Skip to main content Accessibility help
×
Hostname: page-component-76fb5796d-x4r87 Total loading time: 0 Render date: 2024-04-26T23:12:55.570Z Has data issue: false hasContentIssue false

Part I - GPS for Primatologists

Published online by Cambridge University Press:  29 January 2021

Francine L. Dolins
Affiliation:
University of Michigan, Dearborn
Christopher A. Shaffer
Affiliation:
Grand Valley State University, Michigan
Leila M. Porter
Affiliation:
Northern Illinois University
Jena R. Hickey
Affiliation:
University of Georgia
Nathan P. Nibbelink
Affiliation:
University of Georgia
Get access
Type
Chapter
Information
Spatial Analysis in Field Primatology
Applying GIS at Varying Scales
, pp. 7 - 120
Publisher: Cambridge University Press
Print publication year: 2021

Access options

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

References

References

Asensio, N., Schaffner, C. M., and Aureli, F. 2015. Quality and overlap of individual core areas are related to group tenure in female spider monkeys. American Journal of Primatology 77(7): 777785.Google Scholar
Bayart, F. and Simmen, B. 2005. Demography, range use, and behavior in black lemurs (Eule mur macaco macaco) at Ampasikely, northwest Madagascar. American Journal of Primatology 67(3): 299312.CrossRefGoogle Scholar
Brockmeyer, T., Kappeler, P. M., Willaume, E., et al. 2015. Social organization and space use of a wild mandrill (Mandrillus sphinx) group. American Journal of Primatology 77(10): 10361048.Google Scholar
Chiarello, A. G. 1995. Role of loud calls in brown howlers, Alouatta fusca. American Journal of Primatology 36(3): 213222.CrossRefGoogle ScholarPubMed
Corbin, G. D. and Schmid, J. 1995. Insect secretions determine habitat use patterns by a female lesser mouse lemur (Microcebus murinus). American Journal of Primatology 37(4): 317324.CrossRefGoogle ScholarPubMed
de la Torre, S., Campos, F., and Devries, T. 1995. Home-range and birth seasonality of Saguinus nicricollis graelssi in Ecuadorian Amazonia. American Journal of Primatology 37(1): 3956.Google Scholar
Di Bitetti, M. S., Vidal, E. M. L., Baldovino, M. C., and Benesovsky, V. 2000. Sleeping site preferences in tufted capuchin monkeys (Cebus apella nigritus). American Journal of Primatology 50(4): 257274.Google Scholar
Fan, P. F., Garber, P., Chi, M., et al. 2015. High dietary diversity supports large group size in Indo-Chinese gray langurs in Wuliangshan, Yunnan, China. American Journal of Primatology 77(5): 479491.Google Scholar
Fashing, P. J. and Cords, M. 2000. Diurnal primate densities and biomass in the Kakamega Forest: an evaluation of census methods and a comparison with other forests. American Journal of Primatology 50(2): 139152.Google Scholar
Hill, D. A. and Agetsuma, N. 1995. Supra-annual variation in the influence of Myrica rubra fruit on the behavior of a troop of Japanese macaques in Yakushima. American Journal of Primatology 35(3): 241250.Google Scholar
José-Dominguez, J. M., Savini, T., and Asensio, N. 2015. Ranging and site fidelity in northern pigtailed macaques (Macaca leonina) over different temporal scales. American Journal of Primatology 77(8): 841853.Google Scholar
Kurihara, Y. and Hanya, G. 2015. Comparison of feeding behavior between two different-sized groups of Japanese macaques (Macaca fuscata yakui). American Journal of Primatology 77(9): 9861000.Google Scholar
Phoonjampa, R., Koenig, A., Borries, C., Gale, G. A., and Savini, T. 2010. Selection of sleeping trees in pileated gibbons (Hylobates pileatus). American Journal of Primatology 72(7): 617625.CrossRefGoogle ScholarPubMed
Pimley, E. R., Bearder, S. K., and Dixson, A. F. 2005. Social organization of the Milne-Edward’s potto. American Journal of Primatology 66(4): 317330.CrossRefGoogle ScholarPubMed
R Core Team 2013. R: A Language and Environment for Statistical Computing. R Foundation for Statistical Computing, Vienna.Google Scholar
Riley, E. P., MacKinnon, K. C., Fernandez-Duque, E., Setchell, J. M., and Garber, P. A. 2014. Code of best practices for field primatology. Resource document, International Primatological Society & American Society of Primatologists.Google Scholar
Schwab, D. 2000. A preliminary study of spatial distribution and mating system of pygmy mouse lemurs (Microcebus cf myoxinus). American Journal of Primatology 51(1): 4160.3.0.CO;2-7>CrossRefGoogle ScholarPubMed
Spehar, S. N., Link, A., and Di Fiore, A. 2010. Male and female range use in a group of white-bellied spider monkeys (Ateles belzebuth) in Yasuni National Park, Ecuador. American Journal of Primatology 72(2): 129141.Google Scholar
Sprague, D. S. 2000. Topographic effects on spatial data at a Japanese macaque study site. American Journal of Primatology 52(3): 143147.3.0.CO;2-W>CrossRefGoogle Scholar
Van Belle, S. 2015. Female participation in collective group defense in black howler monkeys (Alouatta pigra). American Journal of Primatology 77(6): 595604.Google Scholar
Wong, C. L. and Ni, I. H. 2000. Population dynamics of the feral macaques in the Kowloon Hills of Hong Kong. American Journal of Primatology 50(1): 5366.Google Scholar
Yepez, P., de la Torre, S., and Snowdon, C. T. 2005. Interpopulation differences in exudate feeding of pygmy marmosets in Ecuadorian Amazonia. American Journal of Primatology 66(2): 145158.Google Scholar
Zhang, S. Y. 1995. Sleeping habits of brown capuchin monkeys (Cebus apella) in French Guiana. American Journal of Primatology 36(4): 327335.Google Scholar

