Skip to main content Accessibility help
×
Hostname: page-component-8448b6f56d-m8qmq Total loading time: 0 Render date: 2024-04-19T22:22:49.524Z Has data issue: false hasContentIssue false

Part II - GIS Analysis in Fine-Scale Space

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. 121 - 306
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

Akers, A. A., Islam, M. A., and Nijman, V. 2013. Habitat characterization of western hoolock gibbons Hoolock hoolock by examining home range microhabitat use. Primates 54: 341348.Google Scholar
Anderson, D. J. 1982. The home range: a new nonparametric estimation technique. Ecology 63: 103112.Google Scholar
Bartlett, T. Q. 2007. The Hylobatidae: small apes of Asia. Pages 274289 in Primates in Perspective. Campbell, C. J., Fuentes, A., MacKinnon, K. C., Panger, M., and Bearder, S. (Eds.). Oxford University Press, Oxford.Google Scholar
Bearder, S. 1987. Lorises, bushbabies, and tarsiers: diverse societies in solitary foragers. Pages 1124 in Primate Societies. Smuts, B. B., Cheney, D. L., Seyfarth, R. M., Wrangham, R. W., and Struhsaker, T. T. (Eds.). University of Chicago Press, Chicago, IL.Google Scholar
Beyer, H. L. 2010. Geospatial modeling environment for ArcGIS. Available at: www.spatialecology.com/gme.Google Scholar
Biebouw, K. 2009. Home range size and use in Allocebus trichotis in Analamazaotra Special Reserve, Central Eastern Madagascar. International Journal of Primatology 30: 367386.Google Scholar
Blair, W. F. 1942. Size of home range and notes on the life history of the woodland deer-mouse and eastern chipmunk in northern Michigan. Journal of Mammalogy 23: 2736.Google Scholar
Blundell, F. M., Maier, J. A. K., and Debevec, E. M. 2001. Linear home ranges: effects of smoothing, sample size, and autocorrelation on kernel estimates. Ecological Monographs 71: 469489.Google Scholar
Börger, L., Franconi, N., De Michelle, G., et al. 2006. Effects of sampling regime on the mean and variance of home range size estimates. Journal of Animal Ecology 75: 13931405.CrossRefGoogle ScholarPubMed
Boyle, S. A. and Smith, A. T. 2010a. Behavioral modifications in northern bearded saki monkeys (Chiropotes satanas chiropotes) in forest fragments of central Amazonia. Primates 51: 4351.CrossRefGoogle ScholarPubMed
Boyle, S. A. and Smith, A. T. 2010b. Can landscape and species characteristics predict primate presence in forest fragments in the Brazilian Amazon? Biological Conservation 143: 11341143.Google Scholar
Boyle, S. A., Lourenço, W. C., da Silva, L. R., and Smith, A. T. 2009a. Travel and spatial patterns change when northern bearded saki monkeys (Chiropotes satanas chiropotes) live in forest fragments. International Journal of Primatology 30: 515531.CrossRefGoogle Scholar
Boyle, S. A., Lourenço, W. C., da Silva, L. R., and Smith, A. T. 2009b. Home range estimates vary with sample size and methods. Folia Primatologica 80: 3342.Google Scholar
Burt, W. H. 1940. Territorial behavior and populations of small mammals in southern Michigan. Miscellaneous Publications Museum of Zoology, University of Michigan 45: 158.Google Scholar
Burt, W. H. 1943. Territoriality and home range concepts as applied to mammals. Journal of Mammalogy 24: 346352.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
Campbell-Smith, G., Campbell-Smith, M., Singleton, I., and Linkie, M. 2011. Apes in space: saving an imperiled orangutan population in Sumatra. PLoS ONE 6: e17210.CrossRefGoogle ScholarPubMed
Campos, F. A. and Fedigan, L. M. 2009. Behavioral adaptations to heat stress and water scarcity in white-faced capuchins (Cebus capucinus) in Santa Rosa National Park, Costa Rica. American Journal of Physical Anthropology 138: 101111.Google Scholar
Chapman, C. 1988. Patterns of foraging and range use by three species of Neotropical primates. Primates 29: 177194.Google Scholar
Chaves, O. M., Stoner, K. E., and Arroyo-Rodríguez, V. 2011. Seasonal differences in activity patterns of Geoffroyi’s spider monkeys (Ateles geoffroyi) living in continuous and fragmented forests in southern Mexico. International Journal of Primatology 32: 960973.Google Scholar
Cresswell, W. J. and Smith, G. C. 1992. The effects of temporally autocorrelated data on methods of home range analysis. Pages 272284 in Wildlife Telemetry: Remote Monitoring and Tracking of Animals. Priede, I. G. and Swift, S. M. (Eds.). Ellis Horwood, London.Google Scholar
Cumming, G. S. and Cornélis, D. 2012. Quantitative comparison and selection of home range metrics for telemetry data. Diversity and Distributions 18: 10571065.Google Scholar
Cushman, S. A., Chase, M., and Griffin, C. 2005. Elephants in space and time. Oikos 109: 331341.Google Scholar
de Solla, S. R., Bonduriansky, R., and Brooks, R. J. 1999. Eliminating autocorrelation reduces biological relevance of home range estimates. Journal of Animal Ecology 68: 221234.CrossRefGoogle Scholar
Di Fiore, A. and Campbell, C. J. 2007. The Atelines: variation in ecology, behavior, and social organization. Pages 155185 in Primates in Perspective. Campbell, C. J., Fuentes, A., MacKinnon, K. C., Panger, M., and Bearder, S. (Eds.). Oxford University Press, Oxford.Google Scholar
Digby, L. J., Ferrari, S. F., and Saltzman, W. 2007. Callitrichines: the role of competition in cooperatively breeding species. Pages 85106 in Primates in Perspective. Campbell, C. J., Fuentes, A., MacKinnon, K. C., Panger, M., and Bearder, S. (Eds.). Oxford University Press, Oxford.Google Scholar
Dixon, K. R. and Chapman, J. A. 1980. Harmonic mean measure of animal activity areas. Ecology 61: 10401044.Google Scholar
Ecological Software Solutions, LLC. 2013. Biotas. Software.Google Scholar
Enstam, K. L. and Isbell, L. A. (2007. The guenons (genus Cercopithecus) and their allies: behavioral ecology of polyspecific associations. Pages 252274 in Primates in Perspective. Campbell, C. J., Fuentes, A., MacKinnon, K. C., Panger, M., and Bearder, S. (Eds.). Oxford University Press, Oxford.Google Scholar
ESRI 2013. ArcGIS 10.X. Software.Google Scholar
Fashing, P. J. 2007. African colobine monkeys: patterns of between-group interaction. Pages 201224 in Primates in Perspective. Campbell, C. J., Fuentes, A., MacKinnon, K. C., Panger, M., and Bearder, S. (Eds.). Oxford University Press, Oxford.Google Scholar
Fashing, P. J., Mulindahabi, F., Gakima, J. B., et al. 2007. Activity and ranging patterns of Colobus angolensis ruwenzorii in Nyungwe Forest, Rwanda: possible costs of large group size. International Journal of Primatology 28: 529550.Google Scholar
Fedigan, L. M., Fedigan, L., Chapman, C., and Glander, K. E. 1988. Spider monkey home ranges: a comparison of radio telemetry and direct observation. American Journal of Primatology 16: 1929.Google Scholar
Fernandez-Duque, E. 2007. Aotinae: social monogramy in the only nocturnal haplorhines. Pages 139154 in Primates in Perspective. Campbell, C. J., Fuentes, A., MacKinnon, K. C., Panger, M., and Bearder, S. (Eds.). Oxford University Press, Oxford.Google Scholar
Fieberg, J. 2007a. Kernel density estimators of home range: smoothing and the autocorrelation red herring. Ecology 88: 10591066.Google Scholar
Fieberg, J. 2007b. Utilization distribution estimation using weighted kernel density estimators. The Journal of Wildlife Management 71: 16691675.Google Scholar
Fieberg, J. and Börger, L. B. 2012. Could you please phrase “home range” as a question? Journal of Mammalogy 93: 890902.Google Scholar
Garber, P. A., Pruetz, J. D., and Isaacson, J. 1993. Patterns of range use, range defense, and intergroup spacing in moustached tamarin monkeys (Saguinus mystax). Primates 34: 1125.Google Scholar
Gerber, B. D., Arrigo-Nelson, S., Karpanty, S. M., Kotschwar, M., and Wright, P. C. 2012. Spatial ecology of the endangered Milne-Edwards’ sifaka (Propithecus edwardsi): do logging and season affect home range and daily ranging patterns? International Journal of Primatology 33: 305321.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
Getz, W. M., Fortmann-Roe, S., Cross, P. C., et al. 2007. LoCoH: nonparametric kernel methods for constructing home range and utilization distributions. PLoS ONE 2: e207.Google Scholar
Girard, I., Ouellet, J. P., Courtois, R., Dussault, C., and Breton, L. 2002. Effects of sampling effort based on GPS telemetry on home-range size estimations. Journal of Wildlife Management 66: 12901300.Google Scholar
Gitzen, R. A. and Millspaugh, J. J. 2003. Comparison of least-squares cross-validation bandwidth options for kernel home-range estimation. Wildlife Society Bulletin 31: 823831.Google Scholar
Gould, L. and Sauther, M. 2007. Lemuriformes. Pages 4672 in Primates in Perspective. Campbell, C. J., Fuentes, A., MacKinnon, K. C., Panger, M., and Bearder, S. (Eds.). Oxford University Press, Oxford.Google Scholar
Grueter, C. C., Li, D., Ren, B., and Wei, F. 2009. Choice of analytical method can have dramatic effects on primate home range estimates. Primates 50: 8184.Google Scholar
Gula, R. and Theuerkauf, J. 2013. The need for standardization in wildlife science: home range estimators as an example. European Journal of Wildlife Research 59: 713718.Google Scholar
Gursky, S. 2007. Tarsiiformes. Pages 7385 in Primates in Perspective. Campbell, C. J., Fuentes, A., MacKinnon, K. C., Panger, M., and Bearder, S. (Eds.). Oxford University Press, Oxford.Google Scholar
Hadi, S., Ziegler, T., Waltert, M., et al. 2012. Habitat use and trophic niche overlap of two sympatric colobines, Presbytis potenziani and Simias concolor, on Siberut Island, Indonesia. International Journal of Primatology 33: 218232.Google Scholar
Harris, T. R. and Chapman, C. A. 2007. Variation in diet and ranging of black and white colobus monkeys in Kibale National Park, Uganda. Primates 48: 208221.Google Scholar
Haugen, A. O. 1942. Home range of cottontail rabbit. Ecology 23: 354367.Google Scholar
Hayne, D. W. 1949. Calculation of size of home range. Journal of Mammalogy 30: 118.Google Scholar
Hemingway, C. A. and Bynum, N. 2005. The influence of seasonality on primate diet and ranging. Pages 57104 in Seasonality in Primates: Studies of Living and Extinct Human and Non-human Primates. Brockman, D. K. and van Schaik, C. P. (Eds.). Cambridge University Press, Cambridge.Google Scholar
Hemson, G., Johnson, P., South, A., et al. 2005. Are kernels the mustard? Data from global positioning system (GPS) collars suggest problems for kernel home range analyses with least-squares-cross-validation. Journal of Animal Ecology 74: 455463.Google Scholar
Henzi, S. P., Brown, L. R., Barrett, L., and Marais, A. J. 2011. Troop size, habitat use, and diet of chacma baboons (Papio hamadryas ursinus) in commercial pine plantations: implications for management. International Journal of Primatology 32: 10201032.Google 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: 143167.Google Scholar
Hoffman, T. S. and O’Riain, M. J. 2011. The spatial ecology of chacma baboons (Papio ursinus) in a human-modified environment. International Journal of Primatology 32: 308328.Google Scholar
Hooge, P. N. and Eichenlaub, B. 2000. Animal movement extension to ArcView. Ver. 2.0. Alaska Science Center, Biological Science Office, U.S. Geological Survey, Anchorage, AK.Google Scholar
Horne, J. S. and Garton, E. O. 2006a. Likelihood cross-validation versus least squares cross-validation for choosing the smoothing parameter in kernel home range analysis. Journal of Wildlife Management 70: 641648.Google Scholar
Horne, J. S. and Garton, E. O. 2006b. Selecting the best home range model: an information theoretic approach. Ecology 87: 11461152.Google Scholar
Horne, J. S. and Garton, E. O. 2009. Animal space use 1.3. Available at: www.cnr.uidaho.edu/population_ecology/animal_space_use.htm.Google Scholar
Horne, J. S., Garton, E. O., Krone, S. S., and Lewis, J. L. 2007. Analyzing animal movements using Brownian bridges. Ecology 88: 23542363.Google Scholar
Huck, M., Davison, J., and Roper, T. J. 2008. Comparison of two sampling protocols and four home-range estimators using radio-tracking data from urban badgers Meles meles. Wildlife Biology 14: 467477.Google Scholar
Irwin, M. T. 2008. Diademed sifaka (Propithecus diadema) ranging and habitat use in continuous and fragmented forest: higher density but lower viability in fragments? Biotropica 40: 231240.CrossRefGoogle Scholar
Janmaat, K. R. L., Olupot, W., Chancellor, R. L., Arlet, M. E., and Waser, P. M. 2009. Long-term site fidelity and individual home range shifts in Lophocebus albigena. International Journal of Primatology 30: 443466.Google Scholar
Jennrich, R. I. and Turner, F. B. 1969. Measurement of non-circular home range. Journal of Theoretical Biology 22: 227237.Google Scholar
Kie, J. G., Matthiopoulos, J., Fieberg, J., et al. 2010. The home-range concept: are traditional estimators still relevant with modern telemetry technology? Philosophical Transactions of the Royal Society B 365: 22212231.Google Scholar
