mature male sperm whales on the Galapagos Islands breeding grounds. Canadian Journal of Zoology, 71, 689–99.CrossRefGoogle Scholar
Whitehead, H. (1996). Babysitting, dive synchrony, and indications of alloparental care in sperm whales. Behavioral Ecology and Sociobiology, 38, 237–44.CrossRefGoogle Scholar
Whitehead, H. (2003). Sperm Whales: Social Evolution in the Ocean. Chicago: The University of Chicago Press.Google Scholar
Whitehead, H. and Mann, J. (2000). Female reproductive strategies of cetaceans: Life histories and calf care. In Cetacean Societies: Field Studies of Dolphins and Whales, eds. Mann, J., Connor, R. C., Tyack, P. L. and Whitehead, H.. Chicago: University of Chicago Press, pp. 219–47.Google Scholar
Whitehead, H. and Weilgart, L. (2000). The sperm whale: Social females and roving males. In Cetacean Societies: Field Studies of Dolphins and Whales, eds. Mann, J., Connor, R. C., Tyack, P. L. and Whitehead, H.. Chicago: University of Chicago Press, pp. 154–73.Google Scholar
Whitehead, H., Waters, S. and Lyrholm, T. (1991). Social organization of female sperm whales and their offspring: constant companions and casual acquaintances. Behavioral Ecology and Sociobiology, 29, 385–9.CrossRefGoogle Scholar
Whiting, B. and Edwards, C. (1973). A cross-cultural analysis of sex-differences in the behavior of children age 3 through 11. Journal of Social Psychology, 91, 171–88.CrossRefGoogle Scholar
Whiting, M. J. and Greeff, J. M. (1997). Facultative frugivory in the Cape flat lizard, Platysaurus capensis (Sauria: Cordylidae). Copeia, 1997, 811–18.CrossRefGoogle Scholar
Wickings, E. J. and Dixon, A. F. (1992). Testicular function, secondary sexual development, and social status in male mandrills (Mandrillus sphinx). Physiology and Behavior, 52, 909–16.CrossRefGoogle Scholar
Wikelski, M. and Trillmich, F. (1994). Foraging strategies of the Galapagos marine iguana (Amblyrhynchus cristatus): adapting behavioral rules to ontogenic size change. Behaviour, 128, 255–79.CrossRefGoogle Scholar
Wilkinson, G. S. (1992). Information transfer at evening bat colonies. Animal Behaviour, 44, 501–18.CrossRefGoogle Scholar
Wilkinson, G. S. (1995). Information transfer in bats. In Ecology, Evolution and Behaviour of Bats, eds. Racey, P. A. and Swift, S. M.. Symposium of the Zoological Society of London 67. Oxford: Clarendon Press, pp. 345–360.Google Scholar
Wilkinson, L. C. and Barclay, R. M. R. (1997). Differences in the foraging behaviour of male and female big brown bats (Eptesicus fuscus) during the reproductive period. Ecoscience, 4, 279–85.CrossRefGoogle Scholar
Wilkinson, G. S. and Boughman, J. W. (1998). Social calls co-ordinate foraging in greater spear-nosed bats. Animal Behaviour, 55, 337–50.CrossRefGoogle Scholar
Williams, A. J. (1982). Sexual size-dimorphism in the growth of Macronectes. Acta XVIII Congressus Internationalis Ornithologici, 2, 1190–1.Google Scholar
Williams, J. M., Oehlert, G. W., Carlis, J. V. and Pusey, A. E. (2004). Why do male chimpanzees defend a group range?Animal Behaviour, 68, 523–32.CrossRefGoogle Scholar
Williams, J. M., Pusey, A. E., Carlis, J. V., , Goodall J. and Farm, B. E. (2002). Space use and group membership in female chimpanzees: alternative strategies. Animal Behaviour, 63, 347–60.CrossRefGoogle Scholar
Williams, T. A. and Christiansen, J. L. (1981). The niches of two sympatric softshell turtles, Trionyx muticus and Trionyx spiniferus, in Iowa. Journal of Herpetology, 15, 30–8.CrossRefGoogle Scholar
Williams, T. C., Ireland, L. C. and Williams, J. M. (1973). High altitude flights of the free-tailed bat, Tadarida brasiliensis observed with radar. Journal of Mammalogy, 54, 807–21.CrossRefGoogle Scholar
Williams, T. D. (1991). Foraging ecology and diet of gentoo penguins Pygoscelis papua at South Georgia during winter and an assessment of their winter prey consumption. Ibis, 133, 3–13.CrossRefGoogle Scholar
Williams, T. M. (1999). The evolution of low cost efficient swimming in marine mammals: limits to energetic optimization. Philosophical Transactions of the Royal Society of London, 354, 193–201.CrossRefGoogle Scholar
Wilson, B. (1995). The ecology of bottlenose dolphin in Moray Firth, Scotland: a population at the northern extreme of the species' range. Ph.D. thesis, Scotland: University of Aberdeen.