References

Bolstad, P. 2016. GIS Fundamentals: A First Text on Geographic Information Systems, 5th edition. XanEdu Publishing Inc., Ann Arbor, MI.Google Scholar
Burrough, P. A., McDonnell, R. A., and Lloyd, C. D. 2015. Principles of Geographical Information Systems. Oxford University Press, Oxford.Google Scholar
Clarke, K. C. 1986. Advances in geographic information systems. Computers, Environment and Urban Systems 10(3–4): 175184.CrossRefGoogle Scholar
Chang, K. 2012. Introduction to Geographic Information Systems, 6th edition. McGraw-Hill, New York.Google Scholar
Channan, S., Collins, K., and Emanuel, W. R. 2014. Global Mosaics of the Standard MODIS Land Cover Type Data. University of Maryland and the Pacific Northwest National Laboratory, College Park, MD.Google Scholar
Coppock, J. T. and Rhind, D. W. 1991. The history of GIS. Geographical Information Systems: Principles and Applications 1(1): 2143.Google Scholar
Environmental Systems Research Institute (ESRI). 1990. Understanding GIS: The ARC/INFO Method. ESRI, Redlands, CA.Google Scholar
Goodchild, M. F. 1992. Geographical information science. International Journal of Geographical Information Systems 6(1): 3145.CrossRefGoogle Scholar
Hickey, J. R., Nackoney, J., Nibbelink, N. P., et al. 2013. Human proximity and habitat fragmentation are key drivers of the rangewide bonobo distribution. Biodiversity and Conservation 22: 30853104.Google Scholar
Hofmann-Wellenhof, B., Lichtenegger, H., and Collins, J. 2001. Global Positioning System: Theory and Practice. Springer Science & Business Media, New York.Google Scholar
Howard, A. M., Nibbelink, N. P., Madden, M., et al. 2015. Landscape influences on the natural and artificially manipulated movements of bearded capuchin monkeys. Animal Behaviour 106: 5970.Google Scholar
Longley, P. A., Goodchild, M. F., Maguire, D. J., and Rhind, D. W. 2001. Geographic Information Systems and Science, Wiley, New York.Google Scholar
Longley, P. A., Goodchild, M. F., Maguire, D. J., and Rhind, D. W. 2005. Geographic Information Systems and Science. Wiley, Chichester.Google Scholar
Madden, M. (Ed.) 2009. Manual of Geographic Information Systems. American Society for Photogrammetry and Remote Sensing, Bethesda, MD.Google Scholar
Mark, D. M. 2003. Geographic information science: defining the field. Pages 318 in Foundations of Geographic Information Science. Duckham, M., Goodchild, M. F., and Worboys, M. (Eds.). Taylor & Francis, New York.Google Scholar
McGarigal, K., Cushman, S. A., and Ene, E. 2012. FRAGSTATS v4: Spatial Pattern Analysis Program for Categorical and Continuous Maps. Computer software program produced by the authors at the University of Massachusetts, Amherst. Available at: www.umass.edu/landeco/research/fragstats/fragstats.html.Google Scholar
Monmonier, M. 1996. How to Lie With Maps, 2nd edition. University of Chicago Press, Chicago, IL.Google Scholar
Parkinson, B. W. 1996. GPS error analysis. Pages 478483 in Global Positioning System: Theory and Applications, Volume II. Parkinson, B. W. and Spilker, J. J. Jr. (Eds.). American Institute of Astronautics and Aeronautics, Washington, DC.Google Scholar
R Core Team. 2016. R: A Language and Environment for Statistical Computing. R Foundation for Statistical Computing, Vienna.Google Scholar
Tao, C. V., and Li, J. 2007. Advances in Mobile Mapping Technology. Taylor & Francis, London.Google Scholar
Tilman, D., and Kareiva, P. (Eds.) 2001. Spatial Ecology. Princeton University Press, Princeton, NJ.Google Scholar
Tomlinson, R. F. 2007. Thinking about GIS: Geographic Information System Planning for Managers. ESRI, Redlands, CA.Google Scholar
Turner, M. G., Gardner, R. H., and O’Neill, R. V. 2001. Landscape Ecology in Theory and Practice. Springer, New York.Google Scholar
US Geological Survey. 1996. 30 arc-second DEM of South America. Digital elevation model. US Geological Survey’s Center for Earth Resources Observation and Science (EROS).Google Scholar
Wiens, J. A. 1989. Spatial scaling in ecology. Functional Ecology 3(4): 385397.Google Scholar