Kirkpatrick, R. C. 2007. The Asian colobines: diversity among leaf-eating monkeys. Pages 186200 in Primates in Perspective. Campbell, C. J., Fuentes, A., MacKinnon, K. C., Panger, M., and Bearder, S. (Eds.). Oxford University Press, Oxford.Google Scholar
Knight, C. M., Kenward, R. E., Gozlan, R. E., et al. 2009. Home range estimation within complex restricted environments: importance of method selection in detecting seasonal change. Wildlife Research 36: 213224.Google Scholar
Lahann, P. 2008. Habitat utilization of three sympatric cheirogaleid lemur species in a littoral rain forest of southeastern Madagascar. International Journal of Primatology 29: 117134.Google Scholar
Laver, P. N. and Kelly, M. J. 2008. A critical review of home range studies. Journal of Wildlife Management 72: 290298.Google Scholar
Lawson, E. J. G. and Rodgers, A. R. 1997. Differences in home-range size computed in commonly used software programs. Wildlife Society Bulletin 25: 721729.Google Scholar
Lehmann, J. and Boesch, C. 2003. Social influences on ranging patterns among chimpanzees (Pan troglodytes verus) in the Taï National Park, Côte d’Ivoire. Behavioral Ecology 14: 642649.Google Scholar
Li, Y., Jiang, Z., Li, C., and Grueter, C. C. 2010. Effects of seasonal folivory and frugivory on ranging patterns in Rhinopithecus roxellana. International Journal of Primatology 31: 609626.Google Scholar
Lichti, N. I. and Swihart, R. K. 2011. Estimating utilization distributions with kernel versus local convex hull methods. Journal of Wildlife Management 75: 413422.CrossRefGoogle Scholar
Long, J. A. and Nelson, T. A. 2012. Time geography and wildlife home range delineation. Journal of Wildlife Management 76: 407413.Google Scholar
Lyons, A. J., Turner, W. C., and Getz, W. M. 2013. Home range plus: a space-time characterization of movement over real landscapes. Movement Ecology 1: 2.Google Scholar
Matsuda, I., Tuuga, A., and Higashi, S. 2009. Ranging behavior of proboscis monkeys in a riverine forest with special reference to ranging in inland forest. International Journal of Primatology 30: 313325.Google Scholar
Mekonnen, A., Bekele, A., Fashing, P. L., Hemson, G., and Atickem, A. 2010. Diet, activity patterns, and ranging ecology of the bale monkey (Chlorocebus djamdjamensis) in Odobullu Forest, Ethiopia. International Journal of Primatology 31: 339362.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
Milton, K. and May, M. L. 1976. Body weight, diet and home range area in primates. Nature 259: 459462.Google Scholar
Mochizuki, S. and Murakami, T. 2013. Scale dependent effects in resource selection by crop-raiding Japanese macaques in Niigata Prefecture, Japan. Applied Geography 42: 1322.Google Scholar
Mohr, C. O. 1947. Table of equivalent populations of North American small mammals. The American Midland Naturalist 37: 223449.Google Scholar
Moorcroft, P. R. 2012. Mechanistic approaches to understanding and predicting mammalian space use: recent advances, future directions. Journal of Mammalogy 93: 903916.Google Scholar
Moorcroft, P. R. and Lewis, M. A. 2006. Mechanistic Home Range Analysis. Princeton University Press, Princeton, NJ.Google Scholar
Moorcroft, P. R., Lewis, M. A., and Crabtree, R. L. 1999. Home range analysis using a mechanistic home range model. Ecology: 80: 16561665.Google Scholar
Nekaris, A. and Bearder, S. K. 2007. The lorisiform primates of Asia and mainland Africa: diversity shrouded in darkness. Pages 2445 in Primates in Perspective. Campbell, C. J., Fuentes, A., MacKinnon, K. C., Panger, M., and Bearder, S. (Eds.). Oxford University Press, Oxford.Google Scholar
Neri-Arboleda, I., Stott, P., and Arboleda, N. P. 2002. Home ranges, spatial movements and habitat associations of the Philippine tarsier (Tarsius syrichta) in Corella, Bohol. Journal of Zoology 257: 387402.Google Scholar
Newton-Fisher, N. E. 2003. The home range of the Sonso community of chimpanzees from the Budongo Forest, Uganda. African Journal of Ecology 41: 150156.Google Scholar
Norconk, M. A. 2007. Sakis, uakaris, and titi monkeys: behavioral diversity in a radiation of primate seed predators. Pages 123138 in Primates in Perspective. Campbell, C. J., Fuentes, A., MacKinnon, K. C., Panger, M., and Bearder, S. (Eds.). Oxford University Press, Oxford.Google Scholar
Norscia, I. and Borgognini-Tarli, S. M. 2008. Ranging behavior and possible correlates of pair-living in southeastern avahis (Madagascar). International Journal of Primatology 29: 153171.Google Scholar
Ostro, L. E. T., Young, T. P., Silver, S. C., and Koontz, F. W. 1999. A geographic information system method for estimating home range size. Journal of Wildlife Management 63: 748755.Google Scholar
Otis, D. L. and White, G. C. 1999. Autocorrelation of location estimates and the analysis of radiotracking data. The Journal of Wildlife Management 63: 10391044.Google Scholar
Palminteri, S. and Peres, C. A. 2012. Habitat selection and use of space by bald-faced sakis (Pithecia irrorata) in southwestern Amazonia: lessons from a multiyear, multigroup study. International Journal of Primatology 33: 401417.Google Scholar
Pebsworth, P. A., MacIntosh, A. J. J., Morgan, H. R., and Huffman, M. A. 2012a. Factors influencing the ranging behavior of chacma baboons (Papio hamadryas ursinus) living in a human-modified habitat. International Journal of Primatology 33: 872887.Google Scholar
Pebsworth, P. A., Morgan, H. R., and Huffman, M. A. 2012b. Evaluating home range techniques: use of global positioning system (GPS) collar data from chacma baboons. Primates 53: 345355.Google Scholar
Pimley, E. R., Bearder, S. K., and Dixson, A. F. 2005. Home range analysis of Perodicticus potto edwardsi and Sciurocheirus cameronensis. International Journal of Primatology 26: 191206.Google Scholar
Porter, L. M., Sterr, S. M., and Garber, P. A. 2007. Habitat use and ranging behavior of Callimico goeldii. International Journal of Primatology 28: 10351058.Google Scholar
Powell, R. A. 2000. Animal home ranges and territories and home range estimators. Pages 65110 in Research Technologies in Animal Ecology: Controversies and Consequences. Boitani, L. and Fuller, T. K. (Eds.). Columbia University Press, New York.Google Scholar
Powell, R. A. and Mitchell, M. S. 2012. What is a home range? Journal of Mammalogy 93: 948958.Google Scholar
Pyritz, L. W., Kappeler, P. M., and Fichtel, C. 2011. Coordination of group movements in wild red-fronted lemurs (Eulemur rufifrons): processes and influence of ecological and reproductive seasonality. International Journal of Primatology 32: 13251347.Google Scholar
Raboy, B. E. and Dietz, J. M. 2004. Diet, foraging, and use of space in wild golden-headed lion tamarins. American Journal of Primatology 63: 115.Google Scholar
Ramos-Fernández, G., Aguilar, S. E. S., Schaffner, C. M., Vick, L. G., and Aureli, F. 2013. Site fidelity in space use by spider monkeys (Ateles geoffroyi) in the Yucatan Peninsula, Mexico. PLoS ONE 8: e62813.Google Scholar
Ren, B., Li, M., Long, Y., Wu, R., and Wei, F. 2009. Home range and seasonality of Yunnan snub-nosed monkeys. Integrative Zoology 4: 162171.Google Scholar
Robbins, M. M. 2007. Gorillas: diversity in ecology and behavior. Pages 305321 in Primates in Perspective. Campbell, C. J., Fuentes, A., MacKinnon, K. C., Panger, M., and Bearder, S. (Eds.). Oxford University Press, Oxford.Google Scholar
Robbins, M. M. and McNeilage, A. 2003. Home range and frugivory patterns of mountain gorillas in Bwindi Impenetrable National Park, Uganda. International Journal of Primatology 24: 467491.Google Scholar
Rodgers, A. R. and Carr, A. P. 1998. HRE: the Home Range Extension for ArcView. Thunder Bay, Ontario: Centre for Northern Forest Ecosystem Research, Ontario Ministry of Natural Resources, Centre for Northern Forest Ecosystem Research, Thunder Bay, Ontario.Google Scholar
Rodgers, A. R., Carr, A. P., Beyer, H. L., Smith, L., and Kie, J. G. 2007. HRT: Home Range Tools for ArcGIS. Version 1.1. Ontario Ministry of Natural Resources, Centre for Northern Forest Ecosystem Research, Thunder Bay, Ontario.Google Scholar
Row, J. R. and Blouin-Demers, G. 2006. Kernels are not accurate estimators of home-range size for herpetofauna. Copeia 4: 797802.Google Scholar
Sawyer, S. C. 2012. Subpopulation range estimation for conservation planning: a case study of the critically endangered Cross River gorilla. Biodiversity and Conservation 21: 15891606.Google Scholar
Scull, P., Palmer, M., Frey, F., and Kraly, E. 2012. A comparison of two home range modeling methods using Ugandan mountain gorilla data. International Journal of Geographical Information Science 26: 21112121.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: 20752085.Google Scholar
Seaman, D. E., Millspaugh, J. J., Kernohan, B. J., et al. 1999. Effects of sample size on kernel home range estimates. Journal of Wildlife Management 63: 739747.Google Scholar
Shaffer, C. A. 2013. Ecological correlates of ranging behavior in bearded sakis (Chiropotes sagulatus) in a continuous forest in Guyana. International Journal of Primatology 34: 515532.Google Scholar
Silverman, B. W. 1986. Density Estimation for Statistics and Data Analysis. Chapman & Hall, London.Google Scholar
Siniff, D. B. and Tester, J. R. 1965. Computer analysis of animal movement data obtained by telemetry. BioScience 15: 104108.Google Scholar
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 Yasuní National Park, Ecuador. American Journal of Primatology 72: 129141.Google Scholar
Spencer, W. D. 2012. Home range and the value of spatial information. Journal of Mammalogy 93: 929947.Google Scholar
Steiniger, S. and Hunter, A. J. S. 2012. OpenJUMP HoRAE: a free GIS and toolbox for home range analysis. Wildlife Society Bulletin 36: 600608.Google Scholar
Steiniger, S. and Hunter, A. J. S. 2013. A scaled line-based kernel density estimator for the retrieval of utilization distributions and home ranges from GPS movement tracks. Ecological Informatics 13: 18.Google Scholar
Sterling, E., Nguyen, N., and Fashing, P. 2000. Spatial patterning in nocturnal prosimians: a review of methods and relevance to studies of sociality. American Journal of Primatology 51: 319.Google Scholar
Stumpf, R. 2007. Chimpanzees and bonobos: diversity within and between species. Pages 321344 in Primates in Perspective. Campbell, C. J., Fuentes, A., MacKinnon, K. C., Panger, M., and Bearder, S. (Eds.). Oxford University Press, Oxford.Google Scholar
Swihart, R. K. and Slade, N. A. 1985. Testing for independence of observations in animal movements. Ecology 66: 11761184.Google Scholar
Swihart, R. K. and Slade, N. A. 1997. On testing for independence of animal movements. Journal of Agricultural, Biological, and Environmental Statistics 2: 4863.Google Scholar
Takasaki, H. 1981. Troop size, habitat quality, and home range area in Japanese macaques. Behavioral Ecology and Sociobiology 9: 277281.Google Scholar
Tsuji, Y. and Takatsuki, S. 2009. Effects of yearly change in nut fruiting on autumn home-range use by Macaca fuscata on Kinkazan Island, northern Japan. International Journal of Primatology 30: 169181.Google Scholar
Volampeno, M. S. N., Masters, J. C., and Downs, C. T. 2011. Home range size in the blue-eyed black lemur (Eulemur flavifrons): a comparison between dry and wet seasons. Mammalian Biology 76: 157164.Google Scholar
Walter, W. D., Fischer, J. W., Baruch-Mordo, S., and VerCauteren, K. C. 2011. What is the proper method to delineate home range of an animal using today’s advanced GPS telemetry systems: the initial step. Pages 249268 in Modern Telemetry. Krejcar, O. (Ed.). InTech Open Access Publisher, London.Google Scholar
Wartmann, F. M., Purves, R. S., and van Schaik, C. P. 2010. Modelling ranging behavior of female orang-utans: a case study in Tuanan, Central Kalimantan, Indonesia. Primates 51: 119130.Google Scholar
Wauters, L. A., Preatoni, D. G., Molinari, A. and Tosi, G. 2007. Radio tracking squirrels: performance of home range density and linkage estimators with small range and sample size. Ecological Modelling 3 -4: 333344.Google Scholar
Weber, K. T., Burcham, M., and Marcum, C. L. 2001. Assessing independence of animal locations with association matrices. Journal of Range Management 54: 2124.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
Willems, E. P. and Hill, R. A. 2009. Predator-specific landscapes of fear and resource distribution: effects on spatial range use. Ecology 90: 546555.Google Scholar
Worton, B. J. 1987. A review of models of home range for animal movement. Ecological Modelling 38: 277298.Google Scholar
Worton, B. J. 1989. Kernel methods for estimating the utilization distribution in home-range studies. Ecology 70: 164168.Google Scholar
Worton, B. J. 1995. Using Monte Carlo simulation to evaluate kernel-based home range estimators. Journal of Wildlife Management 59: 794800.Google Scholar