Wilson, E. O. (1980). Sociobiology: The Abridged Edition. Cambridge: Belknap Press.Google Scholar
Wilson, M. and Wrangham, R. W. (2003). Intergroup relations in chimpanzees. Annual Review of Anthropology, 32, 363–92.CrossRef
Wilson, R. P., Copper, J. and Plotz, J. (1992). Can we determine when marine endotherms feed? A case study of seabirds. Journal of Experimental Biology, 167, 267–75.Google Scholar
Wilson, M. L., Wallauer, W. R. and Pusey, A. E. (2004). Intergroup violence in chimpanzees: new cases from Gombe National Park, Tanzania. International Journal of Primatology, 25, 523–50.CrossRefGoogle Scholar
Wirtz, P. and Morato, T. (2001). Unequal sex ratios in longline catches. Journal of Marine Biology, 81, 1–2.Google Scholar
Witter, M. S. and Cuthill, I. C. (1993). The ecological costs of avian fat storage. Philosophical Transactions of the Royal Society of London B, 340, 73–90.CrossRefGoogle ScholarPubMed
Wolf, N. G. (1985). Odd fish abandon mixed-species groups when threatened. Behavioral Ecology and Sociobiology, 17, 47–52.CrossRefGoogle Scholar
Wood, A. G., Naef-Daenzer, B., Prince, P. A. and Croxall, J. P. (2000). Quantifying habitat use in satellite-tracked pelagic seabirds: application of kernel estimation to albatross locations. Journal of Avian Biology, 31, 278–86.CrossRefGoogle Scholar
Wood, C. M. and McDonald, D. G. (1997). Global Warming: Implications for Freshwater and Marine Fish. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Woodroffe, R. and Ginsberg, J. R. (2000). Ranging behaviour and vulnerability to extinction in carnivores. In Behavior and Conservation, eds. Gosling, L. M. and Sutherland, W. J.. Cambridge, United Kingdom: Cambridge University Press, pp. 125–40.Google Scholar
Woolington, D. W. (1993). Sex ratios of canvasbacks wintering in Louisiana. Journal of Wildlife Management, 57, 751–8.CrossRefGoogle Scholar
Wooller, R. D., Richardson, K. C., Garavanta, C. A. M., Saffer, V. M. and Bryant, K. A. (2000). Opportunistic breeding in the polyandrous honey possum, Tarsipes rostratus. Australian Journal of Zoology, 48, 669–80.CrossRefGoogle Scholar
Woolnough, A. P. and du Toit, J. T. (2001). Vertical zonation of browse quality in tree canopies exposed to a size-structured guild of African browsing ungulates. Oecologia, 129, 585–90.CrossRefGoogle ScholarPubMed
Wootton, R. J. (1976) The Biology of the Sticklebacks. London: Academic Press.Google Scholar
Wootton, R. J. (1998) Ecology of Teleost Fishes. Dordrecht, The Netherlands: Kluwer Academic Pulishers.Google Scholar
Worrell, E. (1958). Song of the Snake. Sydney: Angus and Robertson.Google Scholar
Wourms, J. P. and Demski, L. S. (1993). The reproduction and development of sharks, skates, rays and ratfishes: introduction, history, overview, and future prospects. Environmental Biology of Fishes, 38, 7–21.CrossRefGoogle Scholar
Wrangham, R. W. (1979). On the evolution of ape social systems. Social Science Information, 18, 335–68.CrossRefGoogle Scholar
Wrangham, R. W. (1999). Evolution of coalitionary killing. Yearbook of Physical Anthropology, 42, 1–30.3.0.CO;2-E>CrossRefGoogle Scholar