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 258: 131137.Google Scholar
Adams, A. L., Dickinson, K. J. M., Robertson, B. C., and van Heezik, Y. 2013. An evaluation of the accuracy and performance of lightweight GPS collars in a suburban environment. PLoS ONE 8. DOI: 10.1371/journal.pone.0068496.Google Scholar
Adelman, J. S., Moyers, S. C., and Hawley, D. M. 2014. Using remote biomonitoring to understand heterogeneity in immune-responses and disease-dynamics in small, free-living animals. Integrative and Comparative Biology 54: 377386.Google Scholar
Alexander, R. D. 1974. The evolution of social behavior. Annual Review of Ecology and Systematics 5: 325383.Google Scholar
Arseneau, T. J. M., Taucher, A.-L., van Schaik, C. P., and Willems, E. P. 2015. Male monkeys fight in between-group conflicts as protective parents and reluctant recruits. Animal Behaviour 110: 3950.Google Scholar
Bowles, S. 2009. Did warfare among ancestral hunter-gatherers affect the evolution of human social behaviors? Science 324: 12931298.CrossRefGoogle ScholarPubMed
Breed, G. A., Costa, D. P., Goebel, M. E., and Robinson, P. W. 2011. Electronic tracking tag programming is critical to data collection for behavioral time-series analysis. Ecosphere 2. DOI: 10.1890/ES10-00021.1.Google Scholar
Bridge, E. S., Thorup, K., Bowlin, M. S., et al. 2011. Technology on the move: recent and forthcoming innovations for tracking migratory birds. BioScience 61: 689698.Google Scholar
Brown, D., LaPoint, S., Kays, R., et al. 2012. Accelerometer-informed GPS telemetry: reducing the trade-off between resolution and longevity. Wildlife Society Bulletin 36: 139146.Google Scholar
Brown, D., Kays, R., Wikelski, M., Wilson, R., and Klimley, A. 2013. Observing the unwatchable through acceleration logging of animal behavior. Animal Biotelemetry 1: 116.Google Scholar
Brown, M. and Crofoot, M. C. 2013. Social and spatial relationships between primate groups. Pages 151176 in Primate Ecology and Conservation: A Handbook of Techniques. Sterling, E. J., Bynum, N., and Blair, M. E. (Eds.). Oxford University Press, Oxford.Google Scholar
Burger, A. E. and Shaffer, S. A. 2008. Application of tracking and data-logging technology in research and conservation of seabirds. Auk 125: 253264.Google Scholar
Byrne, R. W. 2000. How monkeys find their way: leadership, coordination, and cognitive maps of African baboons. Pages 491518 in On the Move: How and Why Animals Travel in Groups. Boinski, S. and Garber, P.. Chicago University Press, Chicago, IL.Google Scholar
Cain, J. W., Krausman, P. R., Jansen, B. D., and Morgart, J. R. 2005. Influence of topography and GPS fix interval on GPS collar performance. Wildlife Society Bulletin 33: 926934.Google Scholar
Caine, N. G. 1989. Unrecognized anti-predator behaviour can bias observational data. Animal Behaviour 39: 195197.CrossRefGoogle Scholar
Campbell, A. F. and Sussman, R. W. 1994. The value of radio tracking in the study of neotropical rain-forest monkeys. American Journal of Primatology 32: 291301.Google Scholar
Campbell, C. J., Crofoot, M. C., MacKinnon, J. R., and Stumpf, R. 2011. Behavioral data collection in primate field studies. Pages 358367 in Primates in Perspective. Stumpf, R., Campbell, C. J., Fuentes, A., MacKinnon, J. R., and Bearder, S. K. (Eds.). Oxford University Press, Oxford.Google Scholar
Cargnelutti, B., Coulon, A., Hewison, A. J. M., et al. 2007. Testing global positioning system performance for wildlife monitoring using mobile collars and known reference points. Journal of Wildlife Management 71: 13801387.Google Scholar
Carpenter, C. R. 1934. A field study of the behavioral and social relations of howling monkeys (Alouatta palliata). Comparative Psychology Monographs 10: 1168.Google Scholar
Charles Dominique, P. 1977. Urine marking and territoriality in Galago alleni: field-study by radio-telemetry. Zeitschrift Fur Tierpsychologie – Journal of Comparative Ethology 43: 113138.Google Scholar
Cheney, D. L. 1987. Interactions and relations between groups. Pages 267281 in Primate Societies. Smuts, B. B., Cheney, D. L., Seyfarth, R. M., et al. (Eds.). University of Chicago Press, Chicago, IL.Google Scholar
Conradt, L. and Roper, T. J. 2003. Group decision-making in animals. Nature 421: 155158.Google Scholar
Conradt, L. and Roper, T. J. 2010. Deciding group movements: where and when to go. Behavioural Processes 84: 675677.Google Scholar
Conradt, L., Krause, J., Couzin, I. D., and Roper, T. J. 2009. Leading according to need in self-organizing groups. American Naturalist 173: 304312.Google Scholar
Couzin, I. D., Krause, J., Franks, N. R., and Levin, S. A. 2005. Effective leadership and decision-making in animal groups on the move. Nature 433: 513516.Google Scholar
Crofoot, M. C. 2007. Mating and feeding competition in white-faced capuchins (Cebus capucinus): the importance of short- and long-term strategies. Behaviour 144: 14731495.CrossRefGoogle Scholar
Crofoot, M. C. 2008. Intergroup Competition in White-Faced Capuchin Monkeys (Cebus capucinus): Automated Radio-Telemetry Reveals How Intergroup Relationships Shape Space-Use and Foraging Success. Harvard University, Cambridge, MA.Google Scholar
Crofoot, M. C. 2013. The cost of defeat: capuchin groups travel further, faster and later after losing conflicts with neighbors. American Journal of Physical Anthropology 152: 7985.CrossRefGoogle ScholarPubMed
Crofoot, M. C. and Wrangham, R. W. 2010. Intergroup aggression in primates and humans: the case for a unified theory. Pages 171195 in Mind the Gap. Kappeler, P. M. and Silk, J. (Eds.). Springer, New York.CrossRefGoogle Scholar
Crofoot, M. C., Gilby, I. C., Wikelski, M. C., and Kays, R. W. 2008. Interaction location outweighs the competitive advantage of numerical superiority in Cebus capucinus intergroup contests. Proceedings of the National Academy of Sciences of the United States of America 105: 577581.CrossRefGoogle ScholarPubMed
Crofoot, M. C., Lambert, T. D., Kays, R., and Wikelski, M. C. 2010. Does watching a monkey change its behaviour? Quantifying observer effects in habituated wild primates using automated radiotelemetry. Animal Behaviour 80: 475480.Google Scholar
Crofoot, M. C., Kays, R. W., and Wikelski, M. 2015. Shared decision-making drives collective movement in wild baboons. Movebank data repository.Google Scholar
Davies, A., Radford, A., and Nicol, C. 2014. Behavioural and physiological expression of arousal during decision-making in laying hens. Physiology & Behavior 123: 9399.Google Scholar
D’Eon, R. G. and Delparte, D. 2005. Effects of radio-collar position and orientation on GPS radio-collar performance, and the implications of PDOP in data screening. Journal of Applied Ecology 42: 383388.CrossRefGoogle Scholar
D’Eon, R. G., Serrouya, R., Smith, G., and Kochanny, C. O. 2002. GPS radiotelemetry error and bias in mountainous terrain. Wildlife Society Bulletin 30: 430439.Google Scholar
Fairbanks, L. A. and Pereira, M. E. 2002. Juvenile primates: dimensions for future research. Pages 359366 in Juvenile Primates: Life History, Development and Behavior. Pereira, M. E. and Fairbanks, L. A. (Eds.). University of Chicago Press, Chicago, IL.Google Scholar
Farine, D. R., Strandburg-Peshkin, A., Berger-Wolf, T., et al. 2016. Both nearest neighbours and long-term affiliates predict individual locations during collective movement in wild baboons. Scientific Reports 6: 27704.Google Scholar
Flack, A., Ákos, Z., Nagy, M., Vicsek, T., and Biro, D. 2013. Robustness of flight leadership relations in pigeons. Animal Behaviour 86: 723732.Google Scholar
Flack, A., Fiedler, W., Blas, J., et al. 2016. Costs of migratory decisions: a comparison across eight white stork populations. Science Advances 2: e1500931.CrossRefGoogle ScholarPubMed
