Skip to main content Accessibility help
×
Hostname: page-component-848d4c4894-mwx4w Total loading time: 0 Render date: 2024-06-19T04:49:35.567Z Has data issue: false hasContentIssue false

Part VI - Innovation and Problem-Solving

Published online by Cambridge University Press:  01 July 2021

Allison B. Kaufman
Affiliation:
University of Connecticut
Josep Call
Affiliation:
University of St Andrews, Scotland
James C. Kaufman
Affiliation:
University of Connecticut
Get access

Summary

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

Aplin, L. M., Farine, D. R., Morand-Ferron, J., Cockburn, A., Thornton, A., & Sheldon, B. C. (2015). Experimentally induced innovations lead to persistent culture via conformity in wild birds. Nature, 518, 538541. https://doi.org/10.1038/nature13998Google Scholar
Benson-Amram, S., Dantzer, B., Stricker, G., Swanson, E. M., & Holekamp, K. E. (2016). Brain size predicts problem-solving ability in mammalian carnivores. Proc. Natl. Acad. Sci. USA, 113(9), 25322537. https://doi.org/10.1073/pnas.1505913113CrossRefGoogle ScholarPubMed
Boogert, N. J., Madden, J. R., Morand-Ferron, J., & Thornton, A. (2018). Measuring and understanding individual differences in cognition. Phil. Trans. Roy. Soc. Lon. B, 373(1756), 20170280. https://doi.org/10.1098/rstb.2017.0280Google Scholar
Ducatez, S., Clavel, J., & Lefebvre, L. (2015). Ecological generalism and behavioural innovation in birds: Technical intelligence or the simple incorporation of new foods? J. Anim. Ecol., 84(1), 7989. doi: 10.1111/1365-2656.12255.Google Scholar
Ducatez, S., Sol, D., Sayol, F., & Lefebvre, L. (2020). Behavioural plasticity is associated with reduced extinction risk in birds. Nat. Ecol. Evol, 4, 788–793. https://doi.org/10.1038/s41559-020-1168-8.Google Scholar
Dunbar, R. I. M. & Shultz, S. (2007). Evolution in the Social Brain. Science, 317(5843), 13441347. https://doi.org/10.1126/science.1145463Google Scholar
Ebel, S. J., Schmelz, M., Herrmann, E., & Call, J. 2019. Innovative problem solving in great apes: The role of visual feedback in the floating peanut task. Anim. Cogn., 22(5), 791805. doi: 10.1007/s10071-019-01275-0.Google Scholar
Fisher, J. & Hinde, R. A. (1949). The opening of milk bottles by birds. Brit. Birds, 42(347), 347357. https://doi.org/10.1038/1691006a0Google Scholar
Flynn, E., Turner, C., & Giraldeau, L.-A. (2016). Selectivity in social and asocial learning: Investigating the prevalence, effect and development of young children’s learning preferences. Phil. Trans. Roy. Soc. B, 371, 20150189. https://doi.org/10.1098/rstb.2015.0189Google Scholar
Griffin, A. S. & Guez, D. (2014). Innovation and problem solving: A review of common mechanisms. Behav. Proc., 109(Pt B), 121134.Google Scholar
Henrich, J. & McElreath, R. (2003). The evolution of cultural evolution. Evol. Anthropol., 12, 123135. https://doi.org/10.1002/evan.10110Google Scholar
Kummer, H. & Goodall, J. (1985). Conditions of innovative behavior in primates. Phil. Trans. Roy. Soc. Lon. Series, 308, 203214.Google Scholar
Kummer, H. & Goodall, J. (2003). Conditions of Innovative Behaviour in Primates. In Reader, S. M. & Laland, K. N. (Eds.), Anim. Innov., (223236). Oxford University Press. https://doi.org/10.1093/acprof:oso/9780198526223.003.0010Google Scholar
Lefebvre, L. (2011). Taxonomic counts of cognition in the wild. Biol. Lett., 7, 631633. https://doi.org/10.1098/rsbl.2010.0556Google Scholar
Lefebvre, L., Whitle, P., Lascaris, E. & Finkelstein, A. (1997). Feeding innovations and forebrain size in birds. Anim. Behav., 53(3), 549560. https://doi.org/10.1006/anbe.1996.0330Google Scholar
Lefebvre, L., Reader, S. M., & Sol, D. (2004). Brains, innovations and evolution in birds and primates. Brain Behav. Evol., 63, 233246. https://doi.org/10.1159/000076784CrossRefGoogle ScholarPubMed
Liker, A. & Bokony, V. (2009). Larger groups are more successful in innovative problem solving in house sparrows. Proc. Natl. Acad. Sci. USA, 106(19), 78937898. https://doi.org/10.1073/pnas.0900042106Google Scholar
McCabe, C. M., Reader, S. M., & Nunn, C. L. (2015). Infectious disease, behavioural flexibility, and the evolution of culture in primates. Proc. Roy. Soc. B, 282, 20140862. https://doi.org/10.1098/rspb.2014.0862Google Scholar
Mendes, N., Hanus, D., & Call, J. (2007). Raising the level: Orangutans use water as a tool. Biol Lett, 3(5), 453455. https://doi.org/10.1098/rsbl.2007.0198Google Scholar
Morand-Ferron, J. & Quinn, J. L. (2011). Larger groups of passerines are more efficient problem solvers in the wild. Proc. Natl. Acad. Sci. USA, 108, 1589815903. https://doi.org/10.1073/pnas.1111560108CrossRefGoogle ScholarPubMed
Morand-Ferron, J. & Quinn, J. L. (2015). The evolution of cognition in natural populations. Trends Cogn. Sci., 19(5), 235237. https://doi.org/10.1016/j.tics.2015.03.005Google Scholar
Morand-Ferron, J., Cole, E. F., & Quinn, J. L. (2016). Studying the evolutionary ecology of cognition in the wild: A review of practical and conceptual challenges. Biol. Rev., 91(2), 367389. https://doi.org/10.1111/brv.12174Google Scholar
Nelson, D. W. M., Crossland, M. R., & Shine, R. (2011). Behavioural responses of native predators to an invasive toxic prey species. Austral. Ecol., 36, 605611. https://doi.org/https://doi.org/10.1111/j.1442-9993.2010.02187.xGoogle Scholar
Quinn, J. L., Cole, E. F., Reed, T. E., & Morand-Ferron, J. (2016). Environmental and genetic determinants of innovativeness in a natural population of birds. Phil. Trans. Roy. Soc. B, 371(1690), 20150184. https://doi.org/10.1098/rstb.2015.0184Google Scholar
Ramsey, G., Bastian, M. L., & van Schaik, C. P. (2007). Animal innovation defined and operationalized. Behav. Brain Sci., 30, 393437. https://doi.org/10.1017/S0140525X07002373Google Scholar
Reader, S. (2003). Innovation and social learning: Individual variation and brain evolution. Anim. Biol., 53(2), 147158. https://doi.org/10.1163/157075603769700340Google Scholar
Reader, S. M. & Laland, K. N. (2003). Animal Innovation. Oxford: Oxford University Press.Google Scholar
Reader, S. M., Morand-Ferron, J., & Flynn, E. (2016). Animal and human innovation: Novel problems and novel solutions. Phil. Trans. Roy. Soc. B, 371(1690). https://doi.org/10.1098/rstb.2015.0182Google Scholar
Reader, S. M. & Laland, K. N. (2002). Social intelligence, innovation, and enhanced brain size in primates. Proc. Natl. Acad. Sci. USA, 99(7), 44364441. https://doi.org/10.1073/pnas.062041299Google Scholar