References

Altmann, J. 1974. Observational study of behavior: sampling methods. Behaviour 49: 227267.Google Scholar
Anderson, D. P., Nordheim, E. V., Boesch, C., and Moermond, T. C. 2002. Factors Influencing Fission-Fusion Grouping in Chimpanzees in the Taï National Park, Côte d’Ivoire: Behavioural Diversity in Chimpanzees and Bonobos. Cambridge University Press, Cambridge.Google Scholar
Anselin, L. 1995. Local indicators of spatial association: LISA. Geographical Analysis 27: 93115.Google Scholar
Baily, T. C. and Gatrell, A. C. 1995. Interactive Spatial Data Analysis. Longman, New York.Google Scholar
Boyle, S. A., Lourenco, W. C., da Silva, L. R., and Smith, A. T. 2009. Home range estimates vary with sample size and methods. Folia Primatologica 80: 3342.Google Scholar
Byrne, R. W., Whiten, A., and Henzi, S. P. 1990. Measuring the food constraints of mountain baboons. Pages 105122 in Baboons: Behavior and Ecology, Use and Care. de Mello, M. T., Whiten, A., and Byrne, R. W.. International Primatological Society, Brazil.Google Scholar
Chapman, C. A. 1988. Patterns of foraging and range use by three species of neotropical primates. Primates 29(2): 177194.Google Scholar
Chapman, C. A. 1990. Ecological constraints on group size in three species of neotropical primates. Folia Primatologica 55: 19.Google Scholar
Chapman, C. A. and Chapman, L. J. 2000. Determinants of group size in social primates: the importance of travel costs. Pages 2442 in On the Move: How and Why Animals Travel in Groups. Boinski, S. and Garber, P. (Eds.). University of Chicago Press, Chicago.Google Scholar
Chapman, C. A., Chapman, L. J., Wrangham, R., Gebo, D., and Gardner, L. 1992. Estimators of fruit abundance of tropical trees. Biotropica 24: 527531.Google Scholar
Chapman, C. A., White, F., Wrangham, R. W. 1994. Party size in chimpanzees and bonobos: a reevaluation of theory based on two similarly forested sites. Pages 4557 in Chimpanzee Cultures. McGrew, W. C., Marchant, L. F., and Nishida, T. (Eds.). Harvard University Press, Cambridge, MA.Google 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 36: 5970.Google Scholar
Charnov, E. L. 1976. Optimal foraging, the marginal value theorem. Theoretical Population Biology 9: 129136.Google Scholar
Clark, P. J. and Evans, F. C. 1954. Distance to nearest neighbor as a measure of spatial relationships in populations. Ecology 35: 445453Google Scholar
Coder, K. D. 2000. Crown Shape Factors & Volumes. University of Georgia, Warnell School of Forest Resources, Athens, GA.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
Dale, M. R. T. 1999. Spatial Pattern Analysis in Plant Ecology. Cambridge University Press, Cambridge.Google Scholar
Dale, M. R. T. and Powell, R. D. 2001. A new method for characterizing point patterns in plant ecology. Journal of Vegetation Science 12(5): 597608.Google Scholar
Dale, M. R. T., Dixon, P. M., Fortin, M. J., et al. 2002. Conceptual and mathematical relationships among methods for spatial analysis. Ecography 25: 558577.Google Scholar
Dammhahn, M. and Kappeler, P. M. 2008. Comparative feeding ecology of sympatric Microcebus berthae and M. murinus. International Journal of Primatology 29(6): 1567.Google Scholar
De Knegt, H. J., van Langevelde, F., Coughenour, M. B., et al. 2010. Spatial autocorrelation and the scaling of species–environment relationships. Ecology 91: 24552465.Google Scholar
Dias, L. G. and Strier, K. B. 2003. Effects of group size on ranging patterns in Brachyteles arachnoides hypoxanthus. International Journal of Primatology 24: 209221.Google Scholar
Di Fiore, A. 1997. Ecology and behavior of lowland woolly monkeys (Lagothrix lagotricha poeppigii, Atelinae) in Eastern Ecuador. PhD Thesis. University of California, Davis.Google Scholar
Diggle, P. J. 1983. The Statistical Analysis of Spatial Point Patterns. Academic Press, New York.Google Scholar
Dixon, P. M. 2002. Ripley’s K function. Encyclopedia of Environmetrics 3: 17961803.Google Scholar
Drake, J. B., Dubayah, R. O., Knox, , et al. 2002. Sensitivity of large-footprint Lidar to canopy structure and biomass in a neotropical rainforest. Remote Sensing of Environment. 81: 378392.Google Scholar
Ebdon, D. 1985. Statistics in Geography. Blackwell, Malden, MA.Google Scholar
Everitt, B. S., Landau, S., and Leese, M. 2001. Cluster Analysis, 4th edition. Arnold, London.Google Scholar
Fatoyinbo, L. 2012. Remote Sensing of Biomass: Principles and Applications. Intech, London.Google Scholar
Furnival, G. M. 1961. An index comparing equations used in constructing volume equations. Forestry Science 7: 337341.Google Scholar
Forman, R. T. T. 1995. Land Mosaics: The Ecology of Landscapes and Regions. Cambridge University Press, Cambridge.Google Scholar
Fortin, M., Mark, M. J., and Dale, R. T. 2005. Spatial Analysis: A Guide for Ecologists. Cambridge University Press, Cambridge.Google Scholar
Fourrier, M. S. 2013. The spatial and temporal ecology of seed dispersal by gorillas in Lopé National Park, Gabon: linking patterns of disperser behavior and recruitment in an Afrotropical forest. Dissertation, Washington University in St. Louis.Google Scholar
Freeman, E. A. and Ford, E. D. 2002. Effects of data quality on analysis of ecological pattern using the K(d) statistical function. Ecology 83: 3546.Google Scholar
Fruth, B. and Hohmann, G. 1996. Nest building behavior in the great apes: the great leap forward. Pages 225240 in Great Ape Societies. McGrew, W., Marchant, L., and Nishida, T. (Eds.). Cambridge University Press, Cambridge.Google Scholar
Haase, P. 1995. Spatial pattern analysis in ecology based on Ripley’s K-function: introduction and methods of edge correction. Journal of Vegetation Science 6: 575582.Google Scholar
Haase, P. 2001. Can isotropy vs. anisotropy in the spatial association of plant species reveal physical vs. biotic facilitation? Journal of Vegetation Science 12: 127136.Google Scholar
Hardesty, B. D., Hubbell, S. P., and Bermingham, E. 2006. Genetic evidence of frequent long-distance recruitment in a vertebrate-dispersed tree. Ecology Letters 9: 516525.Google Scholar
Harding, D. J., Lefsky, M. A., Parker, G. G., and Blair, J. B. 2001. Laser altimeter canopy height profiles: methods and validation for closed-canopy, broadleaf forests. Remote Sensing of Environment 76: 283297.Google Scholar
Hoffman, T. S. and O’Riain, M. J. 2012. Landscape requirements of a primate population in a human-dominated environment. Frontiers in Zoology 9. DOI: 10.1186/1742-9994-9-1.Google Scholar
Isbell, L. A. 1991. Contest and scramble competition: patterns of female aggression and ranging behavior among primates. Behavioral Ecology 2: 143155.Google Scholar
Isbell, L. A., Pruetz, J. D., Young, T. P., and Lewis, M. 1998. Movements of vervets (Cercopithecus aethiops) and patas monkeys (Erythrocebus patas) as estimators of food resource size, density, and distribution. Behavioral Ecology and Sociobiology 42: 123133.Google Scholar
Janson, C. 1985. Aggressive competition and individual food consumption in wild brown capuchin monkeys (Cebus apella). Behavioral Ecology and Sociobiology 18(2): 125138.Google Scholar
Janson, C. H. 1992. Evolutionary ecology of primate social structure. Pages 95130 in Evolutionary Ecology and Human Behavior. Smith, E. A. and Winterhalder, B. (Eds.). Aldine, New York.Google Scholar
Janson, C. H. and Goldsmith, M. L. 1995. Predicting group size in primates: foraging costs and predation risks. Behavioral Ecology 6(3): 326336.Google Scholar
Koenig, A. 2002. Competition for resources and its behavioral consequences among female primates. International Journal of Primatology 23: 759783.Google Scholar
Koenig, A., Beise, J., Chalise, M. K., and Ganzhorn, J. U. 1998. When females should contest for food: testing hypotheses about resource density, distribution, size, and quality with Hanuman langurs (Presbytis entellus). Behavioral Ecology and Sociobiology 42: 225237.Google Scholar
Krebs, C. 1999. Ecological Methodology. Addison-Wesley, Menlo Park, CA.Google Scholar
Lacher, T. E., Bouchardet da Fonseca, G. A., Alves, C., and Magalhaes-Castro, B. 1984. Parasitism of trees by marmosets in a central Brazilian gallery forest. Biotropica 16: 202209.Google Scholar
Legendre, P. and Fortin, M. J. 1989. Spatial pattern and ecological analysis. Plant Ecology 80: 107138.Google Scholar
Leighton, M. and Leighton, D. R. 1982. The relationship of size and feeding aggregate size to size of food patch: howler monkeys (Alouatta palliata) feeding in Trichilia cipo fruit trees on Barro Colorado Island. Biotropica 14: 8190.Google Scholar
Levi, T., Silvius, K. M., Oliveira, L. F., Cummings, A. R., and Fragoso, J. M. 2013. Competition and facilitation in the capuchin–squirrel monkey relationship. Biotropica 45(5): 636643.Google Scholar
Masota, A. M., Zahabu, E., Malimbwi, R. E., Bollandsas, O. E. and Eid, R. H. 2014. Volume models for single trees in tropical rainforests in Tanzania. Journal of Energy and Natural Resources 3: 6676.Google Scholar
Nakagawa, N. 1989. Feeding strategies of Japanese monkeys against the deterioration of habitat quality. Primates 30(1): 116.Google Scholar
Norconk, M. A., Kinzey, W. G. 1994. Challenge of neotropical frugivory: travel patterns of spider monkeys and bearded sakis. Am J Primatol 34: 171183.Google Scholar
Ord, J. K. and Getis, A. 1995. Local spatial autocorrelation statistics: Distributional issues and an application. Geographical Analysis 27: 287306.Google Scholar
Peper, P. J., McPherson, E. G., and Mori, S. M. 2001. Equations for predicting diameter, height, crown width, and leaf area of San Joaquin Valley street trees. Journal of Arboriculture 27: 306317.Google Scholar
Peres, C. A. 1991. Seed predation of Cariniana micrantha (Lecythidaceae) by brown capuchin monkeys in Central Amazonia. Biotropica 23(3): 262270.Google Scholar
Peres, C. A. 1996. Use of space, spatial group structure, and foraging group size of gray woolly monkeys (Lagothrix lagotricha cana) at Urucu, Brazil. Pages 467488 in Adaptive Radiations of Neotropical Primates. Norconk, M. A., Rosenberger, A. L., and Garber, P. A. (Eds.). Plenum Press, New York.Google Scholar
Perry, J. N., Liebhold, A. M., Rosenberg, M. S., et al. 2002. Illustrations and guidelines for selecting statistical methods for quantifying spatial pattern in ecological data. Ecography 25: 578600.Google Scholar
Phillips, K. A. 1995. Resource patch size and flexible foraging in white-face capuchins (Cebus capucinus). International Journal of Primatology 16: 509521.Google Scholar
Pickett, S. T. A. and White, P. S. 1985. The Ecology of Natural Disturbance and Patch Dynamics. Academic Press, New York.Google Scholar
Pruetz, J. D. 1999. Socioecology of adult female vervet (Cercopithecus aethiops) and patas monkeys (Erythrocebus patas) in Kenya: food availability, feeding competition, and dominance relationships. PhD Thesis. University of Illinois.Google Scholar
Pruetz, J. D. and Isbell, L. A. 2000. Correlations of food distribution and patch size with agonistic interactions in female vervets (Chlorocebus aethiops) and patas monkeys (Erythrocebus patas) living in simple habitats. Behavioral Ecology and Sociobiology 49(1): 3847.Google Scholar
Ripley, B. D. 1976. The second-order analysis of stationary point processes. Journal of Applied Probability 13: 255266.Google Scholar
Russo, S. E. and Augspurger, C. K. 2004. Aggregated seed dispersal by spider monkeys limits recruitment to clumped patterns in Virola calophylla. Ecology Letters 7(11): 10581067.Google Scholar
Russo, S. E., Portnoy, S., and Augspurger, C. K. 2006. Incorporating animal behavior into seed dispersal models: implications for seed shadows. Ecology 87(12): 31603174.Google Scholar
Schoener, T. W. 1971. Theory of feeding strategies. Annual Review of Ecology and Systems 2: 369403.Google Scholar
Sezen, U. U., Chazdon, R. L., and Holsinger, K. E. 2009. Proximity is not a proxy for parentage in an animal-dispersed Neotropical canopy palm. Proceedings of the Royal Society B: Biological Sciences 2761664: 2037–2044.Google Scholar
Shaffer, C. A. 2012. Ranging behavior, group cohesiveness, and patch use in northern bearded sakis (Chiropotes sagulatus) in Guyana. PhD Thesis. Washington University in St. Louis.Google Scholar
Shaffer, C. A. 2013a. GIS analysis of patch use and group cohesiveness of bearded sakis (Chiropotes sagulatus) in the Upper Essequibo Conservation Concession, Guyana. American Journal of Physical Anthropology 150: 235246.Google Scholar
Shaffer, C. A. 2013b. Ecological correlates of ranging behavior in bearded sakis (Chiropotes sagulatus) in a continuous forest in Guyana. International Journal of Primatology 34: 515532.Google Scholar
Snaith, T. V. and Chapman, C. A. 2007. Primate group size and interpreting socioecological models: do folivores really play by different rules? Evolutionary Anthropology 16: 94106.Google Scholar
Stephens, D. and Krebs, J. 1986. Foraging Theory. Princeton University Press, Princeton, NJ.Google Scholar
Sterk, E. A., Watts, D. P. and van Schaik, C. P. 1997. The evolution of female social relationships in primates. Behavioral Ecology and Sociobiology 41: 291310.Google Scholar
Stevenson, P. R., Quinones, M. J., and Ahumada, J. A. 1998. Effects of fruit patch availability on feeding subgroup size and spacing patterns in four primate species at Tinigua National Park, Colombia. International Journal of Primatology 19(2): 313324.Google Scholar
Strier, K. B. 1989. Effects of patch size on feeding associations in muriquis (Brachyteles arachnoides). Folia primatologica 52: 7077.Google Scholar
Symington, M. M. 1988. Food competition and foraging party size in the black spider monkey (Ateles paniscus chamek). Behaviour 105: 117134.Google Scholar
Temerin, L. A. and Cant, J. H. 1983. The evolutionary divergence of Old World monkeys and apes. American Naturalist 122: 335351.Google Scholar
Thompson, C. L., Robl, N. J., Oliveira Melo, L. C., et al. 2013. Spatial distribution and exploitation of trees gouged by common marmosets (Callithrix jacchus). International Journal of Primatology 34: 6585.Google Scholar
Tobler, W. 1970. A computer movie simulating urban growth in the Detroit region. Economic Geography 46(Supplement): 234240.Google Scholar
Upton, G. and Fingleton, B. 1985. Spatial Data Analysis by Example: Volume 1: Point Pattern and Quantitative Data. Wiley, Chichester.Google Scholar
van Schaik, C.P. 1989. The ecology of social relationships amongst female primates. Pages 195218 in Comparative Socioecology: The Behavioural Ecology of Humans and Other Mammals. Standen, V. and Foley, R.A. (Eds.). Blackwell Scientific, Oxford.Google Scholar
Vogel, E. R. and Janson, C. H. 2007. Predicting the frequency of food-related agonism in white-faced capuchin monkeys (Cebus capucinus), using a novel focal-tree method. American Journal of Primatology 69: 533550.Google Scholar
Vogel, E. R. and Janson, C. H. 2011. Quantifying primate food distribution and abundance for socioecological studies: an objective consumer-centered method. International Journal of Primatology 32: 737754.Google Scholar
Warren, W.G. 1972. Point processes in forestry. Pages 85116 in Stochastic Point Processes, Statistical Analysis, Theory and Application. Lewis, P.A.W. (Ed.). Wiley, New York.Google Scholar
White, F. J. and Wrangham, R. W. 1988. Feeding competition and patch size in the chimpanzee species Pan paniscus and Pan troglodytes. Behaviour 105: 148164.Google Scholar
Wrangham, R.W., Gittleman, J.L., and Chapman, C.A. 1993. Constraints on group size in primates and carnivores: population density and day-range as assays of exploitation competition. Behavioral Ecology and Sociobiology 32(3): 199209.Google Scholar