Frair, J. L., Fieberg, J., Hebblewhite, M., et al. 2010. Resolving issues of imprecise and habitat-biased locations in ecological analyses using GPS telemetry data. Philosophical Transactions of the Royal Society B: Biological Sciences 365: 21872200.Google Scholar
Fuentes, A., Klegarth, A., Jones-Engel, L., et al. 2014. “Seeing the world through their eyes”: analyses of the first National Geographic Crittercam (TM) deployments on macaques in Singapore and Gibraltar. American Journal of Physical Anthropology 153: 122.Google Scholar
Gau, R. J., Mulders, R., Ciarniello, L. J., et al. 2004. Uncontrolled field performance of Televilt GPS-Simplex (TM) collars on grizzly bears in western and northern Canada. Wildlife Society Bulletin 32: 693701.Google Scholar
Gleiss, A. C., Wilson, R. P., and Shepard, E. L. C. 2011. Making overall dynamic body acceleration work: on the theory of acceleration as a proxy for energy expenditure. Methods in Ecology and Evolution 2: 2333.Google Scholar
Gursky, S. 2005. Associations between adult spectral tarsiers. American Journal of Physical Anthropology 128: 7483.Google Scholar
Halsey, L. G., Green, J. A., Wilson, R. P., and Frappell, P. B. 2009a. Accelerometry to estimate energy expenditure during activity: best practice with data loggers. Physiological and Biochemical Zoology 82: 396404.Google Scholar
Halsey, L. G., Shepard, E. L. C., Quintana, F., et al. 2009b. The relationship between oxygen consumption and body acceleration in a range of species. Comparative Biochemistry and Physiology A: Molecular & Integrative Physiology 152: 197202.Google Scholar
Hansen, M. C. and Riggs, R. A. 2008. Accuracy, precision, and observation rates of global positioning system telemetry collars. Journal of Wildlife Management 72: 518526.Google Scholar
Hebblewhite, M. and Haydon, D. T. 2010. Distinguishing technology from biology: a critical review of the use of GPS telemetry data in ecology. Philosophical Transactions of the Royal Society B: Biological Sciences 365: 23032312.Google Scholar
Hebblewhite, M., Percy, M., and Merrill, E. H. 2007. Are all global positioning system collars created equal? Correcting habitat-induced bias using three brands in the Central Canadian Rockies. Journal of Wildlife Management 71: 20262033.Google Scholar
Hulbert, I. A. R. and French, J. 2001. The accuracy of GPS for wildlife telemetry and habitat mapping. Journal of Applied Ecology 38: 869878.Google Scholar
Isbell, L. A. 1994. Predation on primates: ecological patterns and evolutionary consequences. Evolutionary Anthropology: Issues, News, and Reviews 3: 6171.Google Scholar
Isbell, L. A. and Young, T. P. 1993. Human presence reduces predation in a free-ranging vervet monkey population in Kenya. Animal Behaviour 45: 12331235.Google Scholar
Jachowski, D. S., Slotow, R., and Millspaugh, J. J. 2014. Good virtual fences make good neighbors: opportunities for conservation. Animal Conservation 17: 187196.Google Scholar
Jolly, A. 1966. Lemur Behavior: A Madagascar Field Study. University of Chicago Press, Chicago, IL.Google Scholar
Juarez, C. P., Rotundo, M. A., Berg, W., and Fernandez-Duque, E. 2011. Costs and benefits of radio-collaring on the behavior, demography, and conservation of owl monkeys (Aotus azarai) in Formosa, Argentina. International Journal of Primatology 32: 6982.Google Scholar
Kaplan, E. D. and Hegarty, C. J. 2005. Understanding GPS: Principles and Applications. Artech House, London.Google Scholar
Kappeler, P. 1997. Intrasexual selection in Mirza coquereli: evidence for scramble competition polygyny in a solitary primate. Behavioral Ecology and Sociobiology 41: 115127.Google Scholar
Kappeler, P., Barrett, L., Blumstein, D. T., and Clutton-Brock, T. H. E. 2013. Flexibility and constraint in the evolution of mammalian social behaviour. Philosophical Transactions of the Royal Society B: Biological Sciences 368: 20120337.Google Scholar
Kays, R., Crofoot, M. C., Jetz, W., and Wikelski, M. 2015. Terrestrial animal tracking as an eye on life and planet. Science 348: DOI: 10.1126/science.aaa2478.Google Scholar
King, A. J., Wilson, A. M., Wilshin, S. D., et al. 2012. Selfish-herd behaviour of sheep under threat. Current Biology 22: R561R562.Google Scholar
Kranstauber, B., Kays, R., LaPoint, S. D., Wikelski, M., and Safi, K. 2012. A dynamic Brownian bridge movement model to estimate utilization distributions for heterogeneous animal movement. Journal of Animal Ecology 81: 738746.Google Scholar
Krause, J., Krause, S., Arlinghaus, R., et al. 2013. Reality mining of animal social systems. Trends in Ecology & Evolution 28: 541551.Google Scholar
Lesku, J. A., Rattenborg, N. C., Valcu, M., et al. 2012. You snooze, you lose: adaptive sleep loss in polygynous pectoral sandpipers. Journal of Sleep Research 21: 5.Google Scholar
Lynch, E., Angeloni, L., Fristrup, K., Joyce, D., and Wittemyer, G. 2013. The use of on-animal acoustical recording devices for studying animal behavior. Ecology and Evolution 3: 20302037.Google Scholar
Macdonald, D. W. and Amlaner, C. J. 1980. A practical guide to radio tracking. Pages 143159 in A Handbook on Biotelemetry and Radio Tracking. Amlaner, C. J. and MacDonald, D. W. (Eds.). Pergamon Press, Oxford.Google Scholar
Manson, J. H. and Wrangham, R. W. 1991. Intergroup aggression in chimpanzees and humans. Current Anthropology 32: 369390.Google Scholar
Markham, A. C. and Altmann, J. 2008. Remote monitoring of primates using automated GPS technology in open habitats. American Journal of Primatology 70: 15.Google Scholar
Markham, A. C., Alberts, S. C., and Altmann, J. 2012. Intergroup conflict: ecological predictors of winning and consequences of defeat in a wild primate population. Animal Behaviour 84: 399403.Google Scholar
Markham, A. C., Guttal, V., Alberts, S. C., and Altmann, J. 2013. When good neighbors don’t need fences: temporal landscape partitioning among baboon social groups. Behavioral Ecology and Sociobiology 67: 875884.Google Scholar
Merker, S. 2006. Habitat-specific ranging patterns of Dian’s tarsiers (Tarsius dianae) as revealed by radiotracking. American Journal of Primatology 68: 111125.Google Scholar
Moil, R. J., Millspaugh, J. J., Beringer, J., Sartwell, J., and He, Z. 2007. A new “view” of ecology and conservation through animal-borne video systems. Trends in Ecology & Evolution 22: 660668.CrossRefGoogle Scholar
Nagy, M., Akos, Z., Biro, D., and Vicsek, T. 2010. Hierarchical group dynamics in pigeon flocks. Nature 464: 890899.Google Scholar
Nagy, M., Vásárhelyi, G., Pettit, B., et al. 2013. Context-dependent hierarchies in pigeons. Proceedings of the National Academy of Sciences 110: 1304913054.Google Scholar
Nathan, R., Getz, W. M., Revilla, E., et al. 2008. A movement ecology paradigm for unifying organismal movement research. Proceedings of the National Academy of Sciences of the United States of America 105: 1905219059.Google Scholar
Nathan, R., Spiegel, O., Fortmann-Roe, S., et al. 2012. Using tri-axial acceleration data to identify behavioral modes of free-ranging animals: general concepts and tools illustrated for griffon vultures. Journal of Experimental Biology 215: 986996.Google Scholar
Newmaster, S. G., Thompson, I. D., Steeves, R. A., et al. 2013. Examination of two new technologies to assess the diet of woodland caribou: video recorders attached to collars and DNA barcoding. Canadian Journal of Forest Research 43: 897900.Google Scholar
Nowak, K., le Roux, A., Richards, S. A., Scheijen, C. P. J., and Hill, R. A. 2014. Human observers impact habituated samango monkeys’ perceived landscape of fear. Behavioral Ecology 25. DOI: 10.1093/beheco/aru110.Google Scholar