Wrangham, R. W. (2000). Why are male chimpanzees more gregarious than mothers? A scramble competition hypothesis. In Primate Males, ed. Kappeler, P. M.. Cambridge: Cambridge University Press, pp. 248–58.Google Scholar
Wrangham, R. W. and Rubenstein, D. I. (1986). Social evolution in birds and mammals. In Ecological Aspects of Social Evolution, ed. Wrangham, R.. Princeton, New Jersey: Princeton University Press, pp. 452–71.Google Scholar
Wrangham, R. W. and Smuts, B. B. (1980). Sex differences in the behavioural ecology of chimpanzees in the Gombe National Park, Tanzania. Journal of Reproduction and Fertility, Suppl. 28, S13–S31.Google ScholarPubMed
Wrangham, R. W., Chapman, C. A., Clark-Arcadi, A. P. and Isabirye-Basuta, G. (1996). Social ecology of Kanyawara chimpanzees: implications for understanding the costs of great ape groups. In Great Ape Societies, eds. McGrew, W. C., Marchant, L. F. and Toshisada, N.. Cambridge: Cambridge University Press, pp. 45–57.CrossRefGoogle Scholar
Wright, P. H. (1988). Interpreting research on gender differences in friendship: A case for moderation and plea for caution. Journal of Social and Personal Relationships, 5(3), 367–73.CrossRefGoogle Scholar
Wright, P. H. and Scanlon, M. B. (1991). Gender role orientations and friendship: some attenuation, but gender differences abound. Sex Roles, 24(9–10), 551–66.CrossRefGoogle Scholar
Wrona, F. J. and , Dixon W. J. (1991). Group size and predation risk: a field analysis of encounter and dilution effects. American Naturalist, 137, 186–201.CrossRefGoogle Scholar
Wunderle, J. M. (1995). Population characteristics of black-throated blue warblers wintering in three sites on Puerto Rico. Auk, 112, 931–46.CrossRefGoogle Scholar
Würsig, B. and Bastida, R. (1986). Long-range movement and individual associations of two dusky dolphins (Lagenorhynchus obscurus) off Argentina. Journal of Mammalogy, 67, 773–4.CrossRefGoogle Scholar
Würsig, B. and Würsig, M. (1980). Behavior and ecology of the Dusky dolphin, Lagenorhynchus obscurus, in the South Atlantic. Fishery Bulletin, 77, 871–90.Google Scholar
Xavier, J. C., Croxall, J. P., Trathan, P. N. and Rodhouse, P. G. (2003a). Inter-annual variation in the cephalopod component of the diet of wandering albatrosses Diomedea exulans breeding at Bird Island, South Georgia. Marine Biology, 142, 611–22.CrossRefGoogle Scholar
Xavier, J. C., Croxall, J. P. and Reid, K. (2003b). Interannual variation in the diets of two albatross species breeding at South Georgia: implications for breeding performance. Ibis, 145, 593–610.CrossRefGoogle Scholar
Xavier, J. C., Croxall, J. P., Trathan, P. N. and Wood, A. G. (2003c). Feeding strategies and diets of breeding grey-headed and wandering albatrosses at South Georgia. Marine Biology, 143, 221–32.CrossRefGoogle Scholar
Xavier, J. C., Trathan, P. N., Croxall, J. P.et al. (2004). Foraging ecology and interactions with fisheries of wandering albatrosses (Diomedea exulans) breeding at South Georgia. Fisheries Oceanography, 13(5), 324–44.CrossRefGoogle Scholar
Yamagiwa, J. (1987). Intra- and inter-group interactions of an all-male group of Virunga mountain gorillas (Gorilla gorilla beringei). Primates, 28, 1–30.CrossRefGoogle Scholar