Roth, T. C., LaDage, L. D., & Pravosudov, V. V. (2010). Learning capabilities enhanced in harsh environments: A common garden approach. Proc. Roy. Soc. B, 277, 31873193. https://doi.org/10.1098/rspb.2010.0630Google Scholar
Sayol, F., Lefebvre, L., & Sol, D. (2016). Relative brain size and its relation with the associative pallium in birds. Brain, Behav. Evol., 87(2), 6977. https://doi.org/10.1159/000444670Google Scholar
Sol, D. (2003) Behavioural Flexibility: A Neglected Issue in the Ecological and Evolutionary Literature? In Reader, S. M. & Laland, K. N. (Eds.), Animal Innovation (pp. 6382). Oxford: Oxford University Press..CrossRefGoogle Scholar
Sol, D. (2009). The Cognitive-Buffer Hypothesis for the Evolution of Large Brains. In Dukas, R & Ratcliffe, J. M. (Eds.), Cognitive Ecology II (pp. 111134). Chicago: University of Chicago Press.Google Scholar
Sol, D., Duncan, R. P., Blackburn, T. M., Cassey, P., & Lefebvre, L. (2005). Big brains, enhanced cognition, and response of birds to novel environments. Proc. Nat. Acad. Sci. USA, 102(15), 54605465. https://doi.org/10.1073/pnas.0408145102Google Scholar
Sol, D., Griffin, A. S., Bartomeus, I., & Boyce, H. (2011). Exploring or avoiding novel food resources? The novelty conflict in an invasive bird. PLoS ONE, 6(5). https://doi.org/10.1371/journal.pone.0019535Google Scholar
Sol, D., Sayol, F., Ducatez, S., & Lefebvre, L. (2016). The life-history basis of behavioural innovations. Phil. Trans. Roy. Soc. B, 371(1690), 20150187. https://doi.org/10.1098/rstb.2015.0187Google Scholar
Street, S. E., Navarrete, A. F., Reader, S. M., & Laland, K. N. (2017). Coevolution of cultural intelligence, extended life history, sociality, and brain size in primates. Proc. Natl. Acad. Sci. USA, 114(30), 79087914. https://doi.org/10.1073/pnas.1620734114Google Scholar
Taylor, A. H., Elliffe, D. M., Hunt, G. R., Emery, N. J., Clayton, N. S., & Gray, R. D. (2011). New Caledonian crows learn the functional properties of novel tool types. PLoS ONE, 6(12), 26887. https://doi.org/10.1371/journal.pone.0026887Google Scholar
Teschke, I., Wascher, C. A. F., Scriba, M. F., von Bayern, A. M. P., Huml, V., Siemers, B., & Tebbich, S. (2013). Did tool-use evolve with enhanced physical cognitive abilities? Phil. Trans. Roy. Soc. B, 368, 20120418. https://doi.org/10.1098/rstb.2012.0418Google Scholar
Thorndike, E. L. (1898). Animal intelligence: An experimental study of the associative processes in animals. Psychol. Monogr. Gen. Appl., 2, 11251127.Google Scholar
Tomasello, M., Carpenter, M., Call, J., Behne, T., & Moll, H. (2005). Understanding and sharing intentions: The origins of cultural cognition. Behav. Brain Sci., 28(05), 675735. https://doi.org/10.1017/S0140525X05000129Google Scholar
Tomasello, M. & Herrmann, E. (2010). Ape and human cognition: What’s the difference? Curr. Direct. Psycholog. Sci., 19(1), 38. https://doi.org/10.1177/0963721409359300Google Scholar
von Bayern, A. M. P., Heathcote, R. J. P., Rutz, C. & Kacelnik, A. (2009). The role of experience in problem solving and innovative tool use in crows. Curr. Biol., https://doi.org/10.1016/j.cub.2009.10.037Google Scholar
Webster, S. J. & Lefebvre, L. (2001). Problem solving and neophobia in a columbiform-passeriform assemblage in Barbados. Anim. Behav.. https://doi.org/10.1006/anbe.2000.1725Google Scholar
Wyles, J. S., Kunkel, J. G., & Wilson, A. C. (1983). Birds, behavior, and anatomical evolution. Proc. Natl. Acad. Sci. USA, 80(14), 43944397. https://doi.org/10.1073/pnas.80.14.4394Google Scholar

References

Anastasi, A. (1961). Psychological Testing, 3rd ed. New York: Macmillan.Google Scholar
Anderson, B. (1993). Evidence from the rat for a general factor that underlies cognitive performance and that relates to brain size. Intelligence? Neuroscience Letters, 153, 98112.Google Scholar
Barkow, J. H., Cosmides, L., & Tooby, J. (1992). The Adapted Mind: Evolutionary Psychology and the Generation of Culture. New York: Oxford University Press.Google Scholar
Burkart, J. M., Schubiger, M. N., & van Schaik, C. P. (2016). The evolution of general intelligence. Behavioral and Brain Sciences, doi: 10.1017/S0140525X16000959 e195Google Scholar
Chomsky, N. (1988). Language and Problems of Knowledge: The Managua Lectures. Cambridge, MA: The MIT Press.Google Scholar
Chopik, W. J., Bremner, R. H., Defever, A. M., & Keller, V. N. (2018). How (and whether) to teach undergraduates about the replication crisis in psychological science. Teaching of Psychology, 45, 158163. doi: 10.1177/0098628318762900Google Scholar
Collias, N. E. & Collias, E. C. (2004). Comparison of vocal signals of three species of African finches. Behaviour, 141, 11511171. doi: 10.1163/1568539042664588Google Scholar
Cronbach, L. J. 1957). The two disciplines of scientific psychology. The American Psychologist, 12, 671684.Google Scholar
Emmerton, J. (1998). Numerosity differences and effects of stimulus density on pigeons’ discrimination performance. Animal Learning & Behavior, 26, 243256. doi: 10.3758/BF03199218Google Scholar
Gallistel, C. R. (2003). The Principle of Adaptive Specialization as it Applies to Learning and Memory. In Kluwe, R. H., Lüer, G., Rösler, F., Kluwe, R. H., Lüer, G., & Rösler, F. (Eds.), Principles of Learning and Memory (pp. 259280). Cambridge, MA: Birkhäuser. doi: 10.1007/978-3-0348-8030-5_15Google Scholar
Galsworthy, M. J., Paya-Cano, J. L., Monleon, S., & Plomin, R. (2002). Evidence for general cognitive ability (g) in heterogeneous stock mice and analysis of potential confounds. Genes, Brain and Behavior, 1, 8895.CrossRefGoogle ScholarPubMed
Galsworthy, M. J., Paya-Cano, J. L., Monleon, S., Gregoryan, G., Fernandes, C., Schalkwyk, L. C., & Plomin, R. (2005). Assessing reliability, heritability and general cognitive ability in a battery of cognitive tasks for laboratory mice. Behavior Genetics, 35, 675692.Google Scholar
Grossman, H. C., Hale, G., Light, K., Kolata, S., Townsend, D. A., Goldfarb, Y., & … Matzel, L. D. (2007). Pharmacological modulation of stress reactivity dissociates general learning ability from the propensity for exploration. Behavioral Neuroscience, 121, 949964. doi: 10.1037/0735-7044.121.5.949Google Scholar
Herrmann, E., Hernández-Lloreda, M. V., Call, J., Hare, B., & Tomasello, M. (2010). The structure of individual differences in the cognitive abilities of children and chimpanzees. Psychological Science, 21, 102110. http://dx.doi.org/10.1177/0956797609356511Google Scholar
Jacobs, I. & Gärdenfors, P. (2017). The false dichotomy of domain-specific versus domain-general cognitionBehavioral and Brain Sciences40, E207. doi:10.1017/S0140525X16001679Google Scholar