Perez-Escudero, A. and de Polavieja, G. G. 2011. Collective animal behavior from Bayesian estimation and probability matching. PLoS Computational Biology 7. DOI: 10.1371/journal.pcbi.1002282.Google Scholar
Pettit, B., Perna, A., Biro, D., and Sumpter, D. J. T. 2013. Interaction rules underlying group decisions in homing pigeons. Journal of the Royal Society Interface 10. DOI: 10.1098/rsif.2013.0529.Google Scholar
Phillips, K. A., Elvey, C. R., and Abercrombie, C. L. 1998. Applying GPS to the study of primate ecology: A useful tool? American Journal of Primatology 46: 167172.Google Scholar
Qasem, L., Cardew, A., Wilson, A., et al. 2012. Tri-axial dynamic acceleration as a proxy for animal energy expenditure: should we be summing values or calculating the vector? PLoS ONE 7. DOI: 10.1371/journal.pone.0031187.Google Scholar
Rasmussen, D. R. 1991. Observer influence on range use of Macaca arctoides after 14 years of observation? Laboratory Primate Newsletter 30: 611.Google Scholar
Rattenborg, N. C., Voirin, B., Vyssotski, A. L., et al. 2008. Sleeping outside the box: electroencephalographic measures of sleep in sloths inhabiting a rainforest. Biology Letters 4: 402405.Google Scholar
Recio, M. R., Mathieu, R., Denys, P., Sirguey, P., and Seddon, P. J. 2011. Lightweight GPS-tags, one giant leap for wildlife tracking? An assessment approach. PLoS ONE 6. DOI: 10.1371/journal.pone.0028225.Google Scholar
Rempel, R. S., Rodgers, A. R., and Abraham, K. F. 1995. Performance of a GPS animal location system under boreal forest canopy. Journal of Wildlife Management 59: 543551.Google Scholar
Ren, B., Li, M., Long, Y., Grüter, C. C., and Wei, F. 2008. Measuring daily ranging distances of Rhinopithecus bieti via a global positioning system collar at Jinsichang, China: a methodological consideration. International Journal of Primatology 29(3): 783.Google Scholar
Rodgers, A. R. 2001. Recent telemetry technology. Pages 82121 in Radio Tracking and Animal Populations. Millspaugh, J. J. and Marzluff, J. M. (Eds.). Academic Press, San Diego, CA.Google Scholar
Rodman, P. S. and Mitani, J. C. 1987. Orangutans: sexual dimorphism in a solitary species. Pages 146154 in Primate Societies. Smuts, B., Cheney, D. L., Seyfarth, R. M., Struhsaker, T. T., and Wrangham, R. (Eds.). Chicago University Press, Chicago, IL.Google Scholar
Rutz, C. and Troscianko, J. 2013. Programmable, miniature video-loggers for deployment on wild birds and other wildlife. Methods in Ecology and Evolution 4: 114122.Google Scholar
Sapir, N., Wikelski, M., McCue, M. D., Pinshow, B., and Nathan, R. 2010. Flight modes in migrating European bee-eaters: heart rate may indicate low metabolic rate during soaring and gliding. PLoS ONE 5: e13956.Google Scholar
Sapir, N., Rotics, S., Kaatz, M., et al. 2013. Multi-year tracking of white storks (Ciconia ciconia): how the environment shapes the movement and behavior of a soaring-gliding inter-continental migrant. Integrative and Comparative Biology 53: E189E189.Google Scholar
Shepard, E. L. C., Wilson, R. P., Halsey, L. G., et al. 2009. Derivation of body motion via appropriate smoothing of acceleration data. Aquatic Biology 4: 235241.Google Scholar
Shumaker, R. 2007. Orangutans. Voyager Press, St. Paul, MI.Google Scholar
Sih, A. 2013. Understanding variation in behavioural responses to human-induced rapid environmental change: a conceptual overview. Animal Behaviour 85: 10771088.Google Scholar
Singleton, I. and van Schaik, C. P. 2001. Orangutan home range size and its determinants in a Sumatran swamp forest. International Journal of Primatology 22: 877911.Google Scholar
Spiegel, O., Getz, W. M., and Nathan, R. 2013. Factors influencing foraging search efficiency: why do scarce lappet-faced vultures outperform ubiquitous white-backed vultures? The American Naturalist 181: E102E115.Google Scholar
Sprague, D. 2004. GPS collars for monkeys: the state of the technology. American Journal of Physical Anthropology 186: 151154.Google Scholar
Sprague, D. S., Kabaya, M., and Hagihara, K. 2004. Field testing a global positioning system (GPS) collar on a Japanese monkey: reliability of automatic GPS positioning in a Japanese forest. Primates 45: 151154.Google Scholar
Strandburg-Peshkin, A., Twomey, C. R., Bode, N. W. F., et al. 2013. Visual sensory networks and effective information transfer in animal groups. Current Biology 23: R709R711.Google Scholar
Strandburg-Peshkin, A., Farine, D. R., Couzin, I. D., and Crofoot, M. C. 2015. Shared decision-making drives collective movement in wild baboons. Science 348: 13581361.Google Scholar
Thorington, R. W., Muckenhirn, N. A., and Montgomery, G. G. 1976. Movements of a wild night monkey (Aotus trivirgatus). Pages 3234 in Neotropical Primates. Thorington, R. W. and Heltne, P. G. (Eds.). National Academy of Sciences, Washington, DC.Google Scholar
Tomkiewicz, S. M., Fuller, M. R., Kie, J. G., and Bates, K. K. 2010. Global positioning system and associated technologies in animal behaviour and ecological research. Philosophical Transactions of the Royal Society B: Biological Sciences 365: 21632176.Google Scholar
Tomlinson, S., Arnall, S. G., Munn, A., et al. 2014. Applications and implications of ecological energetics. Trends in Ecology & Evolution 29: 280290.Google Scholar
van Schaik, C. P. 1983. On the ultimate causes of primate social systems. Behaviour 85: 91117.Google Scholar
Voirin, B., Scriba, M. F., Martinez-Gonzalez, D., et al. 2014. Ecology and neurophysiology of sleep in two wild sloth species. Sleep 37: 753.Google Scholar
Wall, J., Wittemyer, G., Klinkenberg, B., and Douglas-Hamilton, I. 2014. Novel opportunities for wildlife conservation and research with real-time monitoring. Ecological Applications 24: 593601.Google Scholar
Watts, I., Nagy, M., Biro, T. B., and de Perera, D. 2016. Misinformed leaders lose influence over pigeon flocks. Biology Letters 12: 20160544.Google Scholar
White, E. C., Dikangadissi, J. T., Dimoto, E., et al. 2010. Home-range use by a large horde of wild Mandrillus sphinx. International Journal of Primatology 31: 627645.Google Scholar
Wikelski, M. and Kays, R. 2011. Movebank: archive, analysis and sharing of animal movement data. Available at: www.movebank.org.Google Scholar
Wikelski, M., Kays, R. W., Kasdin, N. J., et al. 2007. Going wild: what a global small-animal tracking system could do for experimental biologists. Journal of Experimental Biology 210: 181186.Google Scholar
Wilcove, D. S. and Wikelski, M. 2008. Going, going, gone: is animal migration disappearing. PLOS Biology 6: e188.Google Scholar
Williams, D. M., Quinn, A. D., and Porter, W. F. 2012. Impact of habitat-specific GPS positional error on detection of movement scales by first-passage time analysis. PLoS ONE 7. DOI: 10.1371/journal.pone.0048439.Google Scholar
Wilson, A. M., Lowe, J. C., Roskilly, K., et al. 2013. Locomotion dynamics of hunting in wild cheetahs. Nature 498: 185189.Google Scholar
Wilson, R. P., White, C. R., Quintana, F., et al. 2006. Moving towards acceleration for estimates of activity-specific metabolic rate in free-living animals: the case of the cormorant. Journal of Animal Ecology 75: 10811090.Google Scholar
Wilson, R. P., Shepard, E., and Liebsch, N. 2008. Prying into the intimate details of animal lives: use of a daily diary on animals. Endangered Species Research 4: 123137.Google Scholar
Yoda, K., Murakoshi, M., Tsutsui, K., and Kohno, H. 2011. Social interactions of juvenile brown boobies at sea as observed with animal-borne video cameras. PLoS ONE 6. DOI: 10.1371/journal.pone.0019602.CrossRefGoogle ScholarPubMed
Zinner, D., Hindahl, J., and Kaumanns, W. 2001. Experimental intergroup encounters in lion-tailed macaques (Macaca silenus). Primate Report 59: 7792.Google Scholar