References

Akers, A. A., Islam, M. A., and Nijman, V. 2013. Habitat characterization of western hoolock gibbons Hoolock hoolock by examining home range microhabitat use. Primates 54 : 341348.Google Scholar
Bergman, C. M., Fryxell, J. M., Gates, C. C., and Fortin, D. 2001. Ungulate foraging strategies: energy maximizing or time minimizing? Journal of Animal Ecology 70 : 289300.Google Scholar
Börger, L., Franconi, N., Ferretti, F., et al. 2006. An integrated approach to identify spatiotemporal and individual-level determinants of animal home range size. American Naturalist 168 : 471485.Google Scholar
Boyle, S. A., Lourenco, W. C., da Silva, L. R., and Smith, A. T. 2009. Travel and spatial patterns change when Chiropotes satanas chiropotes inhabit forest fragments. International Journal of Primatology 30 : 515531.Google Scholar
Burt, W. H. 1943. Territoriality and home range concepts as applied to mammals. Journal of Mammalogy 24 : 346352.Google Scholar
Carbone, C., Cowlishaw, G., Isaac, N. J. B., and Rowcliffe, J. M. 2005. How far do animals go? Determinants of day range in mammals. American Naturalist 165 : 290297.Google Scholar
Chapman, C. A. and Chapman, L. J. 2000. Determinants of group size in primates: the importance of travel costs. Pages 2442 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
Clissold, F. J., Sanson, G. D., Read, J., and Simpson, S. J. 2009. Gross vs. net income: how plant toughness affects performance of an insect herbivore. Ecology 90 : 33933405.Google Scholar
Fieberg, J. and Börger, L. 2012. Could you please phrase “home range” as a question? Journal of Mammalogy 93 : 890902.Google Scholar
Foerster, S., Zhong, Y., Pintea, L., et al. 2017. Feeding habitat quality and behavioral trade-offs in chimpanzees: a case for species distribution models. Behavioral Ecology 27 : 10041016.Google Scholar
Fretwell, S. D. 1972. Populations in a Seasonal Environment. Princeton University Press, Princeton, NJ.Google Scholar
Ganzhorn, J. U. 1995. Low-level forest disturbance effects on primary production, leaf chemistry, and lemur populations. Ecology 76 : 20842096.Google Scholar
Hanya, G. and Bernard, H. 2016. Seasonally consistent small home range and long ranging distance in Presbytis rubicunda in Danum Valley, Borneo. International Journal of Primatology 37 : 390404.Google Scholar
Harcourt, A. 1998. Ecological indicators of risk for primates, as judged by species’ susceptibility to logging. Pages 5679 in Behavioral Ecology and Conservation Biology. Caro, T. (Ed.). Oxford University Press, Oxford.Google Scholar
Harper, G. J., Steininger, M. K., Tucker, C. J., Juhn, D., and Hawkins, F. 2007. Fifty years of deforestation and forest fragmentation in Madagascar. Environmental Conservation 34 : 325333.Google Scholar
Harrison, M. E., Morrogh-Bernard, H. C., and Chivers, D. J. 2010. Orangutan energetics and the influence of fruit availability in the nonmasting peat-swamp forest of Sabangau, Indonesian Borneo. International Journal of Primatology 31 : 585607.Google Scholar
Hixon, M. A. 1982. Energy maximizers and time minimizers: theory and reality. American Naturalist 119 : 596599.Google Scholar
Hooge, P. N. and Eichenlaub, B. 1997. Animal Movement Extension to ArcView. 2.0 ed. Anchorage, AK: Alaska Science Center – Biological Science Office, US Geological Survey.Google Scholar
Irwin, M. T. 2006a. Ecological impacts of forest fragmentation on diademed sifakas (Propithecus diadema) at Tsinjoarivo, eastern Madagascar: implications for conservation in fragmented landscapes. PhD Thesis, Stony Brook University.Google Scholar
Irwin, M. T. 2006b. Ecologically enigmatic lemurs: The sifakas of the eastern forests (Propithecus candidus, P. diadema, P. edwardsi, P. perrieri, and P. tattersalli). Pages 305326 in Lemurs: Ecology and Adaptation. Gould, L. and Sauther, M. L. (Ed.). Springer, New York.Google Scholar
Irwin, M. T. 2008a. Diademed sifaka (Propithecus diadema) ranging and habitat use in continuous and fragmented forest: higher density but lower viability in fragments? Biotropica 40 : 231240.Google Scholar
Irwin, M. T. 2008b. Feeding ecology of diademed sifakas (Propithecus diadema) in forest fragments and continuous forest. International Journal of Primatology 29 : 95115.Google Scholar
Irwin, M. T., Raharison, J.-L., and Wright, P. C. 2009. Spatial and temporal variability in predation on rainforest primates: do forest fragmentation and predation act synergistically? Animal Conservation 12 : 220230.Google Scholar
Irwin, M. T., Junge, R. E., Raharison, J. L., and Samonds, K. E. 2010. Variation in physiological health of diademed sifakas across intact and fragmented forest at Tsinjoarivo, eastern Madagascar. American Journal of Primatology 72 : 10131025.Google Scholar
Irwin, M. T., Raharison, J.-L., Raubenheimer, D., Chapman, C. A., and Rothman, J. M. 2014. Nutritional correlates of the “lean season”: effects of seasonality and frugivory on the nutritional ecology of diademed sifakas. American Journal of Physical Anthropology 153 : 7891.Google Scholar
Irwin, M. T., Raharison, J.-L., Raubenheimer, D. R., Chapman, C. A., and Rothman, J. M. 2015. The nutritional geometry of resource scarcity: effects of lean seasons and habitat disturbance on nutrient intakes and balancing in wild sifakas. PLoS ONE 10: e0128046.Google Scholar
Irwin, M. T., Samonds, K. E., Raharison, J.-L., et al. 2019. Morphometric signals of population decline in diademed sifakas occupying degraded rainforest habitat in Madagascar. Scientific Reports 9 : 8776. DOI:10.1038/s41598-019-45426-2.Google Scholar
Johns, A. D. and Skorupa, J. P. 1987. Responses of rain-forest primates to habitat disturbance: a review. International Journal of Primatology 8 : 157191.Google Scholar
Kapos, V., Wandelli, E., Camargo, J. L. and Ganade, G. 1997. Edge-related changes in environment and plant responses due to forest fragmentation in central Amazonia. Pages 3344 in Tropical Forest Remnants: Ecology, Management, and Conservation of Fragmented Communities. Laurance, W. F. and Bierregaard, R. O. Jr. (Eds.). University of Chicago Press, Chicago, IL.Google Scholar
Kelt, D. A. and Van Vuren, D. H. 2001. The ecology and macroecology of mammalian home range area. American Naturalist 157 : 637645.Google Scholar
Laurance, W. F., Camargo, J. L. C., Luizao, R. C. C., et al. 2011. The fate of Amazonian forest fragments: a 32-year investigation. Biological Conservation 144 : 5667.Google Scholar
Lowen, C. and Dunbar, R. I. M. 1994. Territory size and defendability in primates. Behavioral Ecology and Sociobiology 35 : 347354.Google Scholar
Marsh, K. J., Wallis, I. R., Andrew, R. L., and Foley, W. J. 2006. The detoxification limitation hypothesis: where did it come from and where is it going? Journal of Chemical Ecology 32 : 12471266.Google Scholar
Marshall, A. J., Boyko, C. M., Feilen, K. L., Boyko, R. H., and Leighton, M. 2009. Defining fallback foods and assessing their importance in primate ecology and evolution. American Journal of Physical Anthropology 140 : 603614.Google Scholar
Marzluff, J. M., Millspaugh, J. J., Hurvitz, P., and Handcock, M. S. 2004. Relating resources to a probabilistic measure of space use: forest fragments and Steller’s Jays. Ecology 85 : 14111427.Google Scholar
Milich, K. M., Stumpf, R. M., Chambers, J. M., and Chapman, C. A. 2014. Female red colobus monkeys maintain their densities through flexible feeding strategies in logged forests in Kibale National Park, Uganda. American Journal of Physical Anthropology 154 : 5260.Google Scholar
Mitani, J. C. and Rodman, P. S. 1979. Territoriality: the relation of ranging pattern and home range size to defendability, with an analysis of territoriality among primate species. Behavioral Ecology and Sociobiology 5 : 241251.Google Scholar
Ostfeld, R. S. 1994. The fence effect reconsidered. Oikos 70 : 340348.Google Scholar
Palminteri, S. and Peres, C. A. 2012. Habitat selection and use of space by bald-faced sakis (Pithecia irrorata) in Southwestern Amazonia: lessons from a multiyear, multigroup study. International Journal of Primatology 33 : 401417.Google Scholar
Palminteri, S., Powell, G. V. N., and Peres, C. A. 2016. Determinants of spatial behavior of a tropical forest seed predator: the roles of optimal foraging, dietary diversification, and home range defense. American Journal of Primatology 78 : 523533.Google Scholar
Peres, C. A. 2000. Territorial defense and the ecology of group movements in small-bodied neotropical primates. Pages 100123 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
Potts, K. B., Baken, E., Levang, A., and Watts, D. P. 2016. Ecological factors influencing habitat use by chimpanzees at Ngogo, Kibale National Park, Uganda. American Journal of Primatology 78 : 432440.Google Scholar
Powzyk, J. A. 1997. The socio-ecology of two sympatric indriids: Propithecus diadema diadema and Indri indri, a comparison of feeding strategies and their possible repercussions on species-specific behaviors. PhD Thesis, Duke University.Google Scholar
Ries, L., Fletcher, R. J. Jr., Battin, J., and Sisk, T. D. 2004. Ecological responses to habitat edges: mechanisms, models, and variability explained. Annual Review of Ecology, Evolution and Systematics 35 : 491522.Google Scholar
Rode, E. J., Stengel, C. J., and Nekaris, K. A.-I. 2013. Habitat assessment and species niche modeling. Pages 79102 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
Ross, C. and Reeve, N. 2011. Survey and census methods: population distribution and density. Pages 111131 in Field and Laboratory Methods in Primatology: A Practical Guide. Setchell, J. M. and Curtis, D. J. (Eds.). Cambridge University Press, Cambridge.Google Scholar
Santhosh, K., Kumara, H. N., Velankar, A. D., and Sinha, A. 2015. Ranging behavior and resource use by lion-tailed macaques (Macaca silenus) in selectively logged forests. International Journal of Primatology 36 : 288310.Google Scholar
Schatz, G. E. 2001. Generic Tree Flora of Madagascar. Surrey: Cromwell Press.Google Scholar
Schoener, T. W. 1971. Theory of feeding strategies. Annual Review of Ecology and Systematics 2 : 369404.Google Scholar
Shaffer, C. A. 2013a. Ecological correlates of ranging behavior in bearded sakis (Chiropotes sagulatus) in a continuous forest in Guyana. International Journal of Primatology 34 : 515532.Google Scholar
Shaffer, C. A. 2013b. GIS analysis of patch use and group cohesiveness of bearded sakis (Chiropotes sagulatus) in the upper Essequibo Conservation Concession, Guyana. American Journal of Physical Anthropology 150 : 235246.Google Scholar
Singh, M., Cheyne, S. M., and Ehlers Smith, D. A. 2018. How conspecific primates use their habitats: surviving in an anthropogenically-disturbed forest in Central Kalimantan, Indonesia. Ecological Indicators 87 : 167177.Google Scholar
Steudel, K. 2000. The physiology and energetics of movement: effects on individuals 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
Struhsaker, T. 1981. Census methods for estimating densities. Pages 3680 in Techniques for the Study of Primate Population Ecology. National Research Council (Ed.). National Academy Press, Washington, DC.Google Scholar
Teichroeb, J. A. and Sicotte, P. 2009. Test of the ecological-constraints model on ursine colobus monkeys (Colobus vellerosus) in Ghana. American Journal of Primatology 71 : 4959.Google Scholar
Turk, D. 1995. A guide to trees of Ranomafana National Park and Central Eastern Madagascar. Unpublished manuscript.Google Scholar
Turton, S. M. and Freiburger, H. J. 1997. Edge and aspect effects on the microclimate of a small tropical forest remnant on the Atherton Tableland, Northeastern Australia. Pages 4554 in Tropical Forest Remnants: Ecology, Management, and Conservation of Fragmented Communities. Laurance, W. F. and Bierregaard, R. O. Jr. (Eds.). University of Chicago Press, Chicago, IL.Google Scholar
van Horne, B. 1983. Density as a misleading indicator of habitat quality. Journal of Wildlife Management 47 : 893901.Google Scholar
Worton, B. J. 1989. Kernel methods for estimating the utilization distribution in home-range studies. Ecology 70 : 164168.Google Scholar
Wright, P. C., Johnson, S. E., Irwin, M. T., et al. 2008. The crisis of the critically endangered greater bamboo lemur (Prolemur simus). Primate Conservation 23 : 517.Google Scholar
Xiang, Z. F., Huo, S., and Xiao, W. 2011. Habitat selection of black-and-white snub-nosed monkeys (Rhinopithecus bieti) in Tibet: implications for species conservation. American Journal of Primatology 73 : 347355.Google Scholar
Zeng, Y., Xu, J., Wang, Y., and Zhou, C. 2013. Habitat association and conservation implications of endangered Francois’ langur (Trachypithecus francoisi). PLoS ONE 8 : e75661.Google Scholar

References

Asensio, N., Brockelman, W. Y., Malaivijitnond, S., and Reichard, U. H. 2011. Gibbon travel paths are goal oriented. Animal Cognition 14(3): 395405.Google Scholar
Bates, L. A., and Byrne, R. W. 2009. Sex differences in the movement patterns of free-ranging chimpanzees (Pan troglodytes schweinfurthii): foraging and border checking. Behavioral Ecology and Sociobiology 64(2): 247255.Google Scholar
Bruggeman, J. E., Garrott, R. A., White, P. J., Watson, F. G. R., and Wallen, R. 2007. Covariates affecting spatial variability in bison travel behavior in Yellowstone National Park. Ecological Applications 17(5): 14111423.Google Scholar
Byrne, R. W., Noser, R., 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(3): 619631.Google Scholar
Cruse, H., and Wehner, R. 2011. No need for a cognitive map: decentralized memory for insect navigation. PLoS Computational Biology 7(3): e1002009.Google Scholar
Department of Natural Resources WA. 2004. Standards and Guidelines for Land Surveying Using Global Positioning System Methods. State Department of Natural Resources, Olympia, WAGoogle Scholar
Di Fiore, A., and Suarez, S. A. 2007. Route-based travel and shared routes in sympatric spider and woolly monkeys: cognitive and evolutionary implications. Animal Cognition 10(3): 317329.Google Scholar
Dolins, F. L. 2009. Captive cotton-top tamarins (Saguinus oedipus oedipus) use landmarks to localize hidden food items. American Journal of Primatology 71 : 316323.Google Scholar
Dolins, F. L., Klimowicz, C., Kelley, J., and Menzel, C. R. 2014. Using virtual reality to investigate comparative spatial cognitive abilities in chimpanzees and humans. American Journal of Primatology 76 : 496513.Google Scholar
Galdikas, B. M. F. 1988. Orangutan diet, range, and activity at Tanjung Puting, Central Borneo. International Journal of Primatology 9(1): 135.Google Scholar
Galdikas, B. M. F. and Vasey, P. 1992. Why are orangutans so smart? Ecological and social hypotheses. Pages 183224 in Social Processes and Mental Abilities in Non-Human Primates. Burton, F. D. (Ed.). The Edwin Mellen Press, Lewiston, NY.Google Scholar
Gallistel, C. R. and Cramer, A. E. 1996. Computations on metric maps in mammals: getting oriented and choosing a multi-destination route. Journal of Experimental Biology 199(1): 211217.Google Scholar
Hopkins, M. E. 2010. Mantled howler (Alouatta palliata) arboreal pathway networks: relative impacts of resource availability and forest structure. International Journal of Primatology 32(1): 238258.Google Scholar
Janmaat, K. R. L., Byrne, R. W., and Zuberbühler, K. 2006. Evidence for a spatial memory of fruiting states of rainforest trees in wild mangabeys. Animal Behaviour 72(4): 797807.Google Scholar
Janson, C. H. and Byrne, R. 2007. What wild primates know about resources: opening up the black box. Animal Cognition 10(3): 357367.Google Scholar
Knott, C., Beaudrot, L., Snaith, T., et al. 2008. Female–female competition in Bornean orangutans. International Journal of Primatology 29(4): 975997.Google Scholar
Leighton, M. 1993. Modeling dietary selectivity by Bornean orangutans: evidence for integration of multiple criteria in fruit selection. International Journal of Primatology 14(2): 257313.Google Scholar
Lührs, M.-L., Dammhahn, M., Kappeler, P. M., and Fichtel, C. 2009. Spatial memory in the grey mouse lemur (Microcebus murinus). Animal Cognition 12(4): 599609.Google Scholar
MacKinnon, J. 1974. The behaviour and ecology of wild orangutans. Animal Behaviour 22: 374.Google Scholar
Milton, K. 1981. Distribution patterns of tropical plant foods as an evolutionary stimulus to primate mental development. American Anthropologist 83: 534548.Google Scholar
Ministry of Environment BC. 2001. British Columbia Standards, Specifications and Guidelines for Resource Surveys Using Global Positioning System (GPS) Technology. Ministry of Environment, Lands and Parks, Victoria, BC.Google Scholar
Morrogh-Bernard, H., Husson, S. J., Knott, C. D., et al. 2009. Orangutan activity budgets and diet. Pages 119134 in Orangutans: Geographic Variation in Behavioural Ecology and Conservation. Wich, S. A., Utami Atmoko, S. S., Mitra Setia, T., and van Schaik, C. P. (Eds.). Oxford University Press, Oxford.Google Scholar
Normand, E. and Boesch, C. 2009. Sophisticated Euclidean maps in forest chimpanzees. Animal Behaviour 77(5): 11951201.Google Scholar
Noser, R. and Byrne, R. W. 2010. How do wild baboons (Papio ursinus) plan their routes? Travel among multiple high-quality food sources with inter-group competition. Animal Cognition 13(1): 145155.Google Scholar
Porter, L. M. and Garber, P. A. 2012. Foraging and spatial memory in wild Weddell’s saddleback tamarins (Saguinus fuscicollis weddelli) when moving between distant and out-of-sight goals. International Journal of Primatology 34(1): 3048.Google Scholar
Poucet, B. 1993. Spatial cognitive maps in animals: new hypotheses on their structure and neural mechanisms. Psychological Review 100(2): 163182.Google Scholar
Schwartz, B. L. and Evans, S. 2001. Episodic memory in primates. American Journal of Primatology 55(2): 7185.Google Scholar
Setiawan, A., Nugroho, T. S., and Pudyatmoko, S. 2009. A survey of Miller’s grizzled surili, Presbytis hosei canicrus, in East Kalimantan, Indonesia. Primate Conservation 24(1): 139143.Google Scholar
Singleton, I., Knott, C. D., Morrogh-Bernard, H. C., Wich, S. A., and van Schaik, C. P. 2009. Ranging behavior of orangutan females and social organization. Pages 205214 in Orangutans: Geographic Variation in Behavioural Ecology and Conservation. Wich, S. A., Utami Atmoko, S. S., Mitra Setia, T., and van Schaik, C. P. (Eds.). Oxford University Press, Oxford.Google Scholar
Squires, J. R., DeCesare, N. J., Olson, L. E., et al. 2013. Combining resource selection and movement behavior to predict corridors for Canada lynx at their southern range periphery. Biological Conservation 157: 187195.Google Scholar
Thorpe, S. K. and Crompton, R. H. 2009. Orangutan positional behaviour. Pages 3347 in Orangutans: Geographic Variation in Behavioural Ecology and Conservation. Wich, S. A., Utami Atmoko, S. S., Mitra Setia, T., and van Schaik, C. P. (Eds.). Oxford University Press, Oxford.Google Scholar