Kolata, S., Light, K., Townsend, D. A., Hale, G., Grossman, H. C., & Matzel, L. D. (2005). Variations in working memory capacity predict individual differences in general learning abilities among genetically diverse mice. Neurobiology of Learning and Memory, 84, 241246.CrossRefGoogle ScholarPubMed
Light, K., Kolata, S., Townsend, D. A., Grossman, H., Hale, G., & Matzel, L. D. (2008). Up-regulation of exploratory tendencies does not enhance general learning abilities in juvenile or young-adult outbred mice. Neurobiology of Learning and Memory, 90, 317329.Google Scholar
Lilienfeld, S. O. (2017). Psychology’s replication crisis and the grant culture: Righting the ship. Perspectives on Psychological Science, 12, 660664. doi: 10.1177/1745691616687745Google Scholar
Locurto, C. (1991). Sense and Nonsense about IQ: The Case for Uniqueness. New York: Praeger Press.Google Scholar
Locurto, C. (1997). On the Comparative Generality of g. In Tomic, W. & Kingma, J. (Eds.), Advances in Cognition and Education (pp. 79100). Greenwich: JAI Press.Google Scholar
Locurto, C. (2016). G and g: Two markers of a general cognitive ability, or none? In Burkart, Schubiger, & van Schaik. (2016). The evolution of general intelligence. Behavioral and Brain Sciences, 40, 38. doi: 10.1017/S0140525X16001709 e211Google Scholar
Locurto, C. & Scanlon, C. (1998). Individual differences and a spatial learning factor in two strains of mice. Journal of Comparative Psychology, 112, 344352.Google Scholar
Locurto, C., Fortin, E., & Sullivan, R. (2003). The structure of individual differences in heterogeneous strain mice across problem types and motivational conditions. Genes, Brain, & Behavior, 2, 116.Google Scholar
Locurto, C., Benoit, A., Crowley, C., & Miele, A. (2006). The structure of individual differences in batteries of rapid acquisition tasks in mice. Journal of Comparative Psychology, 120, 378388.Google Scholar
Locurto, C., Fox, M., & Mazzella, A. (2015). Implicit learning in cotton-top tamarins (Saguinus oedipus) and pigeons (Columba livia). Learning & Behavior, 43, 129142. doi: 10.3758/s13420–015-0167-0Google Scholar
Mackintosh, N. J. (1998). IQ and Human Intelligence. Oxford: Oxford University Press. ISBN 978-0-19-852367-3.Google Scholar
Mackintosh, N. J., Wilson, B., & Boakes, R. A. (1985). Differences in Mechanisms of Intelligence among Vertebrates. In Weiskrantz, L. (Ed.), Animal Intelligence (pp. 5366). Oxford: Clarendon Press.Google Scholar
Macphail, E. M. (1982). Brain and Intelligence in Vertebrates. Oxford: Clarendon Press.Google Scholar
Macphail, E. M. (1985). Comparative studies of animal intelligence. Is Spearman’s g really Hull’s D? Behavioral and Brain Sciences, 8, 234235.Google Scholar
Macphail, E. M. (1987). The comparative psychology of intelligence. Behavioral and Brain Sciences, 10, 645656.Google Scholar
Matzel, L. D., Han, Y. R., Grossman, H. Karnik, M. S., Patel, D., Scott, N., Specht, S. M., & Gandhi, C. C. (2003). Individual differences in the expression of a “general” learning ability in mice. Journal of Neuroscience, 23, 64236433.CrossRefGoogle ScholarPubMed
Matzel, L. D., Wass, C., & Kolata, S. (2011) Individual differences in animal intelligence: Learning, reasoning, selective attention and inter-species conservation of a cognitive trait. Neurobiology of Learning and Memory 116, 181192.Google Scholar
Matzel, L. D., Kolata, S. Light, K., & Sauce, B. (2016). The tendency for social submission predicts superior cognitive performance in previously isolated male mice. Behavioural Processes, 134, 1221.Google Scholar
Matzel, L. D. & Sauce, B. (2017). Individual differences: Case studies of rodent and primate intelligence. Journal of Experimental Psychology: Animal Learning and Cognition, 43, 32340. doi: 10.1037/xan0000152Google Scholar
Miller, L. T. & Vernon, P. A. (1992). The general factor in short-term memory, intelligence, and reaction time. Intelligence, 16, 530.Google Scholar
Royce, J. R. (1950). The factorial analysis of animal behavior. Psychological Bulletin, 47, 235259. doi: 10.1037/h0061384Google Scholar
Royce, J. R. (1966). Concepts Generated in Comparative and Physiological Psychological Observations. In Cattell, R. B. (Ed.), Handbook of Multivariate Experimental Psychology (pp. 642683). Chicago: Rand McNally.Google Scholar
Sauce, B., Wass, C., Smith, A., Kwan, S., & Matzel, L. D. (2014) The external-internal loop of interference: Two types of attention and their influence on the learning abilities of mice. Neurobiology of Learning and Memory, 116, 181192.Google Scholar
Seligman, M. E. & Hager, J. L. (1972). Biological Boundaries of Learning. East Norwalk, CT: Appleton-Century-Crofts.Google Scholar
Spearman, C. E. (1904). “General intelligence,” objectively determined and measured. American Journal of Psychology15, 201293. doi: 10.2307/1412107Google Scholar
Spearman, C. (1927). The Abilities of Man. Oxford, England: Macmillan.Google Scholar
Terrace, H. (2010). The comparative psychology of serially organized behavior. Comparative Cognition & Behavior Reviews, 5, 523558. doi: 10.3819/ccbr.2010.50002Google Scholar
Wahlsten, D. (1978). Behavioral Genetics and Animal Learning. In Anisman, H. & Bignami, G. (Eds.), Psychopharmacology of Aversively Motivated Behavior (pp. 63118). New York: Plenum.Google Scholar
Warren, J. M. (1977). A Phylogenetic Approach to Learning and Intelligence. In Oliverio, A. (Ed.), Genetics, Environment and Intelligence (pp. 3756). New York: North-Holland.Google Scholar

References

Arden, R., Gottfredson, L. S., & Miller, G. (2009). Does a fitness factor contribute to the association between intelligence and health outcomes? Evidence from medical abnormality counts among 3654 US Veterans. Intelligence, 37(6), 581591. https://doi.org/10.1016/j.intell.2009.03.008Google Scholar
Audet, J.-N. & Lefebvre, L. (2017). What’s flexible in behavioral flexibility? Behavioral Ecology, 28(4), 943947. https://doi.org/10.1093/beheco/arx007Google Scholar
Baldwin, J. M. (1896). A new factor in evolutionThe American Naturalist30, 441451, 536–553. https://doi.org/10.1086/276408.Google Scholar
Bensky, M. K. & Bell, A. M. (2020). Predictors of individual variation in reversal learning performance in three-spined sticklebacks. Animal Cognition, 23925938. https://doi.org/10.1007/s10071-020-01399-8Google Scholar
Boogert, N. J., Anderson, R. C., Peters, S., & Searcy, W. A. (2011). Song repertoire size in male song sparrows correlates with detour reaching, but not with other cognitive measures. Animal Behaviour, 81(6), 12091216. https://doi.org/10.1016/j.anbehav.2011.03.004Google Scholar