References

American Society of Mammalogists. 1998. Guidelines for the capture, handling, and care of mammals as approved by the American Society of Mammalogists. Journal of Mammalogy 74: 14161431.Google Scholar
Côté, S. D., Festa-Bianchet, M., and Fournier, F. 1998. Life-history effects of chemical immobilization and radio collars on mountain goats. Journal of Wildlife Management 62: 745752.Google Scholar
Crofoot, M. C., Norton, T. M., Lessnau, R. C., et al. 2009. Field anesthesia and health assessment of free ranging white-faced capuchin monkeys (Cebus capucinus) in Panama. International Journal of Primatology 30: 125141.Google Scholar
Crofoot, M. C., Lambert, T. D., Kays, R., and Wikelski, M. R. 2010. Does watching a monkey change its behaviour? Quantifying observer effects in habituated wild primates using automated radiotelemetry. Animal Behaviour 80: 475480Google Scholar
Crofoot, M. C., Kays, R., Alavi, S., and Wikelski, M. 2014. Democracy or despotism? How do baboons decide? American Journal of Physical Anthropology 153: 99100.Google Scholar
Di Fiore, A. and Link, A. 2013. Evaluating the utility of GPS collars for studies of ranging by large-bodied, arboreal, forest-dwelling primates. American Journal of Physical Anthropology S56: 112.Google Scholar
ESRI. 2014. ArcGIS Desktop: Release 10. Environmental Systems Research Institute, Redlands, CA.Google Scholar
Fuentes, A., Kalchik, S., Gettler, L., et al. 2008. Characterizing human–macaque interactions in Singapore. American Journal of Primatology 70: 15.Google Scholar
Gursky, S. 1998. Effects of radio transmitter weight on a small nocturnal primate. American Journal of Primatology 46: 145155.Google Scholar
Hebblewhite, M. and Haydon, D. T. 2010. Distinguishing technology from biology: a critical review of the use of GPS telemetry data in ecology. Philosophical Transactions of the Royal Society of London B: Biological Sciences 365: 23032312.Google Scholar
Hilpert, A. L. and Jones, C. B. 2005. Possible costs of radio-tracking a young adult female mantled howler monkey (Alouatta palliata) in deciduous habitat of Costa Rican tropical dry forest. Journal of Applied Animal Welfare Science 8: 227232.Google Scholar
Jones-Engel, L. and Engel, G. A. 2009. The risks and contexts of emerging primate-borne zoonoses. Pages 5277 in Health, Risk and Adversity. Fuentes, A. and Panter-Brick, C. (Eds.). Berghahn Books, Oxford.Google Scholar
Jones-Engel, L., Engel, G. A., Schillaci, M., et al. 2005. Primate to human retroviral transmission in Asia. Emerging Infectious Diseases 11: 10281035.Google Scholar
Klegarth, A. R. (2017). Measuring movement: how remote telemetry facilitates our understanding of the human–macaque interface. Pages 7087 in Ethnoprimatology: A Practical Guide to Research on the Human–Nonhuman Primate Interface. Dore, K., Riley, E., and Fuentes, A. (Eds.). Cambridge University Press, Cambridge.Google Scholar
Klegarth, A. R., Hollocher, H., Jones-Engel, L., et al. (2017). Urban primate ranging patterns: GPS-collar deployments for Macaca fascicularis and M. sylvanus. American Journal of Primatology 79. DOI: 10.1002/ajp.22633.Google Scholar
Langley, R. B. 1999. Dilution of precision. GPS World. Available at: www.gpsworld.com.Google Scholar
MacKinnon, K. C. and Riley, E. P. 2013. Ethical issues in field primatology. Pages 98107 in Ethics in the Field: Contemporary Challenges. Fuentes, A. and MacClancy, J. (Eds.). Berghahn Books, New York.Google Scholar
Markham, A. C. and Altmann, J. 2008. Remote monitoring of primates using automated GPS technology in open habitats. American Journal of Primatology 70: 15.Google Scholar
Markham, A. C., Guttal, V., Alberts, S. C., and Altmann, J. 2013. When good neighbors don’t need fences: temporal landscape partitioning among baboon social groups. Behavioral Ecology and Sociobiology 67: 875884.Google Scholar
Matthews, A., Ruykys, L., Ellis, B., et al. 2013. The success of GPS collar deployments on mammals in Australia. Australian Mammalogy 35: 6583.Google Scholar
Moorhouse, T. P. and MacDonald, D. W. J. 2005. Indirect negative impacts of radio-collaring: sex ratio variation in water voles. Journal of Applied Ecology 42: 9198.Google Scholar
Morton, D. B., Hawkins, P., Bevan, R., et al. 2003. Refinements in telemetry procedures: Seventh Report of the BVAAWF/FRAME/RSPCA/UFAW Joint Working Group on Refinement, Part A. Laboratory Animals 37, 261299.Google Scholar
Nash, L. T. 2005. Studies of primates in the field and in captivity: similarities and differences in ethical concerns. Pages 2748 in Biological Anthropology and Ethics: From Repatriation to Genetic Identity. Turner, T.R. (Ed.). SUNY Press, Albany, NY.Google Scholar
Oberste, M. S., Feeroz, M. M., Maher, K., et al. 2013. Characterizing the picornavirus landscape among synanthropic nonhuman primates in Bangladesh, 2007–2008. Journal of Virology 87: 558571.Google Scholar
Setchell, J.M. and Curtis, D.J. (Eds.). 2010. Field and Laboratory Methods in Primatology: A Practical Guide. Cambridge University Press, Cambridge.Google Scholar
Sprague, D. S., Kabaya, H., and Hagihara, K. 2004. Field testing a global positioning systems (GPS) collar on a Japanese monkey: reliability of automatic GPS positioning in a Japanese forest. Primates 45: 151154.Google Scholar
Todd, D. E. and Shah, S. F. 2012. Assessing acute effects of trapping, handling, and tagging on the behavior of wildlife using GPS telemetry: a case study of the common brushtail possum. Journal of Applied Animal Welfare Science 15: 189207.Google Scholar
Tuyttens, F. A. M., MacDonald, D. W. J., and Roddam, A. W. 2002. Effects of radio-collars on European badgers (Meles meles). Journal of Zoology 257: 3742.Google Scholar