References

Adamescu, G. S., Plumptre, A. J., Abernethy, K. A., et al. 2018. Annual cycles are the most common reproductive strategy in African tropical tree communities. Biotropica 50(3): 418430.Google Scholar
Alverson, W., Moskovitz, D. K., Shopland, J. M. (Eds.). 2000. Rapid Biological Inventories for Conservation Action: Northwestern Pando, Bolivia. Field Museum, Chicago, IL.Google Scholar
Baayen, R. H. 2008. Analyzing Linguistic Data: A Practical Introduction to Statistics Using R. Cambridge University Press, Cambridge.Google Scholar
Ban, S. D., Boesch, C., and Janmaat, K. R. L. 2014. Taï chimpanzees anticipate revisiting high-valued fruit trees from further distances. Animal Cognition 17: 13531364.Google Scholar
Ban, S. D., Boesch, C., N’Guessan, A., et al. 2016. Taï chimpanzees change their travel direction for rare feeding trees providing fatty fruits. Animal Behaviour 118: 135147.Google Scholar
Barr, D. J., Levy, R., Scheepers, C., and Tily, H. J. 2013. Random effects structure for confirmatory hypothesis testing: keep it maximal. Journal of Memory and Language 68 : 255278.Google Scholar
Bates, D., Mächler, M., Bolker, B., and Walker, S. 2015. Fitting linear mixed-effects models using lme4. Journal of Statistical Software 67: 148.Google Scholar
Bertolani, M. P. 2013. Ranging and travelling patterns of wild chimpanzees at Kibale, Uganda: a GIS approach. Doctoral dissertation, University of Cambridge.Google Scholar
Boesch, C. and Boesch, A. 2000. The Chimpanzees of the Taï Forest: Behavioral Ecology and Evolution. Oxford University Press, Oxford.Google Scholar
Di Fiore, A. and Suarez, S. A. 2007. Route-based travel and shared routes in sympatric spider and woolly monkeys: cognitive and evolutionary implications. Animal Cognition 10: 317329.Google Scholar
Digby, L. J., Ferrari, S. F., and Saltzman, W. 2011. The role of competition in cooperatively breeding species. Pages 91107 in Primates in Perspective, 2nd edition. Campbell, C. J., Fuentes, A., MacKinnon, K. C., Bearder, S., and Stumpf, R. M. (Eds.). Oxford University Press, New York.Google Scholar
Dobson, A. J. and Barnett, A. 2008. An Introduction to Generalized Linear Models. Chapman & Hall/CRC, Boca Raton, FL.Google Scholar
Garber, P. A. 1988. Foraging decisions during nectar feeding by tamarin monkeys (Saguinus mystax and Saguinus fuscicollis, Callitrichidae, Primates) in Amazonian Peru. Biotropica 20: 100106.Google Scholar
Garber, P. A. 1989. Role of spatial memory in primate foraging patterns: Saguinus mystax and Saguinus fuscicollis. American Journal of Primatology 19: 203216.Google Scholar
Garber, P. A. 2000. The ecology of group movement: evidence for the use of spatial, temporal, and social information in some primate foragers. Pages 261298 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
Garber, P. A. in press. Primate cognitive ecology: challenges and solutions to locating and acquiring resources in social foragers. In Primate Diet and Nutrition: Needing, Finding, and Using Food. Lambert, J. E. and Rothman, J. (Eds.). University of Chicago Press, Chicago, IL.Google Scholar
Garber, P. A., and Porter, L. M. 2010. The ecology of exudate feeding and exudate production in Saguinus and Callimico. Pages 89108 in The Evolution of Exudativory in Primates. Burrows, A. and Nash, L. (Eds). Springer, New York.Google Scholar
Garber, P. A., and Porter, L. M. 2014. Navigating in small-scale space: the role of landmarks and resource monitoring in understanding saddleback tamarin travel. American Journal of Primatology 76: 447459.Google Scholar
Garber, P.A., Porter, L.M., Spross, J., and Di Fiore, A. 2016. Tamarins: insights into monogamous and non-monogamous single female breeding systems. American Journal of Primatology 78: 298314.Google Scholar
Goné Bi, Z. B. 2007. Régime alimentaire des chimpanzés, distribution spatiale et phénologie des plantes dont les fruits sont consommés par les chimpanzés du Parc National de Taï, en Côte d’Ivoire. PhD Thesis, University of Cocody.Google Scholar
Hothorn, T. and Hornik, K. 2013. exactRankTests: exact distributions for rank and permutation tests. R package version 0.8-24. Available at: http://CRAN.R-project.org/package=exactRankTests.Google Scholar
Jang, H., Boesch, C., Mundry, R., Ban, S. D., and Janmaat, K. R. L. 2019. Travel linearity and speed of human foragers and chimpanzees during their daily search for food in tropical rainforests. Scientfic Reports 9: 11066.Google Scholar
Janmaat, K. R. L. and Chancellor, R. L. 2010. Exploring new areas: how important is long-term spatial memory for mangabey (Lophocebus albigena johnstonii) foraging efficiency? International Journal of Primatology 31: 863886.Google Scholar
Janmaat, K. R. L., Olupot, W., Chancellor, R. L., Arlet, M. E., and Waser, P. M. 2009. Long-term site fidelity and individual home range shifts in Lophocebus albigena. International Journal of Primatology 30: 443466.Google Scholar
Janmaat, K. R. L., Ban, S. D., and Boesch, C. 2013a. Chimpanzees use long-term spatial memory to monitor large fruit trees and remember feeding experiences across seasons. Animal Behaviour 86: 11831205.Google Scholar
Janmaat, K. R. L., Ban, S. D. and Boesch, C. 2013b. Taï chimpanzees use botanical skills to discover fruit: what we can learn from their mistakes. Animal Cognition 16: 851860.Google Scholar
Janmaat, K. R. L., Polansky, L., Ban, S. D., and Boesch, C. 2014. Wild chimpanzees plan their breakfast time, type, and location. Proceedings of the National Academy of Sciences of the United States of America 111: 1634316348.Google Scholar
Janmaat, K. R. L., Boesch, C., Byrne, R., et al. 2016. The spatio-temporal complexity of chimpanzee food: how cognitive adaptations can counteract the ephemeral nature of ripe fruit. American Journal of Primatology 78: 626645.Google Scholar
Kalan, A. K. and Boesch, C. 2015. Audience effects in chimpanzee food calls and their potential for recruiting others. Behavioral Ecology and Sociobiology 69: 17011712.Google Scholar
Koenig, W. D., Kelly, D., Sork, V. L., et al. 2003. Dissecting components of population-level variation in seed production and the evolution of masting behavior. Oikos 102: 581591.Google Scholar
Kouakou, C. Y., Boesch, C., and Kuehl, H. 2011. Identifying hotspots of chimpanzee group activity from transect surveys in Taï National Park, Côte d’Ivoire. Journal of Tropical Ecology 27: 621630.Google Scholar
Laden, G. T. 1992. Ethnoarchaeology and land use ecology of the Efe (pygmies) of the Ituri rain forest, Zaire: a behavioral ecological study of land use patterns and foraging behavior. PhD Thesis, Harvard University.Google Scholar
Lledo-Ferrer, Y., Pelaez, F., and Heymann, E. W. 2011. The equivocal relationship between territoriality and scent marking in wild saddleback tamarins (Saguinus fuscicollis). International Journal of Primatology 32: 974991.Google Scholar
Lopez-Rebellon, N. 2010. Sex and age-based differences in antipredator behavior in saddle-back tamarins (Saguinus fuscicollis) in the department of Pando, Bolivia. MA Thesis, DeKalb, Northern Illinois University.Google Scholar
Martin, P. and Bateson, P. 1993. Measuring Behavior: An Introductory Guide, 2nd edition. Cambridge University Press, Cambridge.Google Scholar
Milton, K. 1980. The Foraging Strategy of Howler Monkeys: A Study in Primate Economics. Columbia University Press, New York.Google Scholar
Nishida, T., Zamma, K., Matsusaka, T., Inaba, A., and McGrew, W. C. 2010. Chimpanzee Behavior in the Wild: An Audio-Visual Encyclopedia. Springer, Tokyo.Google Scholar
Normand, E., Ban, S. D., and Boesch, C. 2009. Forest chimpanzees (Pan troglodytes verus) remember the location of numerous fruit trees. Animal Cognition 12(6): 797807.Google Scholar
Noser, R. and Byrne, R. W. 2015. Wild chacma baboons (Papio ursinus) remember single foraging episodes. Animal Cognition 18, 921929.Google Scholar
Peñuelas, J., Filella, I., Xiaoyang, Z., et al. 2004. Complex spatiotemporal phenological shifts as a response to rainfall changes. New Phytologist 161: 837846.Google Scholar
Polansky, L. and Boesch, C. 2013. Long-term fruit phenology and rainfall trends conflict in a West African lowland tropical rainforest. Journal of Tropical Ecology 45: 409535.Google Scholar
Porter, L. M. 2001. Dietary differences among sympatric Callitrichinae in northern Bolivia: Callimico goeldii, Saguinus fuscicollis and S. labiatus. International Journal of Primatology 22: 961992.Google Scholar
Porter, L. M. and Garber, P. A. 2013. Foraging and spatial memory in wild Weddell’s saddleback tamarins (Saguinus fuscicollis weddelli) when moving between distant and out-of-sight goals. International Journal of Primatology 34: 3048.Google Scholar
Porter, L. M., Sterr, S. M., and Garber, P. A. 2007. Habitat use and ranging behavior of Callimico goeldii. International Journal of Primatology 28: 10351058.Google Scholar
Porter, L. M., Erb, W. M., Saire, C. L., et al. 2015. Temporal changes in the composition of saddleback tamarin (Saguinus fuscicollis weddelli) groups over time. American Association of Physical Anthropologists 156(60): 255256.Google Scholar
Rakoczy, H., Clüver, A., Saucke, L., et al. 2014. Apes are intuitive statisticians. Cognition 131(1): 6068.Google Scholar
Rylands, A., Heymann, E., Matauschek, C., et al. 2016. Taxonomic review of the New World tamarins (Primates: Callitrichidae). Zoological Journal of the Linnean Society 177 (4): 10031028.Google Scholar
Schel, A. M., Machanda, Z., Townsend, S. W., Zuberbühler, K., and Slocombe, K. E. 2013. Chimpanzee food calls are directed at specific individuals. Animal Behaviour 86: 955965.Google Scholar
Stumpf, R. M. 2011. Chimpanzees and bonobos: inter- and intraspecies diversity. Pages 340356 in Primates in Perspective, 2nd edition. Campbell, C. J., Fuentes, A., MacKinnon, K. C., Bearder, S., and Stumpf, R. M. (Eds.) Oxford University Press, New York.Google Scholar
Tecwyn, E. C., Denison, S., Messer, E. J., and Buchsbaum, D. 2017. Intuitive probabilistic inference in capuchin monkeys. Animal Cognition 20(2): 243256.Google Scholar
Terborgh, J. 1983. Five New World Primates: A Study in Comparative Ecology. Princeton University Press, Princeton, NJ.Google Scholar