Borgia, G. (1985a). Bower quality, number of decorations and mating success of male satin bowerbirds (Ptilonorhynchus violaceus): An experimental analysis. Animal Behaviour, 33(1), 266271. https://doi.org/10.1016/S0003-3472(85)80140-8Google Scholar
Borgia, G. (1985b). Bower destruction and sexual competition in the satin bowerbird (Ptilonorhynchus violaceus). Behavioral Ecology and Sociobiology, 18(2), 91100. https://doi.org/10.1007/BF00299037CrossRefGoogle Scholar
Borgia, G. (1993). The cost of display in the non-resource-based mating system of the satin bowerbird. The American Naturalist, 141(5), 729743. https://doi.org/10.1086/285502Google Scholar
Borgia, G. (1995). Why do bowerbirds build bowers? American Scientist, 83(6), 542547.Google Scholar
Borgia, G. & Gore, M. A. (1986). Feather stealing in the satin bowerbird (Ptilonorhynchus violaceus): Male competition and the quality of display. Animal Behaviour, 34(3), 727738. https://doi.org/10.1016/S0003-3472(86)80056-2Google Scholar
Borgia, G., Kaatz, I. M., & Condit, R. (1987). Flower choice and bower decoration in the satin bowerbird Ptilonorhynchus violaceus: A test of hypotheses for the evolution of male display. Animal Behaviour, 35(4), 11291139. https://doi.org/10.1016/S0003-3472(87)80169-0CrossRefGoogle Scholar
Borgia, G. & MuellerU. (1992). Bower destruction, decoration stealing and female choice in the spotted bowerbird Chlamydera maculataEmu, 92(1), 1118https://doi.org/10.1071/MU9920011Google Scholar
Borgia, G. & Presgraves, D. C. (1998). Coevolution of elaborated male display traits in the spotted bowerbird: An experimental test of the threat reduction hypothesis. Animal Behaviour, 56(5), 11211128. https://doi.org/10.1006/anbe.1998.0908Google Scholar
Borgia, G. & Keagy, J. (2006). An inverse relationship between decoration and food colour preferences in satin bowerbirds does not support the sensory drive hypothesis. Animal Behaviour, 72(5), 11251133. https://doi.org/10.1016/j.anbehav.2006.03.015Google Scholar
Borgia, G. & Keagy, J. (2015). Sexual Selection and Cognitive Ability: What Bowerbirds Can Teach Us. In Irschick, D., Briffa, M., & Podos, J. (Eds.), Animal Signaling and Function: An Integrative Approach. Hoboken, NJ: John Wiley and Sons.Google Scholar
Boul, K. E., Funk, W. C., Darst, C. R., Cannatella, D. C., & Ryan, M. J. (2007). Sexual selection drives speciation in an Amazonian frog. Proceedings of the Royal Society of London B: Biological Sciences, 274(1608), 399406. https://doi.org/10.1098/rspb.2006.3736Google Scholar
Bravery, B. D. & Goldizen, A. W. (2007). Male satin bowerbirds (Ptilonorhynchus violaceus) compensate for sexual signal loss by enhancing multiple display features. Naturwissenschaften, 94(6), 473476. https://doi.org/10.1007/s00114-006-0211-1Google Scholar
Cole, E. F., Morand-Ferron, J., Hinks, A. E., & Quinn, J. L. (2012). Cognitive ability influences reproductive life history variation in the wild. Current Biology, 22(19), 18081812. https://doi.org/10.1016/j.cub.2012.07.051Google Scholar
Coleman, S. W. (2005). Variable Female Preferences and the Evolution of Complex Male Displays in the Satin Bowerbird (Ptilonorhynchus violaceus). [Doctoral dissertation, University of Maryland, College Park]. Digital Repository at the University of Maryland.Google Scholar
Coleman, S. W., Patricelli, G. L., & Borgia, G. (2004). Variable female preferences drive complex male displays. Nature, 428, 742745. https://doi.org/10.1038/nature02419Google Scholar
Collis, K. & Borgia, G. (1993). The costs of male display and delayed plumage maturation in the satin bowerbird (Ptilonorhynchus violaceus). Ethology, 94(1), 5971. https://doi.org/10.1111/j.1439-0310.1993.tb00547.xGoogle Scholar
Croston, R., Kozlovsky, D. Y., Branch, C. L., Parchman, T. L., Bridge, E. S., & Pravosudov, V. V. (2016). Individual variation in spatial memory performance in wild mountain chickadees from different elevations. Animal Behaviour, 111(1), 225234. https://doi.org/10.1016/j.anbehav.2015.10.015Google Scholar
Darwin, C. (1871). The Descent of Man and Selection in Relation to Sex. London: John Murray.Google Scholar
Day, L. B., Westcott, D. A., & Olster, D. H. (2005). Evolution of bower complexity and cerebellum size in bowerbirds. Brain, Behavior, and Evolution, 66(1), 6272. https://doi.org/10.1159/000085048Google Scholar
Diamond, J. (1986). Animal art: Variation in bower decorating style among male bowerbirds Ambylornis inornatus. Proceedings of the National Academy of Sciences of the USA, 83(9), 30423046. https://doi.org/10.1073/pnas.83.9.3042Google Scholar
Diamond, J. (1987). Bower building and decoration by the bowerbird Amblyornis inornatus. Ethology, 74(3), 177204. https://doi.org/10.1111/j.1439-0310.1987.tb00932.xGoogle Scholar
Doerr, N. R. (2010). Decoration supplementation and male-male competition in the great bowerbird (Ptilonorhynchus nuchalis): A test of the social control hypothesisBehavioral Ecology and Sociobiology, 64(11), 18871896. https://doi.org/10.1007/s00265-010-1000-6Google Scholar
Doerr, N. R. & Endler, J. A. (2015). Illusions vary because of the types of decorations at bowers, not male skill at arranging them, in great bowerbirds. Animal Behaviour, 99(1), 7382. https://doi.org/10.1016/j.anbehav.2014.10.022Google Scholar
Endler, J. A., Endler, L. C., & Doerr, N. R. (2010). Great bowerbirds create theaters with forced perspective when seen by their audience. Current Biology. 20(18), 16791684. https://doi.org/10.1016/j.cub.2010.08.033Google Scholar
Erwin, D. H. (2015). Novelty and innovation in the history of life. Current Biology, 25(19), R930R940. https://doi.org/10.1016/j.cub.2015.08.019Google Scholar
Federspiel, I. G., Garland, A., Guez, D., Bugnyar, T., Healy, S. D., Güntürkün, O., & Griffin, A. (2017). Adjusting foraging strategies: A comparison of rural and urban common mynas (Acridotheres tristis). Animal Cognition, 20(1), 6574. https://doi.org/10.1007/s10071-016-1045-7Google Scholar
Griffin, A. S., Guez, D., Lermite, F., & Patience, M. (2013). Tracking changing environments: Innovators are fast, but not flexible learners. PLoS ONE, 8(12), e84907. https://doi.org/10.1371/journal.pone.0084907Google Scholar
Griffin, A. S. & Guez, D. (2014). Innovation and problem solving: A review of common mechanisms. Behavioral Processes, 109(Pt B), 121134. https://doi.org/10.1016/j.beproc.2014.08.027Google Scholar