References

Araldi, A., Barelli, C., Hodges, K., and Rovero, F. 2014. Density estimation of the endangered Udzungwa red colobus (Procolobus gordonorum) and other arboreal primates in the Udzungwa Mountains using systematic distance sampling. International Journal of Primatology 35(5): 941956.Google Scholar
Bivand, R., Keitt, T., and Rowlingson, B. 2016. rgdal: Bindings for the Geospatial Data Abstraction Library. R package version 1.2-4.Google Scholar
Burgman, M. A. and Fox, J. C. 2003. Bias in species range estimates from minimum convex polygons: implications for conservation and options for improved planning. Animal Conservation 6(1): 1928.Google Scholar
Calenge, C. 2006. The package adehabitat for the R software: a tool for the analysis of space and habitat use by animals. Ecological Modelling 197: 516519.Google Scholar
Coleman, B. T. and Hill, R. A. 2014. Living in a landscape of fear: the impact of predation, resource availability and habitat structure on primate range use. Animal Behaviour 88: 165173.Google Scholar
Crawley, M. J. 2007. The R Book. Wiley, Chichester.Google Scholar
Dalgaard, P. 2002. Introductory Statistics with R. Springer, New York.Google Scholar
Fleming, C. and Calabrese, J. 2017. ctmm: Continuous-Time Movement Modeling. R package version 0.3.5. Available at: https://CRAN.R-project.org/package=ctmm.Google Scholar
Fleming, C. H., Fagan, W. F., Mueller, T., et al. 2016. Estimating where and how animals travel: an optimal framework for path reconstruction from autocorrelated tracking data. Ecology. DOI: 10.1890/15-1607.Google Scholar
Getz, W. M. and Wilmers, C. C. 2004. A local nearest-neighbor convex-hull construction of home ranges and utilization distributions. Ecography 27: 489505.Google Scholar
Hengl, T., Roudier, P., Beaudette, D., and Pebesma, E. 2015. plotKML: scientific visualization of spatio-temporal data. Journal of Statistical Software 63(5): 125.Google Scholar
Hicks, T. C., Tranquilli, S., Kuehl, H., et al. 2014. Absence of evidence is not evidence of absence: discovery of a large, continuous population of Pan troglodytes schweinfurthii in the Central Uele region of northern DRC. Biological Conservation 171: 107113.Google Scholar
Hijmans, R. 2016a. geosphere: spherical trigonometry. R package version 1.5-5.Google Scholar
Hijmans, R. 2016b. raster: geographic data analysis and modeling. R package version 2.5-8.Google Scholar
Howard, A. M., Nibbelink, N. P., Madden, M., et al. 2015. Landscape influences on the natural and artificially manipulated movements of bearded capuchin monkeys. Animal Behaviour 106: 5970.Google Scholar
Janmaat, K. R. L., Ban, S. D., and Boesch, C. 2013. Chimpanzees use long-term spatial memory to monitor large fruit trees and remember feeding experiences across seasons. Animal Behaviour 86: 11831205.Google Scholar
Lo, C. P. and Yeung, A. K. 2007. Concepts and techniques of geographic information systems. Pearson Prentice Hall, Upper Saddle River, NJ.Google Scholar
Mohr, C. 1947. Table of equivalent populations of North American small mammals. American Midland Naturalist 37: 223249.Google Scholar
Patzelt, A., Kopp, G. H., Ndaod, I., et al. 2014. Male tolerance and male–male bonds in a multilevel primate society. PNAS 41: 1474014745.Google Scholar
Qi, X. G., Garber, P. A., Ji, W., et al. 2014. Satellite telemetry and social modeling offer new insights into the origin of primate multilevel societies. Nature 5(5296): DOI: 10.1038/ncomms6296.Google Scholar
R Core Team. 2016. R: A Language and Environment for Statistical Computing. R Foundation for Statistical Computing, Vienna.Google Scholar
R packages. 2015. Home page. Available at: http://r-pkgs.had.co.nz.Google Scholar
Rowlingson, B. and Diggle, P. 2016. splancs: spatial and space-time point pattern analysis. R package version 2.01-39. Available at: https://CRAN.R-project.org/package=splancs.Google Scholar
Seaman, D. E. and Powell, R. A. 1996. An evaluation of the accuracy of kernel density estimators for home range analysis. Ecology 77 (7): 20752085.Google Scholar
Springer, A., Mellmann, A., Fichtel, C., and Kappeler, P. M. 2016. Social structure and Escherichia coli sharing in a group-living wild primate, Verreaux’s sifaka. BMC Ecology 16(1): 6.Google Scholar
Szantoi, Z., Smith, S. E., Strona, G., Koh, L. P., and Wich, S. A. 2017. Mapping orangutan habitat and agricultural areas using Landsat OLI imagery augmented with unmanned aircraft system aerial photography. International Journal of Remote Sensing 38: 22312245.Google Scholar
Teetor, P. 2011. R Cookbook: Proven Recipes for Data Analysis, Statistics, and Graphics. O’Reilly Media, Inc., Beijing.Google Scholar