References

Al-Safadi, M. M. 1994. The hamadryas baboon, Papio hamadryas (Linnaeus, 1758) in Yemen (Mammalia: Primates: Cercopithecidae). Zoology in the Middle East 10: 516.Google Scholar
Altmann, S. A. and Altmann, J. 1970. Baboon Ecology. University of Chicago Press, Chicago, IL.Google Scholar
Atkinson, R. P. D., Rhodes, C. J., Macdonald, D. W., and Anderson, R. M. 2002. Scale-free dynamics in the movement patterns of jackals. Oikos 98: 134140.Google Scholar
Austin, D., Bowen, W. D., and McMillan, J. I. 2004. Intraspecific variation in movement patterns: modeling individual behaviour in a large marine predator. Oikos 105: 1530.Google Scholar
Bailey, H. and Thompson, P. 2006. Quantitative analysis of bottlenose dolphin movement patterns and their relationship with foraging. Journal of Animal Ecology 75: 456465.Google Scholar
Bartumeus, F. 2007. Levy processes in animal movement: an evolutionary hypothesis. Fractals: Complex Geometry Patterns and Scaling in Nature and Society 15: 151162.Google Scholar
Bartumeus, F., Peters, F., Pueyo, S., Marrase, C., and Catalan, J. 2003. Helical Levy walks: adjusting searching statistics to resource availability in microzooplankton. Proceedings of the National Academy of Sciences of the United States of America 100: 1277112775.Google Scholar
Benedix, J. H. 1993. Area-restricted search by the plains pocket gopher (Geomys bursarius) in tallgrass prairie habitat. Behavioral Ecology 4: 318324.Google Scholar
Benhamou, S. 1992. Efficiency of area-concentrated searching behavior in a continuous patchy environment. Journal of Theoretical Biology 159: 6781.Google Scholar
Benhamou, S. 2007. How many animals really do the Lévy walk? Ecology 88: 19621969.Google Scholar
Berg, H. C. 1983. Random Walks in Biology. Princeton University Press, Princeton, NJ.Google Scholar
Bergman, C. M., Schaefer, J. A., and Luttich, S. N. 2000. Caribou movement as a correlated random walk. Oecologia 123: 364374.Google Scholar
Biquand, S., Biquand-Guyot, V., Boug, A., and Gautier, J.-P. 1992. Group composition in wild and commensal hamadryas baboons: a comparative study in Saudi Arabia. International Journal of Primatology 13: 533543.Google Scholar
Blackwell, P. G. 1997. Random diffusion models for animal movement. Ecological Modelling 100: 87102.Google Scholar
Bond, A. B. 1980. Optimal foraging in a uniform habitat: search mechanism of the green lacewing. Animal Behaviour 28: 1019.Google Scholar
Bovet, P. and Benhamou, S. 1988. Spatial analysis of animals’ movements using a correlated random walk model. Journal of Theoretical Biology 131: 419433.Google Scholar
Boyer, D., Ramos-Fernández, G., Miramontes, O., et al. 2006. Scale-free foraging by primates emerges from their interaction with a complex environment. Proceedings of the Royal Society B: Biological Sciences 273: 17431750.Google Scholar
Brantingham, P. J. 2006. Measuring forager mobility. Current Anthropology 47: 435459.Google Scholar
Brockmann, D., Hufnagel, L., and Geisel, T. 2006. The scaling laws of human travel. Nature 439: 462465.Google Scholar
Brown, C. T., Liebovitch, L. S., and Glendon, R. 2007. Lévy flights in Dobe Ju/’hoansi foraging patterns. Human Ecology 35: 129138.Google Scholar
Brown, J. H., Gupta, V. K., Li, B.-L., et al. 2002. The fractal nature of nature: power laws, ecological complexity and biodiversity. Philosophical Transactions of the Royal Society B: Biological Sciences 357: 619626.Google Scholar
Byrne, R. W., Noser, R., 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
Chapleau, F., Johansen, P. H., and Williamson, M. 1988. The use and abuse of the term strategy. Oikos 53: 136138.Google Scholar
Clauset, A., Shalizi, C. R., and Newman, M. E. J. 2009. Power-law distributions in empirical data. SIAM Review 51: 661703.Google Scholar
Condit, R., Ashton, P. S., Baker, P., et al. 2000. Spatial patterns in the distribution of tropical tree species. Science 288: 14141418.Google Scholar
Crist, T. O., Guertin, D. S., Wiens, J. A., and Milne, B. T. 1992. Animal movement in heterogeneous landscapes: an experiment with Eleodes beetles in shortgrass prairie. Functional Ecology 6: 536544.Google Scholar
Curio, E. 1976. The Ethology of Predation. Springer, New York.Google Scholar
Dai, X., Shannon, G., Slotow, R., Page, B., and Duffy, K. J. 2007. Short-duration daytime movements of a cow herd of African elephants. Journal of Mammalogy 88: 151157.Google Scholar
de Jager, N. R. and Rohweder, J. J. 2011. Spatial scaling of core and dominant forest cover in the Upper Mississippi and Illinois River floodplains, USA. Landscape Ecology 26: 697708.Google Scholar
de Knegt, H. J., Hengeveld, G. M., van Langevelde, F., et al. 2007. Patch density determines movement patterns and foraging efficiency of large herbivores. Behavioral Ecology 18: 10651072.Google Scholar
Edwards, A. M., Phillips, R. A., Watkins, N. W., et al. 2007. Revisiting Lévy flight search patterns of wandering albatrosses, bumblebees and deer. Nature 449: 10441045.Google Scholar
Faugeras, B. and Maury, O. 2007. Modeling fish population movements: from an individual-based representation to an advection–diffusion equation. Journal of Theoretical Biology 247: 837848.Google Scholar
Ford, R. G. 1983. Home range in a patchy environment: optimal foraging predictions. American Zoologist 23: 315326.Google Scholar
Fritz, H., Said, S., and Weimerskirch, H. 2003. Scale-dependent hierarchical adjustments of movement patterns in a long-range foraging seabird. Proceedings of the Royal Society B: Biological Sciences 270: 11431148.Google Scholar
Garber, P. A. and Porter, L. 2014. Navigating in small-scale space: the role of landmarks and resource monitoring in understanding saddleback tamarin travel. American Journal of Primatology 76: 447459.Google Scholar
Gautestad, A. O. 2013. Animal space use: distinguishing a two-level superposition of scale-specific walks from scale-free Levy walk. Oikos 122: 612620.Google Scholar
Gautestad, A. O. and Mysterud, I. 2005. Intrinsic scaling complexity in animal dispersion and abundance. American Naturalist 165: 4455.Google Scholar
Gonzalez, M. C., Hidalgo, C. A., and Barabasi, A.-L. 2008. Understanding individual human mobility patterns. Nature 453: 779782.Google Scholar
Grove, M. 2010. The quantitative analysis of mobility: ecological techniques and archaeological extensions. Pages 83118 in New Perspectives on Old Stones: Analytical Approaches to Palaeolithic Technologies. Lycett, S. and Chuahan, P. (Eds.). Springer, Dordrecht.Google Scholar
Grove, M. 2013. The evolution of spatial memory. Mathematical Biosciences 242: 2532.Google Scholar
Grunbaum, D. 1998. Using spatially explicit models to characterize foraging performance in heterogeneous landscapes. American Naturalist 151: 97115.Google Scholar
Hills, T. T., Kalff, C., and Wiener, J. M. 2013. Adaptive Levy processes and area-restricted search in human foraging. PLoS ONE 8. DOI: 10.1371/journal.pone.0060488.Google Scholar
Janson, C. H. and DiBitetti, M. S. 1997. Experimental analysis of food detection in capuchin monkeys: effects of distance, travel speed, and resource size. Behavioral Ecology and Sociobiology 41: 1724.Google Scholar
Johnson, A. R., Milne, B. T., and Wiens, J. A. 1992. Diffusion in fractal landscapes: simulations and experimental studies of tenebrionid beetle movements. Ecology 73: 19681983.Google Scholar
Johnson, D. S., London, J. M., Lea, M.-A., and Durban, J. W. 2008. Continuous-time correlated random walk model for animal telemetry data. Ecology 89: 12081215.Google Scholar
Kareiva, P. and Odell, G. 1987. Swarms of predators exhibit preytaxis if individual predators use area-restricted search. American Naturalist 130: 233270.Google Scholar
Kareiva, P. M. and Shigesada, N. 1983. Analyzing insect movement as a correlated random walk. Oecologia 56: 234238.Google Scholar
Keasar, T., Shmida, A., and Motro, U. 1996. Innate movement rules in foraging bees: flight distances are affected by recent rewards and are correlated with choice of flower type. Behavioral Ecology and Sociobiology 39: 381388.Google Scholar
Knell, A. S. and Codling, E. A. 2012. Classifying area-restricted search (ARS) using a partial sum approach. Theoretical Ecology 5: 325339.Google Scholar
Kummer, H. 1968. Social Organization of Hamadryas Baboons. University of Chicago Press, Chicago, IL.Google Scholar
Laing, J. 1937. Host-finding by insect parasites: I. Observations on the finding of hosts by Alysia manducator, Mormoniella vitripennis and Trichogramma evanescens. Journal of Animal Ecology 6: 298317.Google Scholar
Laing, J. 1938. Host-finding by insect parasites: II. The chance of Trichogramma evanescens finding its hosts. Journal of Experimental Biology 15: 281302.Google Scholar
Levandowsky, M., Klafter, J., and White, B. S. 1988a. Feeding and swimming behavior in grazing microzooplankton. Journal of Protozoology 35: 243246.Google Scholar
Levandowsky, M., Klafter, J., and White, B. S. 1988b. Swimming behavior and chemosensory responses in the protistan microzooplankton as a function of the hydrodynamic regime. Bulletin of Marine Science 43: 758763.Google Scholar
Levandowsky, M., White, B. S., and Schuster, F. L. 1997. Random movements of soil amebas. Acta Protozoologica 36: 237248.Google Scholar
Lévy, P. 1937. Theorie de l’Addition des Variables Aleatoires. Paris: Gauthier-Villars.Google Scholar
Lode, T. 2000. Functional response and area-restricted search in a predator: seasonal exploitation of anurans by the European polecat, Mustela putorius. Austral Ecology 25: 223231.Google Scholar
Marell, A., Ball, J. P., and Hofgaard, A. 2002. Foraging and movement paths of female reindeer: insights from fractal analysis, correlated random walks, and Lévy flights. Canadian Journal of Zoology/Revue Canadienne De Zoologie 80: 854865.Google Scholar
Miramontes, O., DeSouza, O., Hernandez, D., and Ceccon, E. 2012. Non-Levy mobility patterns of Mexican Me’Phaa peasants searching for fuel wood. Human Ecology 40: 167174.Google Scholar
Newman, M. E. J. 2005. Power laws, Pareto distributions and Zipf’s law. Contemporary Physics 46: 323351.Google Scholar
Normand, E. and Boesch, C. 2009. Sophisticated Euclidean maps in forest chimpanzees. Animal Behaviour 77: 11951201.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. 2010. How do wild baboobs (Papio ursinus) plan their routes? Travel among multiple high-quality food sources with inter-group competition. Animal Cognition 13, 145155.Google Scholar
Patlak, C. S. 1953a. Random walk with persistence and external bias. Bulletin of Mathematical Biophysics 15: 311318.Google Scholar
Patlak, C. S. 1953b. A mathematical contribution to the study of orientation of organisms. Bulletin of Mathematical Biophysics 15: 431476.Google Scholar
Porter, L. and Garber, P. 2013. Foraging and spatial memory in wild Weddell’s saddleback tamarins (Saguinus fuscicollis weddelli) when moving between distant and out-of-sight goals. International Journal of Primatology 34: 3048.Google Scholar
Raichlen, D. A., Wood, B. M., Gordon, A. D., et al. 2014. Evidence of Lévy walk foraging patterns in human hunter-gatherers. Proceedings of the National Academy of Sciences USA 111(2): 728733.Google Scholar
Ramos-Fernández, G., Mateos, J. L., Miramontes, O., et al. 2004. Lévy walk patterns in the foraging movements of spider monkeys (Ateles geoffroyi). Behavioral Ecology and Sociobiology 55: 223230.Google Scholar
Reid, A. T. 1953. On stochastic processes in biology. Biometrics 9: 275289.Google Scholar
Reynolds, A. M., Reynolds, D. R., Smith, A. D., Svensson, G. P., and Lofstedt, C. 2007a. Appetitive flight patterns of male Agrotis segetum moths over landscape scales. Journal of Theoretical Biology 245: 141149.Google Scholar
Reynolds, A. M., Smith, A. D., Menzel, R., et al. 2007b. Displaced honey bees perform optimal scale-free search flights. Ecology 88: 19551961.Google Scholar
Reynolds, A. M., Smith, A. D., Reynolds, D. R., Carreck, N. L., and Osborne, J. L. 2007c. Honeybees perform optimal scale-free searching flights when attempting to locate a food source. Journal of Experimental Biology 210: 37633770.Google Scholar
Rhee, I., Shin, M., Hong, S., et al. 2011. On the Lévy-walk nature of human mobility. IEEE–ACM Transactions on Networking 19: 630643.Google Scholar
Schaefer, J. A., Bergman, C. M., and Luttich, S. N. 2000. Site fidelity of female caribou at multiple spatial scales. Landscape Ecology 15: 731739.Google Scholar
Schreier, A. L. 2009. The influence of resource distribution on the social structure and travel patterns of wild hamadryas baboons (Papio hamadryas) in Filoha, Awash National Park, Ethiopia. Dissertation, City University of New York.Google Scholar
Schreier, A. L. and Grove, M. 2010. Ranging patterns of hamadryas baboons: random walk analyses. Animal Behaviour 80: 7587.Google Scholar
Schreier, A. L. and Grove, M. 2014. Recurrent patterning in the daily foraging routes of hamadryas baboons (Papio hamadryas): spatial memory in large- versus small-scale space. American Journal of Primatology 76: 421435.Google Scholar
Schreier, A. and Swedell, L. 2008. Use of palm trees as a sleeping site for hamadryas baboons (Papio hamadryas) in Ethiopia. American Journal of Primatology 70: 107113.Google Scholar
Schreier, A. L. and Swedell, L. 2009. The fourth level of social structure in a multi-level society: ecological and social functions of clans in hamadryas baboons. American Journal of Primatology 71: 948955.Google Scholar
Schreier, A. L. and Swedell, L. 2012a. The socioecology of network scaling ratios in the multilevel society of hamadryas baboons. International Journal of Primatology 33: 10691080.Google Scholar
Schreier, A. L. and Swedell, L. 2012b. Ecology and sociality in a multilevel society: ecological determinants of social cohesion in hamadryas baboons. American Journal of Physical Anthropology 148: 580588.Google Scholar
Shettleworth, S. J. 1998. Cognition, Evolution, and Behavior. Oxford University Press, New York.Google Scholar
Shlesinger, M. F. and Klafter, J. 1986. Lévy walks versus Lévy flights. Pages 279283 in On Growth and Form. Stanley, H. E. and Ostrowski, N. (Eds.). Martinus Nijhof, Amsterdam.Google Scholar
Shlesinger, M. F., Zaslavsky, G. M., and Klafter, J. 1993. Strange kinetics. Nature 363: 3137.Google Scholar
Sigg, H. 1986. Ranging patterns in hamadryas baboons: evidence for a mental map. Pages 8791 in Primate Ontogeny, Cognition, and Social Behaviour, Vol. 3, Else, J. G. and Lee, P. C. (Eds.). Cambridge University Press, Cambridge.Google Scholar
Sigg, H. and Stolba, A. 1981. Home range and daily march in a hamadryas baboon troop. Folia Primatologica 26: 4075.Google Scholar
Sims, D. W., Righton, D., and Pitchford, J. W. 2007. Minimizing errors in identifying Lévy flight behaviour of organisms. Journal of Animal Ecology 76: 222229.Google Scholar
Sims, D. W., Southall, E. J., Humphries, N. E., et al. 2008. Scaling laws of marine predator search behaviour. Nature 451: 10981102.Google Scholar
Skellam, J. G. 1951. Random dispersal in theoretical populations. Biometrika 38: 196218.Google Scholar
Skellam, J. G. 1973. The formulation and interpretation of mathematical models of diffusionary processes in population biology. Pages 6385 in The Mathematical Theory of the Dynamics of Biological Populations. Bartlett, M. S. and Hiorns, R. W. (Eds.). Academic Press, New York.Google Scholar
Sueur, C. 2011. A non-Lévy random walk in chacma baboons: what does it mean? PLoS ONE 6. DOI: 10.1371/journal.pone.0016131.Google Scholar
Sueur, C., Briard, L., and Petit, O. 2011. Individual analyses of Lévy walk in semi-free ranging tonkean macaques (Macaca tonkeana). PLoS ONE 6. DOI: 10.1371/journal.pone.0026788.Google Scholar
Swedell, L. 2000. Two takeovers in wild hamadryas baboons. Folia Primatologica 71: 169172.Google Scholar
Swedell, L. 2002a. Affiliation among females in wild hamadryas baboons (Papio hamadryas hamadryas). International Journal of Primatology 23: 12051226.Google Scholar
Swedell, L. 2002b. Ranging behavior, group size and behavioral flexibility in Ethiopian hamadryas baboons (Papio hamadryas hamadryas). Folia Primatologica 73: 95103.Google Scholar
Swedell, L. 2006. Strategies of Sex and Survival in Hamadryas Baboons: Through a Female Lens. Prentice Hall, Upper Saddle River, NJ.Google Scholar
Swedell, L. and Schreier, A. 2009. Male aggression towards females in hamadryas baboons: conditioning, coercion, and control. Pages 244268 in Sexual Coercion in Primates: An Evolutionary Perspective on Male Aggression Against Females. Muller, M. N. and Wrangham, R. (Eds.). Harvard University Press, Cambridge, MA.Google Scholar
Swedell, L., Hailemeskel, G., and Schreier, A. 2008. Composition and seasonality of diet in adult male hamadryas baboons: preliminary findings from Filoha. Folia Primatologica 79: 476490.Google Scholar
Swedell, L., Saunders, J., Schreier, A., et al. 2011. Female “dispersal” in hamadryas baboons: transfer among social units in a multi-level society. American Journal of Physical Anthropology 145: 360370.Google Scholar
Turchin, P. 1998. Quantitative Analysis of Movement. Sinauer Associates, Sunderland, MA.Google Scholar
Viswanathan, G. M., Afanasyev, V., Buldyrev, S. V., et al. 1996. Levy flight search patterns of wandering albatrosses. Nature 381: 413415.Google Scholar
Viswanathan, G. M., Bartumeus, F., Buldyrev, S. V., et al. 2002. Levy flight random searches in biological phenomena. Physica A: Statistical Mechanics and Its Applications 314: 208213.Google Scholar
Viswanathan, G. M., Buldyrev, S. V., Havlin, S., et al. 1999. Optimizing the success of random searches. Nature 401: 911914.Google Scholar
Viswanathan, G. M., Raposo, E. P., and da Luz, M. G. E. 2008. Levy flights and superdiffusion in the context of biological encounters and random searches. Physics of Life Reviews 5: 133150.Google Scholar
Ward, D. and Saltz, D. 1994. Foraging at different spatial scales: dorcas gazelles foraging for lilies in the Negev Desert. Ecology 75: 4858.Google Scholar
Zinner, D., Pelaez, F., and Torkler, F. 2001. Distribution and habitat associations of baboons (Papio hamadryas) in central Eritrea. International Journal of Primatology 22: 397413.Google Scholar