Hicks, R. E., Larned, A., & Borgia, G. (2013). Bower paint removal leads to reduced female visits, suggesting bower paint functions as a chemical signal. Animal Behaviour, 85(6), 12091215. https://doi.org/10.1016/j.anbehav.2013.03.007CrossRefGoogle Scholar
Hoekstra, H. E., Hoekstra, J. M., Berrigan, D., Vignieri, S. N., Hoang, A., Hill, C. E., Beerli, P., & Kingsolver, J. P. (2001). Strength and tempo of directional selection in the wild. Proceedings of the National Academy of Sciences of the U.S.A., 98(16), 91579160. https://doi.org/10.1073/pnas.161281098Google Scholar
Isden, J., Panayi, C., Dingle, C., & Madden, J. (2013). Performance in cognitive and problem-solving tasks in male spotted bowerbirds does not correlate with mating success. Animal Behaviour, 86(4), 829838. https://doi.org/10.1016/j.anbehav.2013.07.024Google Scholar
Keagy, J., Savard, J.-F., & Borgia, G. (2009). Male satin bowerbird problem-solving ability predicts mating success. Animal Behaviour, 78(4), 809817. https://doi.org/10.1016/j.anbehav.2009.07.011Google Scholar
Keagy, J., Savard, J.-F., & Borgia, G. (2011). Complex relationship between multiple measures of cognitive ability and male mating success in satin bowerbirds, Ptilonorhynchus violaceus. Animal Behaviour, 81(5), 10631070. https://doi.org/10.1016/j.anbehav.2011.02.018Google Scholar
Keagy, J., Lettieri, L., & Boughman, J. W. (2016). Male competition fitness landscapes predict both forward and reverse speciation. Ecology Letters, 19(1), 7180. https://doi.org/10.1111/ele.12544Google Scholar
Kingsolver, J. G., Hoekstra, H. E., Hoekstra, J. M., Berrigan, D., Vignieri, S. N., Hill, C. E., Hoang, A., Gibert, P., & Beerli, P. (2001). The strength of phenotypic selection in natural populations. The American Naturalist, 157(3), 245261. https://doi.org/10.1086/319193Google Scholar
Kingsolver, J. G., Hoekstra, H. E., Hoekstra, J. M., Berrigan, D., Vignieri, S. N., Hill, C. E., Hoang, A., Gibert, P., & Beerli, P. (2008). Data from: The strength of phenotypic selection in natural populations. Dryad Digital Repository. [Data set]. https://doi.org/10.5061/dryad.166Google Scholar
Kraaijeveld, K., Kraaijeveld-Smit, F. J. L., & Maan, M. E. (2011). Sexual selection and speciation: The comparative evidence revisited. Biological Reviews, 86(3), 367377. https://doi.org/10.1086/319193.Google Scholar
Kusmierski, R., Borgia, B., Uy, A. & Crozier, R. (1997). Labile evolution of display traits in bowerbirds indicates reduced effects of phylogenetic constraint. Proceedings of the Royal Society of London B: Biological Sciences, 264(1380), 307313. https://doi.org/10.1098/rspb.1997.0044Google Scholar
Lande, R. & Arnold, S. J. (1983). The measurement of selection on correlated characters. Evolution, 37(6), 12101226. https://doi.org/10.1111/j.1558-5646.1983.tb00236.x.Google Scholar
Larned, A. F. (2012). The effect of sunlight on decoration placement and mating success in male satin bowerbirds. [Master’s thesis, University of Maryland, College Park]. Digital Repository at the University of Maryland.Google Scholar
Lefebvre, L., Reader, S. M., & Sol, D. (2004). Brains, innovations and evolution in birds and primates. Brain, Behavior, and Evolution, 63(4), 233246. https://doi.org/10.1159/000076784Google Scholar
Loffredo, C. A. & Borgia, G. (1986). Male courtship vocalizations as cues for mate choice in the satin bowerbird (Ptilonorhynchus violaceus). Auk, 103(1) 189195.Google Scholar
Madden, J. (2001). Sex, bowers and brains. Proceedings of the Royal Society of London B: Biological Sciences, 268(1469), 833838. https://doi.org/10.1098/rspb.2000.1425Google Scholar
Madden, J. R. (2002). Bower decorations attract females but provoke other male spotted bowerbirds: Bower owners resolve this trade-off. Proceedings of the Royal Society of London B: Biological Sciences, 269(1498), 13471351. https://doi.org/10.1098/rspb.2002.1988Google Scholar
Madden, J. R. (2003). Bower decorations are good predictors of mating success in the spotted bowerbird. Behavioral Ecology and Sociobiology, 53(5), 269277. https://doi.org/10.1007/s00265-003-0583-6Google Scholar
Madden, J. R. (2006). Interpopulation differences exhibited by spotted bowerbirds Chlamydera maculata across a suite of male traits and female preferences. Ibis, 148(3), 425435. https://doi.org/10.1111/j.1474-919X.2006.00540.xGoogle Scholar
Madden, J. R. (2008). Do bowerbirds exhibit culture? Animal Cognition, 11(1), 1-12. https://doi.org/10.1007/s10071-007-0092-5Google Scholar
Madden, J. R. & Balmford, A. (2004). Spotted bowerbirds Chlamydera maculata do not prefer rare or costly bower decorations. Behavioral Ecology and Sociobiology, 55(6) 589595. https://doi.org/10.1007/s00265-003-0737-6Google Scholar
Marshall, A. J. (1954). Bower-birds. Biological Reviews, 29, 1-45https://doi.org/10.1111/j.1469-185X.1954.tb01395.xGoogle Scholar
Marshall, A. J. (1956). Bowerbirds. Scientific American, 194(6), 4853.Google Scholar
Martinez, J., Keagy, J., Wurst, B., Fetzner, W., & Boughman, J. W. (2016). The relative role of genes and environment on spatial learning ability in recently diverged stickleback fish. Evolutionary Ecology Research, 17(4565581.Google Scholar
Mendelson, T. C. (2003). Sexual isolation evolves faster than hybrid inviability in a diverse and sexually dimorphic genus of fish (Percidae: Etheostoma). Evolution, 57(2), 317327. https://doi.org/10.1111/j.0014-3820.2003.tb00266.xGoogle Scholar
M’Gonigle, L. K., Mazzucco, R., Otto, S. P., & Dieckmann, U. (2012). Sexual selection enables long-term coexistence despite ecological equivalence. Nature, 484(7395), 506509. https://doi.org/10.1038/nature10971Google Scholar
Morand-Ferron, J., Hamblin, S., Cole, E. F., Aplin, L. M., & Quinn, J. L. (2015). Taking the operant paradigm into the field: Associative learning in wild Great Tits. PLoS ONE, 10(8), e0133821. https://doi.org/10.1371/journal.pone.0133821Google Scholar
Morand-Ferron, J., Cole, E. F., & Quinn, J. L. (2016). Studying the evolutionary ecology of cognition in the wild: A review of practical and conceptual challenges. Biological Reviews of the Cambridge Philosophical Society, 91(2), 367389. https://doi.org/10.1111/brv.12174.Google Scholar
Morrison-Scott, T. C. S. (1937). Experiments on colour-vision in the satin bower-bird (Ptilonorhynchus violaceus), with other observations. Proceedings of the Zoological Society of London, Series A, 107(1), 4149. https://doi.org/10.1111/j.1469-7998.1937.tb08498.xGoogle Scholar