Yearsley, J. M. and Pérez-Barbería, F. J. (2005). Does the activity budget hypothesis explain sexual segregation in ungulates?Animal Behaviour, 69, 257–67.CrossRefGoogle Scholar
Young, D. D. and Cockcroft, V. G. (1994). Diet of common dolphins (Delphinus delphis) off the south-east of southern Africa: opportunism or specialization?Journal of Zoology, London, 234, 41–53.CrossRefGoogle Scholar
Young, T. P. and Isbell, L. A. (1991). Sex-differences in giraffe feeding ecology – energetic and social constraints. Ethology, 87, 79–89.CrossRefGoogle Scholar
Yurk, H., Barrett-Lennard, L., Ford, J. K. B. and Matkins, C. O. (2002). Cultural transmission within maternal lineages: vocal clans in resident killer whales in southern Alaska. Animal Behaviour, 63, 1103–19.CrossRefGoogle Scholar
Zari, T. A. (1987). The energetics and thermal physiology of the Wiegmann's skink, Mabuya brevicollis. Ph.D. thesis, Nottingham, UK: University of Nottingham.Google Scholar
Zari, T. A. (1998). Effects of sexual condition on food consumption and temperature selection in the herbivorous desert lizard, Uromastyx philbyi. Journal of Arid Environments, 38, 371–7.CrossRefGoogle Scholar
Zharikov, Y. and Skilleter, G. A. (2002). Sex-specific intertidal habitat use in subtropically wintering bar-tailed godwits. Canadian Journal of Zoology, 80, 1918–29.CrossRefGoogle Scholar
Zinner, H. (1985). On behavioral and sexual dimorphism of Telescopus dhara Forscal 1776 (Reptilia: Serpentes, Colubridae). Journal of the Herpetological Association of Africa, 31, 5–6.CrossRefGoogle Scholar
Zuberbühler, J. K. D. and Bshary, R. (1999). The predator deterrence function of primate alarm calls. Ethology, 105, 477–90.CrossRefGoogle Scholar
Zucker, N. (1986). Perch height preferences of male and female tree lizards, Urosaurus ornatus: a matter of food competition or social role?Journal of Herpetology, 20, 547–53.CrossRefGoogle Scholar
Zuk, M. and McKean, K. A. (1996). Sex differences in parasite infections: patterns and processes. International Journal of Parasitology, 26, 1009–24.CrossRefGoogle ScholarPubMed
Zwarts, L. (1988). Numbers and distribution of coastal waders in Guinea-Bissau. Ardea, 76, 42–55.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 Kathreen Ruckstuhl, University of Cambridge, Peter Neuhaus, University of Cambridge
  • Book: Sexual Segregation in Vertebrates
  • Online publication: 04 September 2009
  • Chapter DOI: https://doi.org/10.1017/CBO9780511525629.022
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 Kathreen Ruckstuhl, University of Cambridge, Peter Neuhaus, University of Cambridge
  • Book: Sexual Segregation in Vertebrates
  • Online publication: 04 September 2009
  • Chapter DOI: https://doi.org/10.1017/CBO9780511525629.022
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 Kathreen Ruckstuhl, University of Cambridge, Peter Neuhaus, University of Cambridge
  • Book: Sexual Segregation in Vertebrates
  • Online publication: 04 September 2009
  • Chapter DOI: https://doi.org/10.1017/CBO9780511525629.022
Available formats
×