References

Amsler, S. J. 2010. Energetic costs of territorial boundary patrols by wild chimpanzees. American Journal of Primatology 72(2): 93103.Google Scholar
Bates, L. A., and Byrne, R. W. 2009. Sex differences in the movement patterns of free-ranging chimpanzees (Pan troglodytes schweinfurthii): foraging and border checking. Behavioral Ecology and Sociobiology 64(2): 247255.Google Scholar
Cachat, J. M., Stewart, A., Utterback, E., et al. 2011. Three-dimensional neurophenotyping of adult zebrafish behaviorPloS One 6(3): e17597.Google Scholar
Chapman, C. A. and Pavelka, M. S. 2005. Group size in folivorous primates: ecological constraints and the possible influence of social factors. Primates 46(1): 19.Google Scholar
Clutton‐Brock, T. H. and Harvey, P. H. 1977. Primate ecology and social organization. Journal of Zoology 183(1): 139.Google 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
Dickens, M. 1955. A statistical formula to quantify the “spread‐of‐participation” in group discussion. Speech Monographs 22: 2831.Google Scholar
Ghadar, N., Zhang, X., Li, K., et al. 2013. Visual hull reconstruction for automated primate behavior observation. Pages 16 in 2013 IEEE International Workshop on Machine Learning for Signal Processing (MLSP). IEEE, Piscataway, NJ.Google Scholar
Gomez-Marin, A., Partoune, N., Stephens, G. J., and Louis, M. 2012. Automated tracking of animal posture and movement during exploration and sensory orientation behaviors. PLoS ONE 7(8): e41642.Google Scholar
Janmaat, K. R., 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(6): 11831205.Google Scholar
Khan, Z., Herman, R. A., Wallen, K., and Balch, T. 2005. An outdoor 3-D visual tracking system for the study of spatial navigation and memory in rhesus monkeys. Behavior Research Methods 37(3): 453463.Google Scholar
Kurtycz, L. M., Shender, M. A., and Ross, S. R. 2014. The birth of an infant decreases group spacing in a zoo-housed lowland gorilla group (Gorilla gorilla gorilla). Zoo Biology 33: 471474.Google Scholar
Landt, J. 2005. The history of RFID. IEEE Potentials 24(4): 811.Google Scholar
Lipton, A. 2012. RFID goes on Safari. RFiD Journal 9(3).Google Scholar
Maple, T. L. and Finlay, T. W. 1987. Post-occupancy evaluation in the zoo. Applied Animal Behaviour Science 18(1): 518.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(5): 495499.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
Newton‐Fisher, N. E. 2003. The home range of the Sonso community of chimpanzees from the Budongo Forest, Uganda. African Journal of Ecology 41(2): 150156.Google Scholar
Ng, M. L., Leong, K. S., Hall, D. M., and Cole, P. H. 2005. A small passive UHF RFID tag for livestock identification. Pages 6770 in MAPE 2005. IEEE International Symposium on Microwave, Antenna, Propagation and EMC Technologies for Wireless Communications, Vol. 1. IEEE, Piscataway, NJ.Google Scholar
Ogden, J. J., Lindburg, D. G., and Maple, T. L. 1993. Preference for structural environmental features in captive lowland gorillas (Gorilla gorilla gorilla). Zoo Biology 12(4): 381395.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(2): 167172.Google Scholar
Radespiel, U. 2007. Ecological diversity and seasonal adaptations of mouse lemurs (Microcebus spp.). Pages 211234 in Lemurs. Gould, L. and Sauther, M. (Eds.). Springer, New York.Google Scholar
Reutebuch, S. E., McGaughey, R. J., Andersen, H. E., and Carson, W. W. 2003. Accuracy of a high-resolution lidar terrain model under a conifer forest canopy. Canadian Journal of Remote Sensing 29(5): 527535.Google Scholar
Roberts, M. and Kohn, F. 1993. Habitat use, foraging behavior, and activity patterns in reproducing western tarsiers, Tarsius bancanus, in captivity: a management synthesis. Zoo Biology 12(2): 217232.Google Scholar
Ross, S. R. and Lukas, K. E. 2000. Conducting a post-occupancy evaluation as part of the design process for a new great ape facility. Pages 140–141 in The Apes: Challenges for the 21st Century, Conference Proceedings, May 10–13, 2000, Brookfield, IL.Google Scholar
Ross, S. R., and Lukas, K. E. 2006. Use of space in a non-naturalistic environment by chimpanzees (Pan troglodytes) and lowland gorillas (Gorilla gorilla gorilla). Applied Animal Behaviour Science 96(1): 143152.Google Scholar
Ross, S. R., Schapiro, S. J., Hau, J., and Lukas, K. E. 2009. Space use as an indicator of enclosure appropriateness: a novel measure of captive animal welfare. Applied Animal Behaviour Science 121(1): 4250.Google Scholar
Ross, S. R., Calcutt, S., Schapiro, S. J., and Hau, J. 2011a. Space use selectivity by chimpanzees and gorillas in an indoor–outdoor enclosure. American Journal of Primatology 73(2): 197208.Google Scholar
Ross, S. R., Wagner, K. E., Schapiro, S. J., and Hau, J. 2011b. Transfer and acclimation effects on the behavior of two species of African great apes moved to a novel and naturalistic zoo environment. International Journal of Primatology 32: 99117.Google Scholar
Shender, M. A. and Ross, S. R. 2013. The effects of available space on a group of captive gorillas (Gorilla gorilla gorilla) and chimpanzees (Pan troglodytes) living in a naturalistic zoo-setting. American Journal of Primatology 75(S1): 51.Google Scholar
Shender, M. A., Anderson, K. E., and Ross, S. R. 2012. Group cohesion in captive gorillas (Gorilla gorilla gorilla) and chimpanzees (Pan troglodytes) living in a naturalistic zoo exhibit. American Journal of Primatology 74(S1): 75.Google Scholar
Spink, A. J., Tegelenbosch, R. A. J., Buma, M. O. S., and Noldus, L. P. J. J. 2001. The EthoVision video tracking system: a tool for behavioral phenotyping of transgenic mice. Physiology & Behavior 73(5): 731744.Google Scholar
Spitzen, J., Spoor, C. W., Grieco, F., et al. 2013. A 3D analysis of flight behavior of Anopheles gambiae sensu stricto malaria mosquitoes in response to human odor and heat. PLoS ONE 8(5): e62995.Google Scholar
Trayford, H. R. and Farmer, K. H. 2012. An assessment of the use of telemetry for primate reintroductions. Journal for Nature Conservation 20: 311325.Google Scholar
Traylor‐Holzer, K. and Fritz, P. 1985. Utilization of space by adult and juvenile groups of captive chimpanzees (Pan troglodytes). Zoo Biology 4(2): 115127.Google Scholar
Vanderploeg, H. A. and Scavia, D. 1979. Two electivity indices for feeding with special reference to zooplankton grazing. Journal of the Fisheries Board of Canada 36(4): 362365.Google Scholar