Nicolakakis, N., Sol, D., & Lefebvre, L. (2003). Behavioural flexibility predicts species richness in birds, but not extinction risk. Animal Behaviour, 65(3) 445452. https://doi.org/10.1006/anbe.2003.2085Google Scholar
Odling-Smee, L. C., Boughman, J. W., & Braithwaite, V. A. (2008). Sympatric species of threespine stickleback differ in their performance in a spatial learning task. Behavioral Ecology and Sociobiology, 62(12), 19351945. https://doi.org/10.1007/s00265-008-0625-1Google Scholar
Overington, S. E., Cauchard, L., Côté, K.-A., & Lefebvre, L. (2011). Innovative foraging behaviour in birds: What characterizes an innovator? Behavioural Processes, 87(3), 274285. https://doi.org/10.1016/j.beproc.2011.06.002.Google Scholar
Patricelli, G. L., Coleman, S. W., & Borgia, G. (2002). Male displays adjusted to female’s response. Nature, 415(6869), 279280https://doi.org/10.1038/415279aGoogle Scholar
Prokosch, M. D., Yeo, R. A., & Miller, G. F. (2005). Intelligence tests with higher g-loadings show higher correlations with body symmetry: Evidence for a general fitness factor mediated by developmental stability. Intelligence. 33(2), 203-213. https://doi.org/10.1016/j.intell.2004.07.007Google Scholar
Reynolds, S. M., Dryer, K., Bollback, J., Uy, J. A. C., Patricelli, G. L., Robson, T., Borgia, G., & Braun, M. (2007). Behavioral paternity predicts genetic paternity in Satin Bowerbirds (Ptilonorhynchus violaceus), a species with a non-resource-based mating system. Auk, 124(3), 857867. https://doi.org/10.1093/auk/124.3.857Google Scholar
Robson, T. E., Goldizen, A. W. & Green, D. J. (2005). The multiple signals assessed by female satin bowerbirds: Could they be used to narrow down females’ choices of mates? Biology Letters, 1(3), 264e267. https://doi.org/10.1098/rsbl.2005.0325Google Scholar
Roth, G. & Dicke, U. (2005). Evolution of the brain and intelligence. Trends in Cognitive Sciences, 9(5), 250257, https://doi.org/10.1016/j.tics.2005.03.005.Google Scholar
Rowe, C. & Healy, S. D. (2014). Measuring variation in cognition. Behavioral Ecology, 25(6) 12871292. https://doi.org/10.1093/beheco/aru090Google Scholar
Seehausen, O., Terai, Y., Magalhaes, I. S., Carleton, K. L., Mrosso, H. D. J., Miyagi, R., van der Sluijs, I., Schneider, M. V., Mann, M. E., Tachida, H., Imai, H., & Okada, N. (2008). Speciation through sensory drive in cichlid fish. Nature, 455(7213) 620626. https://doi.org/10.1038/nature07285Google Scholar
Sol, D. & Lefebvre, L. (2000). Behavioural flexibility predicts invasion success in birds introduced to New Zealand. Oikos, 90(3), 599605. https://doi.org/10.1034/j.1600-0706.2000.900317.xGoogle Scholar
Sol, D., Timmermans, S., & Lefebvre, L. (2002). Behavioural flexibility and invasion success in birds. Animal Behaviour, 63(3), 495502. https://doi.org/10.1006/anbe.2001.1953Google Scholar
Sol, D., Duncan, R. P., Blackburn, T. M., Cassey, P., & Lefebvre, L. (2005). Big brains, enhanced cognition, and response of birds to novel environments. Proceedings of the National Academy of Sciences of the USA, 102(15), 54605465. https://doi.org/10.1073/pnas.0408145102Google Scholar
Tebbich, S., Sterelny, K., & Teschke, I. (2010). The tale of the finch: Adaptive radiation and behavioural flexibility. Philosophical Transactions of the Royal Society of London B: Biological Sciences, 365(1543), 10991109. https://doi.org/10.1098/rstb.2009.0291Google Scholar
Thornton, A. & Samson, J. (2012). Innovative problem solving in wild meerkats. Animal Behaviour, 83(6), 14591468. https://doi.org/10.1016/j.anbehav.2012.03.018Google Scholar
Thornton, A., Isden, J., & Madden, J. R. (2014). Toward wild psychometrics: linking individual cognitive differences to fitness. Behavioral Ecology, 25(6), 12991301. https://doi.org/10.1093/beheco/aru095Google Scholar
Uy, J. A. C. & Borgia, G. (2000). Sexual selection drives rapid divergence in bowerbird display traits. Evolution, 54(1), 273278. https://doi.org/10.1111/j.0014-3820.2000.tb00027.xGoogle Scholar
Uy, J. A. C., Patricelli, G. L., & Borgia, G. (2000). Dynamic mate-searching tactic allows female satin bowerbirds Ptilonorhynchus violaceus to reduce searching. Proceedings of the Royal Society of London B: Biological Sciences. 267(1440), 251256. https://doi.org/10.1098/rspb.2000.0994Google Scholar
Uy, J. A. C., Patricelli, G. L., & Borgia, G. (2001a). Loss of preferred mates forces female satin bowerbirds (Ptilonorhynchus violaceus) to increase mate searching. Proceedings of the Royal Society of London B: Biological Sciences, 268(1467), 633638. https://doi.org/10.1098/rspb.2000.1413Google Scholar
Uy, J. A. C., Patricelli, G. L., & Borgia, G. (2001b). Complex mate searching in the satin bowerbird Ptilonorhynchus violaceus. The American Naturalist, 158(5), 530542. https://doi.org/10.1086/323118.Google Scholar
Wojcieszek, J. M., Nicholls, J. A., &Goldizen, A. W. (2007). Stealing behavior and the maintenance of a visual display in the satin bowerbird. Behavioral Ecology, 18(4), 689695. https://doi.org/10.1093/beheco/arm031Google Scholar

References

Auersperg, A. M. (2015). Exploration Technique and Technical Innovations in Corvids and Parrots. In Kaufman, A. B & Kaufman, J. C. (Eds.), Animal Creativity and Innovation (pp. 4572). London: Elsevier. https://doi.org/10.1016/B978-0-12-800648-1.00003-6.Google Scholar
Auersperg, A. M. I., Gajdon, G. K., & Huber, L. (2010). Kea, Nestor notabilis, produce dynamic relationships between objects in a second-order tool use task. Animal Behaviour, 80(5), 783789. https://doi.org/10.1016/j.anbehav.2010.08.007Google Scholar
Auersperg, A. M. I., von Bayern, A. M. P., Gajdon, G. K., Huber, L., & Kacelnik, A. (2011a). Flexibility in problem solving and tool use of kea and new Caledonian crows in a multi access box paradigm. PLoS One, 6(6), e20231. https://doi.org/10.1371/journal.pone.0020231Google Scholar
Auersperg, A. M. I., Huber, L., & Gajdon, G. K. (2011b). Navigating a tool end in a specific direction: Stick-tool use in kea (Nestor notabilis). Biology Letters, 7(6), 825828. https://doi.org/10.1098/rsbl.2011.0388Google Scholar
Auersperg, A. M. I., Szabo, B., von Bayern, A. M. P., & Kacelnik, A. (2012). Spontaneous innovation in tool manufacture and use in a Goffin’s cockatoo. Current Biology, 22(21), R903R904. https://doi.org/10.1016/j.cub.2012.09.002Google Scholar
Auersperg, A. M. I., Laumer, I. B., & Bugnyar, T. (2013a). Goffin cockatoos wait for qualitative and quantitative gains but prefer “better” to “more.” Biology Letters, 9(3), 2012109220121092. https://doi.org/10.1098/rsbl.2012.1092Google Scholar