References

Asensio, N., Brockelman, W. Y., Malaivijitnond, , Reichard, S., U. H. 2011. Gibbon travel paths are goal oriented. Animal Cognition 14: 395405.Google Scholar
Byrne, R. W., Noser, R. N., Bates, L. A., and Jupp, P. E. 2009. How did they get here from there? Detecting changes of direction in terrestrial ranging. Animal Behaviour 77: 619631.Google Scholar
Coyne, M. S. and Godley, B. J. 2005. Satellite tracking and analysis tool (STAT): an integrated system for archiving, analyzing and mapping animal track data. Marine Ecology Progress Series 301: 17.Google Scholar
Guschanski, K., Vigilant, L., McNeilage, A., et al. 2009. Counting elusive animals: comparison of a field and genetic census of the entire population of mountain gorillas of Bwindi Impenetrable National Park, Uganda. Biological Conservation 142: 290300.Google Scholar
Hickey, J. R., Carroll, J. P., and Nibbelink, N. P. 2012. Applying landscape metrics to characterize potential habitat of bonobos (Pan paniscus) in the Maringa-Lopori-Wamba landscape, Democratic Republic of Congo. International Journal of Primatology 33: 381400.Google Scholar
Janmaat, K. R. L., Byrne, R. W., and Zuberbühler, K. 2006. Evidence for spatial memory of fruiting states of rainforest fruit in wild ranging mangabeys. Animal Behaviour 71: 797807.Google Scholar
Junker, J., Blake, S., Boesch, C., et al. 2012. Recent decline in suitable environmental conditions for African great apes. Diversity and Distribution 18: 10771091.Google Scholar
Kouakou, C. Y., Boesch, C., and Kühl, H. 2009. Estimating chimpanzee population size with nest counts: validating methods in Taï National Park. American Journal of Primatology 71(6): 7176.Google Scholar
Lehmann, J. and Boesch, C. 2005. Bisexually-bonded ranging in chimpanzees (Pan troglodytes verus). Behavioral Ecology and Sociobiology 57: 525535.Google Scholar
Milton, K. 2000. Quo Vadis? Tactics of food search and group movement in primates and other animals. Pages 375417 in On the Move: How and Why Animals Travel in Groups. Boinski, S., and Garber, P. A. (Eds.). University of Chicago Press, Chicago, IL.Google Scholar
Noser, R. and Byrne, R. W. 2007. Travel routes and planning of visits to out-of-sight resources in wild chacma baboons (Papio ursinus). Animal Behaviour 73: 257266.Google Scholar
Noser, R. and Byrne, R. W. 2009. How do wild baboons (Papio ursinus) plan their routes? Travel among multiple high-quality food sources with inter-group competition. Animal Cognition 13: 145155.Google Scholar
Olupot, W., Chapman, C. A., Waser, P. M., and Isabirye-Basuta, G. 1997. Mangabey (Cercocebus albigena) ranging patterns in relation to fruit availability and the risk of parasite infection in Kibale National Park, Uganda. American Journal of Primatology 43: 6578.Google Scholar
Patterson, T. A., Thomas, L., Wilcox, C., Ovaskainen, O., and Matthiopoulos, J. 2008. State-space models of individual animal movement. Trends in Ecology and Evolution 23: 8794.Google Scholar
Pochron, S. 2001. Can concurrent speed and directness of travel indicate purposeful encounter in the yellow baboons (Papio hamadryas cynocephalus) of Ruaha National Park, Tanzania? International Journal of Primatology 22(5): 773785.Google Scholar
R Development Core Team. 2012. R: A Language and Environment for Statistical Computing. R Foundation for Statistical Computing, Vienna.Google Scholar
Steudel, K. 2000. The physiology and energetics of movement effects on individual and groups. Pages 923 in On the Move: How and Why Animals Travel in Groups. Boinski, S. and Garber, P. A. (Eds.). University of Chicago Press, Chicago, IL.Google Scholar
Tremblay, Y., Robinson, P. W., and Costa, D. P. 2009. A parsimonious approach to modelling animal movement data. PLoS ONE 4(3): 4711.Google Scholar
Valero, A. and Byrne, R. W. 2007. Spider monkey ranging patterns in Mexican subtropical forest: do travel routes reflect planning? Animal Cognition 10: 305315.Google Scholar
Walsh, P. D., Biek, R., and Real, L. A. 2005. Wave-like spread of ebola Zaire. PLoS Biology 3(11): e371.Google Scholar

Save book to Kindle

To save this book to your Kindle, first ensure coreplatform@cambridge.org is added to your Approved Personal Document E-mail List under your Personal Document Settings on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part of your Kindle email address below. Find out more about saving to your Kindle.

Note you can select to save to either the @free.kindle.com or @kindle.com variations. ‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi. ‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.

Find out more about the Kindle Personal Document Service.

Available formats
×

Save book to Dropbox

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Dropbox.

Available formats
×

Save book to Google Drive

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Google Drive.

Available formats
×