References

Albeke, S. E., Nibbelink, N. P., and Ben-David, M. 2015. Modeling behavior by coastal river otter (Lontra canadensis) in response to prey availability in Prince William Sound, Alaska: a spatially-explicit individual-based approach. PLoS ONE 10: e0126208.Google Scholar
Allan, R. J. 2010. Survey of agent based modelling and simulation tools. Technical Report DL-TR-2010-007. Science and Technology Facilities Council.Google Scholar
An, G., and Wilensky, U. 2009. From artificial life to in silico medicine: NetLogo as a means of translational knowledge representation in biomedical research. Pages 183214 in Artificial Life Models in Software. Komosinski, M. and Adamatzky, A. (Eds.). Springer-Verlag, London.Google Scholar
Axelrod, R. M. 1997. The Complexity of Cooperation: Agent-Based Models of Competition and Collaboration. Princeton University Press, Princeton, NJ.Google Scholar
Axtell, R. L., Epstein, J. M., Dean, J. S., et al. 2002. Population growth and collapse in a multi-agent model of the Kayenta Anasazi in Long House Valley. Proceedings of the National Academy of Sciences USA 99: 72757279.Google Scholar
Baggio, J. A., Salau, K., Janssen, M. A., Schoon, M. L., and Bodin, Ö. 2011. Landscape connectivity and predator–prey population dynamics. Landscape Ecology 26: 3345.Google Scholar
Balbi, S. and Giupponi, C. 2009. Reviewing agent-based modelling of socio-ecosystems: a methodology for the analysis of climate change adaptation and sustainability. Working Paper 15. Department of Economics, Ca’ Foscari University of Venice.Google Scholar
Barton, C. M. 2008. Patch choice model from optimal foraging theory (Human Behavioral Ecology) (Version 1). CoMSES Computational Model Library. Available at: www.openabm.org/model/2221/version/1.Google Scholar
Barton, C. M. 2015. Diet breadth model from optimal foraging theory (Human Behavioral Ecology) (Version 2). CoMSES Computational Model Library. Available at: www.openabm.org/model/2225/version/2.Google Scholar
Barton, C. M. and Riel-Salvatore, J. 2012. Agents of change: modeling biocultural evolution in Upper Pleistocene western Eurasia. Advances in Complex Systems 15: 1150003.Google Scholar
Barton, C. M., Riel-Salvatore, J., Anderies, J. M., and Popescu, G. 2011. Modeling human ecodynamics and biocultural interactions in the Late Pleistocene of western Eurasia. Human Ecology 39: 705725.Google Scholar
Batty, M. 2005. Cities and Complexity: Understanding Cities with Cellular Automata, Agent-Based Models, and Fractals. MIT Press, Cambridge, MA.Google Scholar
Berger, T. and Troost, C. 2014. Agent-based modelling of climate adaptation and mitigation options in agriculture. Journal of Agricultural Economics 65: 323348.Google Scholar
Berryman, M. 2008. Review of software platforms for agent based models. Report DSTO-GD-0532. Australian Government Department of Defense, Land Operations Division, Defence Science and Technology Organisation.Google Scholar
Bharwani, S., Bithell, M., Downing, T. E., et al. 2005. Multi-agent modelling of climate outlooks and food security on a community garden scheme in Limpopo, South Africa. Philosophical Transactions of the Royal Society B: Biological Sciences 360: 21832194.Google Scholar
Billari, F. C., Fent, T., Prskawetz, A., Scheffran, J. 2006. Agent-based computational modelling: an introduction. Pages 116 in Agent-Based Computational Modelling: Applications in Demography, Social, Economic and Environmental Sciences. Billari, F. C., Fent, T., Prskawetz, A., and Scheffran, J. (Eds.). Physica-Verlag, Heidelberg.Google Scholar
Bithell, M. and Brasington, J. 2009. Coupling agent-based models of subsistence farming with individual-based forest models and dynamic models of water distribution. Environmental Modelling & Software 24: 173190.Google Scholar
Bonabeau, E. 2002. Agent-based modeling: methods and techniques for simulating human systems. Proceedings of the National Academy of Sciences USA 99(S3): 72807287.Google Scholar
Bonnell, T. R., Sengupta, R. R., Chapman, C. A., and Goldberg, T. L. 2010. An agent-based model of red colobus resources and disease dynamics implicates key resource sites as hot spots of disease transmission. Ecological Modelling 221: 24912500.Google Scholar
Bonnell, T. R., Campennì, M., Chapman, C. A., et al. 2013. Emergent group level navigation: an agent-based evaluation of movement patterns in a folivorous primate. PLoS ONE 8: e78264.Google Scholar
Bousquet, F. and Le Page, C. 2004. Multi-agent simulations and ecosystem management: a review. Ecological Modelling 176: 313332.Google Scholar
Brown, D. G., Riolo, R., Robinson, D. T., North, M., and Rand, W. 2005. Spatial process and data models: toward integration of agent-based models and GIS. Journal of Geographical Systems 7: 2547.Google Scholar
Bryson, J. J., Ando, Y., and Lehmann, H. 2007. Agent-based modelling as scientific method: a case study analysing primate social behaviour. Philosophical Transactions of the Royal Society B: Biological Sciences 362: 16851699.Google Scholar
Chiacchio, F., Pennisi, M., Russo, G., Motta, S., and Pappalardo, F. 2014. Agent-based modeling of the immune system: NetLogo, a promising framework. BioMed Research International 2014: 16.Google Scholar
Conradt, L., Zollner, P., Roper, T., Frank, K., and Thomas, C. 2003. Foray search: an effective systematic dispersal strategy in fragmented landscapes. The American Naturalist 161: 905915.Google Scholar
Couzin, I. D., Krause, J., James, R., Ruxton, G. D., and Franks, N. R. 2002. Collective memory and spatial sorting in animal groups. Journal of Theoretical Biology 218: 111.Google Scholar
Crooks, A. T. and Castle, C. J. E. 2012. The integration of agent-based modelling and geographical information for geospatial simulation. Pages 219251 in Agent-Based Models of Geographical Systems. Springer, Dordrecht.Google Scholar
Crooks, A. T. and Heppenstall, A. J. 2012. Introduction to agent-based modelling. Pages 85–105 in Agent-Based Models of Geographical Systems. Springer, Dordrecht.Google Scholar
DeAngelis, D. L. and Mooij, W. M. 2005. Individual-based modeling of ecological and evolutionary processes. Annual Review of Ecology and Systematics 36: 147168.Google Scholar
Di Fiore, A. 2009. Agent-based simulation modeling of primate sociality. American Journal of Physical Anthropology Supplement 48: 118119.Google Scholar
Di Fiore, A. 2010a. Forward-time, individual-based simulations and their use in primate landscape genetics. 23rd Congress of the International Primatological Society, August 2010, Kyoto, Japan.Google Scholar
Di Fiore, A. 2010b. The influence of social systems on primate population genetic structure: an agent-based modeling approach. SOCIOR Conference on Social Systems: Demographic and Genetic Issues, September 2010, University of Rennes, Paimpont, France.Google Scholar
Di Fiore, A. 2012. The interplay between primate social organization and population genetic structure: insights from agent-based simulation models. 24th Congress of the International Primatological Society, August 2012, Cancun, Mexico.Google Scholar
Di Fiore, A. and Valencia, L. M. 2013. Monogamy, polygyny, and polyandry: exploring the genetic consequences of callitrichine social systems using agent-based simulation. 36th Annual Meeting of the American Society of Primatologists, June 2013, San Juan, Puerto Rico.Google Scholar
Di Fiore, A. and Valencia, L. M. 2014. The interplay of landscape features and social system on the genetic structure of a primate population: a simulation study using “tamarin” monkeys. International Journal of Primatology 35: 226257.Google Scholar
Dolado, R. and Beltran, F. S. 2012. Emergent patterns of social organization in captive Cercocebus torquatus: testing the GrooFiWorld agent-based model. Journal of Biosciences 37: 777784.Google Scholar
Dumont, B. and Hill, D. R. C. 2004. Spatially explicit models of group foraging by herbivores: what can agent-based models offer? Animal Research 53: 419428.Google Scholar
Epstein, J. M. 1999. Agent-based computational models and generative social science. Complexity 5: 4160.Google Scholar
Epstein, J. M. 2007. Generative Social Science: Studies in Agent-Based Computational Modeling. Princeton University Press, Princeton, NJ.Google Scholar
Epstein, J. M. and Axtell, R. 1996. Growing Artificial Societies: Social Science from the Bottom Up. MIT Press, Cambridge, MA.Google Scholar
Evers, E., de Vries, H., Spruijt, B. M., and Sterck, E. H. M. 2011. Better safe than sorry: socio-spatial group structure emerges from individual variation in fleeing, avoidance or velocity in an agent-based model. PLoS ONE 6: e26189.Google Scholar
Evers, E., de Vries, H., Spruijt, B. M., and Sterck, E. H. M. 2012. Look before you leap: individual variation in social vigilance shapes socio-spatial group properties in an agent-based model. Behavioural Ecology and Sociobiology 66: 931945.Google Scholar
Evers, E., de Vries, H., Spruijt, B. M., and Sterck, E. H. M. 2014. The EMO-model: an agent-based model of primate social behavior regulated by two emotional dimensions, anxiety-FEAR and satisfaction-LIKE. PLoS ONE 9: e87955.Google Scholar
Gilbert, N. 2008. The idea of agent-based modeling. Pages 120 in Agent-Based Models. Sage, Thousand Oaks, CA.Google Scholar
Gilbert, N. and Terna, P. 2000. How to build and use agent-based models in social science. Mind & Society 2000: 5772.Google Scholar
Gilbert, N. and Troitzsch, K. G. 2005. Simulation for the Social Scientist, 2nd edition. Open University Press (McGraw-Hill Education), Glasgow.Google Scholar
Goldstone, R. L. and Janssen, M. A. 2005. Computational models of collective behavior. Trends in Cognitive Sciences 9: 424430.Google Scholar
Griffith, C., Long, B., and Sept, J. 2010. HOMINIDS: an agent-based spatial simulation model to evaluate behavioral patterns of early Pleistocene hominids. Ecological Modelling 221: 738760.Google Scholar
Grignard, A., Taillandier, P., Gaudou, B., et al. 2013. GAMA 1.6: Advancing the art of complex agent-based modeling and simulation. Pages 117131 in PRIMA 2013: Principles and Practice of Multi-Agent Systems. Springer, New York.Google Scholar
Grimm, V. and Railsback, S. F. 2005. Individual-Based Modeling and Ecology. Princeton University Press, Princeton, NJ.Google Scholar
Grimm, V., Berger, U., Bastiansen, F., et al. 2006. A standard protocol for describing individual-based and agent-based models. Ecological Modelling 198: 115126.Google Scholar
Grimm, V., Berger, U., DeAngelis, D. L., et al. 2010. The ODD protocol: a review and first update. Ecological Modelling 221: 27602768.Google Scholar
Hashemi, M. and Alesheikh, A. A. 2013. GIS: agent-based modeling and evaluation of an earthquake-stricken area with a case study in Tehran, Iran. Natural Hazards 69: 18951917.Google Scholar
Heckbert, S. 2013. MayaSim: An agent-based model of the ancient Maya social-ecological system. Journal of Artificial Societies and Social Simulation 16: 11.Google Scholar
Hemelrijk, C. K. 1999a. An individual-orientated model of the emergence of despotic and egalitarian societies. Proceedings of the Royal Society of London Series B: Biological Sciences 266: 361369.Google Scholar
Hemelrijk, C. 1999b. Effects of cohesiveness on inter-sexual dominance relationships and spatial structure among group-living virtual entities. Pages 524534 in Advances in Artificial Life: Fifth European Conference on Artificial Life. Floreano, D., Nicoud, J.-D., and Mondana, F. (Eds.). Springer-Verlag, Berlin.Google Scholar
Hemelrijk, C. K. 2000. Towards the integration of social dominance and spatial structure. Animal Behaviour 59: 10351048.Google Scholar
Hemelrijk, C. 2001. Sexual attraction and inter-sexual dominance among virtual agents. Lecture Notes in Computer Science 1979: 167180.Google Scholar
Hemelrijk, C. 2002. Despotic societies, sexual attraction and the emergence of male “tolerance”: an agent-based model. Behaviour 139: 729747.Google Scholar
Hemelrijk, C. and Puga-Gonzalez, I. I. 2012. An individual-oriented model on the emergence of support in fights, its reciprocation and exchange. PLoS One 7: e37271.Google Scholar
Hemelrijk, C., Wantia, J., and Dätwyler, M. 2003. Female co-dominance in a virtual world: ecological, cognitive, social and sexual causes. Behaviour 140: 12471273.Google Scholar
Hill, R. A., Logan, B. S., Sellers, W. I., and Zappala, J. 2014. An agent-based model of group decision making in baboons. Pages 454476 in Modelling Natural Action Selection. Seth, A. K., Prescott, T. J., and Bryson, J. J. (Eds.). Cambridge University Press, Cambridge.Google Scholar
Hogeweg, P. 1988. MIRROR beyond MIRROR, puddles of Life. Pages 297315 in Artificial Life I. Langton, C. G. (Ed.). Addison-Wesley, Redwood City, CA.Google Scholar
Hogeweg, P. and Hesper, B. 1985. Socioinformatic processes, a MIRROR modelling methodology. Journal of Theoretical Biology 113: 311330.Google Scholar
Hogeweg, P. and Hesper, B. 1990. Individual oriented modelling in ecology. Mathematical and Computer Modelling 13: 8390.Google Scholar
Janssen, M. 2009. Understanding artificial anasazi. Journal of Artificial Societies and Social Simulation 12: 13.Google Scholar
Janssen, M. and de Vries, B. 1998. The battle of perspectives: a multi-agent model with adaptive responses to climate change. Ecological Economics 26: 4365.Google Scholar
Johnston, K. M. 2013. Agent Analyst: Agent-Based Modeling in ArcGIS. ESRI Press, Redlands, CA.Google Scholar
Klügl, F., Herrler, R., and Fehler, M. 2006. SeSAm: Implementation of agent-based simulation using visual programming. Pages 14391440 in Proceedings of the Fifth International Joint Conference on Autonomous Agents and Multiagent Systems. Association for Computing Machinery, New York.Google Scholar
Kokko, H. 2007. Modelling for Field Biologists and Other Interesting People. Cambridge University Press, Cambridge.Google Scholar
Kolher, T. A., Varien, M. D., Wright, A., and Kuckelman, K. A. 2009. Mesa Verde migrations: new archaeological research and computer simulation suggest why ancestral Puebloans deserted the northern Southwest United States. American Scientist 96: 145153.Google Scholar
Lansing, J. S. and Kremer, J. N. 1993. Emergent properties of Balinese water temple networks: coadaptation on a rugged fitness landscape. American Anthropologist 95: 97114.Google Scholar
Luke, S., Cioffi-Revilla, C., Panait, L., Sullivan, K., and Balan, G. 2005. MASON: a multi-agent simulation environment. Simulation 81: 517527.Google Scholar
Macal, C. M. and North, M. J. 2009. Agent-based modeling and simulation. Pages 8698 in Proceedings of the 2009 Winter Simulation Conference. Rossetti, M. D., Hill, R. R., Johansson, B., et al. (Eds.). IEE, Piscataway, NJ.Google Scholar
Macal, C. M. and North, M. J. 2010. Toward teaching agent-based simulation. Pages 1–10 in Proceedings of the 2010 Winter Simulation Conference, Johansson, B., Jain, S., Montoya-Torres, J., Hugan, J., and Yücesan, E. (Eds.).Google Scholar
Macy, M. W. and Willer, R. 2002. From factors to actors: computational sociology and agent-based modeling. Annual Review of Sociology 28: 143166.Google Scholar
Miller, J. H. and Page, S. E. 2007. Complex Adaptive Systems: An Introduction to Computational Models of Social Life. Princeton University Press, Princeton, NJ.Google Scholar
Nibbelink, N. P. and Carpenter, S. R. 1998. Interlake variation in growth and size structure of bluegill (Lepomis macrochirus): inverse analysis of an individual-based model. Canadian Journal of Fisheries and Aquatic Sciences 55: 387396.Google Scholar
Nikolai, C. and Madey, G. 2009. Tools of the trade: a survey of various agent based modeling platforms. Journal of Artificial Societies and Social Simulation 12: 2.Google Scholar
North, M. J., Collier, N. T., Ozik, J., et al. 2013. Complex Adaptive Systems Modeling With Repast Simphony. Complex Adaptive Systems Modeling. Springer, Heidelberg.Google Scholar
Nunn, C. L. 2009. Using agent-based models to investigate primate disease ecology. Pages 83110 in Primate Parasite Ecology. Huffman, M. A. and Chapman, C. A. (Eds.). Cambridge University Press, Cambridge.Google Scholar
Oom, S., Beecham, J., Legg, C., and Hester, A. 2004. Foraging in a complex environment: from foraging strategies to emergent spatial properties. Ecological Complexity 1: 299327.Google Scholar
Parker, D. C., Manson, S. M., Janssen, M. A., Hoffmann, M. J., and Deadman, P. 2003. Multi-agent systems for the simulation of land-use and land-cover change: a review. Annals of the Association of American Geographers 93: 314337.Google Scholar
Parrott, L. and Kok, R. 2002. A generic, individual-based approach to modelling higher trophic levels in simulation of terrestrial ecosystems. Ecological Modelling 151: 154178.Google Scholar
Pepper, J. W. and Smuts, B. B. 1999. The evolution of cooperation in an ecological context: an agent-based model. Pages 4576 in Dynamics in Human and Primate Societies. Kohler, T. and Gumerman, G. (Eds.). Oxford University Press, Oxford.Google Scholar
Perez, L. and Dragocevic, S. 2009. An agent-based approach for modeling dynamics of contagious disease spread. International Journal of Health Geographics 8: 50.Google Scholar
Peuquet, D. J. 2005. Time in GIS and geographical databases. Pages 91103 in Geographical Information Systems: Principles, Techniques, Management and Applications (Abridged Edition), Longley, P. A., Goodchild, M. F., Maguire, D. J., and Rhind, D. W. (Eds.). Wiley, Hoboken, NJ.Google Scholar
Puga-Gonzalez, I. I., Hildenbrandt, H. H., and Hemelrijk, C. K. C. 2009. Emergent patterns of social affiliation in primates, a model. PLoS Computational Biology 5: e1000630.Google Scholar
Railsback, S. F. and Grimm, V. 2012. Agent-Based and Individual-Based Modeling: A Practical Introduction. Princeton University Press, Princeton, NJ.Google Scholar
Railsback, S. F., Lytinen, S. L., and Jackson, S. K. 2006. Agent-based simulation platforms: Review and development recommendations. Simulation 82: 609623.Google Scholar
Ramos-Fernández, G., Boyer, D., and Gomez, V. P. 2006. A complex social structure with fission–fusion properties can emerge from a simple foraging model. Behavioural Ecology and Sociobiology 60: 536549.Google Scholar
Reas, C. and Fry, B. 2007. Processing: A Programming Handbook for Visual Designers and Artists. MIT Press, Cambridge, MA.Google Scholar
Reas, C. and Fry, B. 2010. Getting Started with Processing. O’Reilly Media, Cambridge, MA.Google Scholar
Roche, B., Guégan, J.-F., and Bousquet, F. 2008. Multi-agent systems in epidemiology: a first step for computational biology in the study of vector-borne disease transmission. BMC Bioinformatics 9: 435.Google Scholar
Rollins, N. D., Barton, C. M., Bergin, S., Janssen, M. A., and Lee, A. 2014. A computational model library for publishing model documentation and code. Environmental Modelling & Software 61: 5964.Google Scholar
Sellers, W., Hill, R., and Logan, B. 2007. An agent-based model of group decision making in baboons. Philosophical Transactions of the Royal Society Series B: Biological Sciences 362: 16991710.Google Scholar
Smaldino, P. E. and Schank, J. C. 2012. Human mate choice is a complex system. Complexity 17: 1122.Google Scholar
Sullivan, K., Coletti, M., and Luke, S. 2010. GeoMason: GeoSpatial support for MASON. Technical Report GMU-CS-TR-2010–16. Department of Computer Science, George Mason University.Google Scholar
te Boekhorst, I. J. A. and Hogeweg, P. 1994a. Self-structuring in artificial “chimps” offers new hypotheses for male grouping in chimpanzees. Behaviour 130: 229252.Google Scholar
te Boekhorst, I. and Hogeweg, P. 1994b. Effect of tree size on travel band formation in orang-utans: data analysis suggested by a model study. Pages 119129 in Artificial Life IV: Proceedings of the Fourth International Workshop on the Synthesis and Simulation of Living Systems. Brooks, R. A. and Maes, P. (Eds.). MIT Press, Cambridge, MA.Google Scholar
Tesfatsion, L. 2002. Agent-based computational economics: growing economies from the bottom up. Artificial Life 8: 5582.Google Scholar
Tesfatsion, L. 2003. Agent-based computational economics: modeling economies as complex adaptive systems. Information Sciences 149: 262268.Google Scholar
Tesfatsion, L. and Judd, K. L. (Eds.). 2006. Handbook of Computational Economics: Volume 2: Agent-Based Computational Economics. Elsevier, Amsterdam.Google Scholar
van der Post, D. J. and Hogeweg, P. 2008. Diet traditions and cumulative cultural processes as side-effects of grouping. Animal Behaviour 75: 133144.Google Scholar
van der Post, D. J. and Hogeweg, P. 2009. Cultural inheritance and diversification of diet in variable environments. Animal Behaviour 78: 155166.Google Scholar
Wang, J., Brown, D. G., Riolo, R. L., Page, S. E., and Agrawal, A. 2013. Exploratory analyses of local institutions for climate change adaptation in the Mongolian grasslands: an agent-based modeling approach. Global Environmental Change 23: 12661276.Google Scholar
Ward, J., Austin, R., and Macdonald, D. 2000. A simulation model of foraging behaviour and the effect of predation risk. Journal of Animal Ecology 69: 1630.Google Scholar
Ward, C., Gobet, F., and Kendall, G. 2001. Evolving collective behavior in an artificial ecology. Artificial Life 7: 191209.Google Scholar
Wilensky, U. 1999. NetLogo. Center for Connected Learning and Computer-Based Modeling, Northwestern University, Evanston, IL. Available at: http://ccl.northwestern.edu/netlogo/.Google Scholar

References

Freeman, N. J., Pasternak, G. M., Rubi, T. L., Barrett, L., and Henzi, S. P. 2012. Evidence for scent marking in vervet monkeys? Primates 53(3): 311315.Google Scholar
Gregory, T., Mullett, A., and Norconk, M. A. 2014. Strategies for navigating large areas: a GIS spatial ecology analysis of the bearded saki monkey, Chiropotes sagulatus, in Suriname. American Journal of Primatology 76(6): 586595.Google Scholar
Josephs, N., Bonnell, T., Dostie, M., Barrett, L., and Henzi, S.P. 2016. Working the crowd: sociable vervets benefit by reducing exposure to risk. Behavioral Ecology 27(4): 988994.Google Scholar
Ramos-Fernández, G., Pinacho-Guendulain, B., Miranda-Pérez, A., and Boyer, D. 2011. No evidence of coordination between different subgroups in the fission–fusion society of spider monkeys (Ateles geoffroyi). International Journal of Primatology 32: 13671382.Google Scholar
Russo, S. E. and Augspurger, C. K. 2004. Aggregated seed dispersal by spider monkeys limits recruitment to clumped patterns in Virola calophylla. Ecology Letters 7(11): 10581067.Google Scholar
Schmitt, C. A. and Di Fiore, A. 2015. Predation risk sensitivity and the spatial organization of primate groups: a case study using GIS in lowland woolly monkeys (Lagothrix lagotricha poeppigii). American Journal of Physical Anthropology 156(1): 158165.Google Scholar
Shaffer, C. A. 2013. GIS analysis of patch use and group cohesiveness of bearded sakis (Chiropotes sagulatus) in the upper Essequibo Conservation Concession, Guyana. American Journal of Physical Anthropology 150(2): 235246.Google Scholar
Sueur, C., Briard, L., and Petit, O. 2011a. Individual analyses of Lévy walk in semi-free ranging Tonkean macaques (Macaca tonkeana). PLoS One 6(10): e26788.Google Scholar
Sueur, C., Jacobs, A., Amblard, F., Petit, O., and King, A. J. 2011b. How can social network analysis improve the study of primate behavior? American Journal of Primatology 73(8): 703719.Google Scholar
Willems, E. P. and Hill, R. A. 2009. Predator‐specific landscapes of fear and resource distribution: effects on spatial range use. Ecology 90(2): 546555.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
×