Auersperg, A. M. I., Kacelnik, A., & von Bayern, A. M. P. (2013b). Explorative learning and functional inferences on a five-step means-means-end problem in Goffin’s cockatoos (Cacatua goffini). PLoS One, 8(7), e68979. https://doi.org/10.1371/journal.pone.0068979Google Scholar
Auersperg, A. M. I., Oswald, N., Domanegg, M., Gajdon, D., & Bugnyar, T. (2014a). Unrewarded object combinations in captive parrots. Animal Behavior and Cognition, 1(4), 470488. https://doi.org/10.12966/abc.11.05.2014Google Scholar
Auersperg, A. M. I., van Horik, J. O., Bugnyar, T., Kacelnik, A., Emery, N. J., & von Bayern, A. M. P. (2014b). Combinatory actions during object play in psittaciformes (Diopsittaca nobilis, Pionites melanocephala, Cacatua goffini) and corvids (Corvus corax, C. monedula, C. moneduloides). Journal of Comparative Psychology, 129(1), 6271. https://doi.org/10.1037/a0038314Google Scholar
Auersperg, A. M. I., von Bayern, A. M. P., Weber, S., Szabadvari, A., Bugnyar, T., & Kacelnik, A. (2014c). Social transmission of tool use and tool manufacture in Goffin cockatoos (Cacatua goffini). Proceedings of the Royal Society B: Biological Sciences, 281(1793), 2014097220140972. https://doi.org/10.1098/rspb.2014.0972Google Scholar
Auersperg, A. M. I., Borasinski, S., Laumer, I., & Kacelnik, A. (2016). Goffin’s cockatoos make the same tool type from different materials. Biology Letters, 12(11), 20160689. https://doi.org/10.1098/rsbl.2016.0689Google Scholar
Auersperg, A. M. I., Köck, C., Pledermann, A., O’Hara, M., & Huber, L. (2017). Safekeeping of tools in Goffin’s cockatoos, Cacatua goffiniana. Animal Behaviour, 128, 125133. https://doi.org/10.1016/j.anbehav.2017.04.010Google Scholar
Auersperg, A. M. I. & von Bayern, A. M. P. (2019). Who’s a clever bird – Now? A brief history of parrot cognition. Behaviour, 1(aop), 117. https://doi.org/10.1163/1568539X-00003550Google Scholar
Bateson, P. & Martin, P. (2013). Play, Playfulness, Creativity and Innovation. Cambridge: Cambridge University Press.Google Scholar
Birch, H. G. (1945). The relation of previous experience to insightful problem-solving. Journal of Comparative Psychology, 38(6), 367.Google Scholar
Biro, D., Haslam, M., & Rutz, C. (2013). Tool use as adaptation. Philosophical Transactions of the Royal Society B: Biological Sciences, 368(1630), 2012040820120408. https://doi.org/10.1098/rstb.2012.0408Google Scholar
Bond, A. B., Kamil, A. C., & Balda, R. P. (2007). Serial reversal learning and the evolution of behavioral flexibility in three species of North American corvids (Gymnorhinus cyanocephalus, Nucifraga columbiana, Aphelocoma californica). Journal of Comparative Psychology, 121(4), 372379. https://doi.org/10.1037/0735-7036.121.4.372CrossRefGoogle ScholarPubMed
Borsari, A. & Ottoni, E. B. (2005). Preliminary observations of tool use in captive hyacinth macaws (Anodorhynchus hyacinthinus). Animal Cognition, 8(1), 4852. https://doi.org/10.1007/s10071-004-0221-3Google Scholar
Boswall, J. (1977). Tool-using by birds and related behaviour. Aviculture Magazine, 83, 8897.Google Scholar
Demery, Z. P., Chappell, J., & Martin, G. R. (2011). Vision, touch and object manipulation in Senegal parrots Poicephalus senegalus. Proceedings of the Royal Society of London B: Biological Sciences. https://doi.org/10.1098/rspb.2011.0374Google Scholar
Diamond, J., & Bond, A. B. (1999). Kea, Bird of Paradox: The Evolution and Behavior of a New Zealand Parrot. Berkeley: University of California Press.Google Scholar
Ducatez, S., Clavel, J., & Lefebvre, L. (2015). Ecological generalism and behavioural innovation in birds: Technical intelligence or the simple incorporation of new foods? Journal of Animal Ecology, 84(1), 7989. https://doi.org/10.1111/1365-2656.12255Google Scholar
Emery, N. J. (2006). Cognitive ornithology: The evolution of avian intelligence. Philosophical Transactions of the Royal Society B: Biological Sciences, 361(1465), 2343. https://doi.org/10.1098/rstb.2005.1736Google Scholar
Gajdon, G. K., Fijn, N., & Huber, L. (2004). Testing social learning in a wild mountain parrot, the kea (Nestor notabilis). Animal Learning & Behavior, 32(1), 6271. https://doi.org/10.3758/BF03196007Google Scholar
Gajdon, G. K., Amann, L., & Huber, L. (2011). Keas rely on social information in a tool use task but abandon it in favour of overt exploration. Interaction Studies, 12(2), 304323. https://doi.org/10.1075/is.12.2.06gajGoogle Scholar
Gajdon, G. K., Lichtnegger, M., & Huber, L. (2014). What a parrot’s mind adds to play: The urge to produce novelty fosters tool use acquisition in kea. Open Journal of Animal Sciences, 4(2), 5158. https://doi.org/10.4236/ojas.2014.42008Google Scholar
Goodman, M., Hayward, T., & Hunt, G. R. (2018). Habitual tool use innovated by free-living New Zealand kea. Scientific Reports, 8(1). https://doi.org/10.1038/s41598-018-32363-9Google Scholar
Gossette, R. L. (1968). Examination of retention decrement explanation of comparative successive discrimination reversal learning by birds and mammals. Perceptual and Motor Skills, 27(3_suppl), 11471152.Google Scholar
Greenberg, R. & Mettke-Hofmann, C. (2001). Ecological Aspects of Neophobia and Neophilia in Birds. In Nolan, V Jr. & Thompson, C. F. (Eds.), Current Ornithology (pp. 119178). Boston: Springer.Google Scholar
Griffin, A. S., Diquelou, M., & Perea, M. (2014). Innovative problem solving in birds: A key role of motor diversity. Animal Behaviour, 92, 221227. https://doi.org/10.1016/j.anbehav.2014.04.009Google Scholar
Griffin, A. S., & Guez, D. (2014). Innovation and problem solving: A review of common mechanisms. Behavioural Processes, 109, 121134. https://doi.org/10.1016/j.beproc.2014.08.027Google Scholar
Güntürkün, O. & Bugnyar, T. (2016). Cognition without cortex. Trends in Cognitive Sciences, 20(4), 291303. https://doi.org/10.1016/j.tics.2016.02.001Google Scholar
Gutiérrez-Ibáñez, C., Iwaniuk, A. N., & Wylie, D. R. (2018). Parrots have evolved a primate-like telencephalic-midbrain-cerebellar circuit. Scientific Reports, 8(1). https://doi.org/10.1038/s41598-018-28301-4Google Scholar
Hansell, M. & Ruxton, G. (2008). Setting tool use within the context of animal construction behaviour. Trends in Ecology & Evolution, 23(2), 7378. https://doi.org/10.1016/j.tree.2007.10.006Google Scholar
Haslam, M. (2013). ‘Captivity bias’ in animal tool use and its implications for the evolution of hominin technology. Philosophical Transactions of the Royal Society B: Biological Sciences, 368(1630), 2012042120120421. https://doi.org/10.1098/rstb.2012.0421Google Scholar
Heinsohn, R., Zdenek, C. N., Cunningham, R. B., Endler, J. A., & Langmore, N. E. (2017). Tool-assisted rhythmic drumming in palm cockatoos shares key elements of human instrumental music. Science Advances, 3(6), e1602399. https://doi.org/10.1126/sciadv.1602399Google Scholar
Herculano-Houzel, S. (2017). Numbers of neurons as biological correlates of cognitive capability. Current Opinion in Behavioral Sciences, 16, 17. https://doi.org/10.1016/j.cobeha.2017.02.004Google Scholar
Homberger, D. G. (1986). The lingual apparatus of the African Grey Parrot, Psittacus erithacus Linné (Aves: Psittacidae): Description and theoretical mechanical analysis.Google Scholar
Homberger, D. G. (2006). Classification and Status of Wild Populations of Parrots. In Luescher, A. U. (Ed.), Manual of Parrot Behavior (1st ed.) (pp. 311). Ames, IA: Blackwell Publishing.Google Scholar
Huber, L. & Gajdon, G. K. (2006). Technical intelligence in animals: The kea model. Animal Cognition, 9(4), 295305. https://doi.org/10.1007/s10071-006-0033-8Google Scholar
Hunt, G. R. (1996). Manufacture and use of hook-tools by New Caledonian crows. Nature, 379(6562), 249.Google Scholar
Kabadayi, C., Taylor, L. A., von Bayern, A. M. P., & Osvath, M. (2016). Ravens, New Caledonian crows and jackdaws parallel great apes in motor self-regulation despite smaller brains. Royal Society Open Science, 3(4), 160104. https://doi.org/10.1098/rsos.160104Google Scholar
Kabadayi, C., Krasheninnikova, A., O’Neill, L., van de Weijer, J., Osvath, M., & von Bayern, A. M. P. (2017). Are parrots poor at motor self-regulation or is the cylinder task poor at measuring it? Animal Cognition, 20(6), 11371146. https://doi.org/10.1007/s10071-017-1131-5Google Scholar
Koepke, A. E., Gray, S. L., & Pepperberg, I. M. (2015). Delayed gratification: A Grey parrot (Psittacus erithacus) will wait for a better reward. Journal of Comparative Psychology, 129(4), 339. https://doi.org/10.1037/a0039553Google Scholar
Koops, K., McGrew, W. C., & Matsuzawa, T. (2013). Ecology of culture: Do environmental factors influence foraging tool use in wild chimpanzees, Pan troglodytes verus? Animal Behaviour, 85(1), 175185. https://doi.org/10.1016/j.anbehav.2012.10.022Google Scholar
Koutsos, E. A., Matson, K. D., & Klasing, K. C. (2001). Nutrition of birds in the order Psittaciformes: A review. Journal of Avian Medicine and Surgery, 14(4), 257275.Google Scholar
Krasheninnikova, A. & Schneider, J. M. (2014). Testing problem-solving capacities: Differences between individual testing and social group setting. Animal Cognition, 17(5), 12271232. https://doi.org/10.1007/s10071-014-0744-1Google Scholar
Lambert, M. L., Seed, A. M., & Slocombe, K. E. (2015). A novel form of spontaneous tool use displayed by several captive greater vasa parrots (Coracopsis vasa): Table 1. Biology Letters, 11(12), 20150861. https://doi.org/10.1098/rsbl.2015.0861Google Scholar
Lambert, M. L., Schiestl, M., Schwing, R., Taylor, A. H., Gajdon, G. K., Slocombe, K. E., & Seed, A. M. (2017). Function and flexibility of object exploration in kea and New Caledonian crows. Royal Society Open Science, 4(9), 170652. https://doi.org/10.1098/rsos.170652Google Scholar
Lambert, M. L., Jacobs, I., Osvath, M., & von Bayern, A. M. P. (2018). Birds of a feather? Parrot and corvid cognition compared. Behaviour, 156(5–8), 505594. https://doi.org/10.1163/1568539X-00003527Google Scholar
Laumer, I. B., Bugnyar, T., & Auersperg, A. M. I. (2016). Flexible decision-making relative to reward quality and tool functionality in Goffin cockatoos (Cacatua goffiniana). Scientific Reports, 6(1). https://doi.org/10.1038/srep28380Google Scholar
Laumer, I., Bugnyar, T., Reber, S., & Auersperg, A. (2017). Can hook-bending be let off the hook? Bending/unbending of pliant tools by cockatoos. Procedures of the Royal Society B, 284(1862), 20171026.Google Scholar
Lefebvre, Louis, Whittle, P., Lascaris, E., & Finkelstein, A. (1997). Feeding innovations and forebrain size in birds. Animal Behaviour, 53(3), 549560.Google Scholar
Lefebvre, Louis, Gaxiola, A., Dawson, S., Timmermans, S., Rosza, L., & Kabai, P. (1998). Feeding innovations and forebrain size in Australasian birds. Behaviour, 135(8), 10771097.Google Scholar
Lefebvre, Louis, Nicolakakis, N., & Boire, D. (2002). Tools and brains in birds. Behaviour, 139(7), 939973.Google Scholar
Lefebvre, L., Reader, S. M., & Sol, D. (2004). Brains, innovations and evolution in birds and primates. Brain, Behavior and Evolution, 63(4), 233246. https://doi.org/10.1159/000076784Google Scholar
Liedtke, J., Werdenich, D., Gajdon, G. K., Huber, L., & Wanker, R. (2011). Big brains are not enough: Performance of three parrot species in the trap-tube paradigm. Animal Cognition, 14(1), 143149. https://doi.org/10.1007/s10071-010-0347-4Google Scholar
Linden, P. G., & Luescher, A. U. (2006). Behavioral Development of Psittacine Companions: Neonates, Neophytes, and Fledglings. In Luescher, A. U. (Ed.), Manual of Parrot Behavior (1st ed.) (pp. 93111). Ames, IA: Blackwell Publishing.Google Scholar
MacLean, E. L., Hare, B., Nunn, C. L., Addessi, E., Amici, F., Anderson, R. C., Aureli, F., Baker, J. M., Bania, A. E., Barnard, A. M., Boogert, N. J., Brannon, E. M., Bray, E. E., Bray, J., Brent, L. J. N., Burkart, J. M., Call, J., Cantlon, J. F., Cheke, L. G., … Zhao, Y. (2014). The evolution of self-control. Proceedings of the National Academy of Sciences, 111(20), E2140E2148. https://doi.org/10.1073/pnas.1323533111Google Scholar
Manrique, H. M., Gross, A. N.-M., & Call, J. (2010). Great apes select tools on the basis of their rigidity. Journal of Experimental Psychology: Animal Behavior Processes, 36(4), 409422. https://doi.org/10.1037/a0019296Google Scholar
Manrique, H. M., Sabbatini, G., Call, J., & Visalberghi, E. (2011). Tool choice on the basis of rigidity in capuchin monkeys. Animal Cognition, 14(6), 775786. https://doi.org/10.1007/s10071-011-0410-9Google Scholar
Manrique, H. M., Völter, C. J., & Call, J. (2013). Repeated innovation in great apes. Animal Behaviour, 85(1), 195202. https://doi.org/10.1016/j.anbehav.2012.10.026Google Scholar
Mettke‐Hofmann, C., Winkler, H., & Leisler, B. (2002). The significance of ecological factors for exploration and neophobia in parrots. Ethology, 108(3), 249272.Google Scholar
Mioduszewska, B., O’Hara, M., Haryoko, T., Auersperg, A., Huber, L., & Prawiradilaga, D. M. (2018). Notes on ecology of wild Goffin´s cockatoo in the late dry season with emphasis on feeding ecology. 45, 85102.Google Scholar
Mulawka, E. J. (2014). The Cockatoos: A Complete Guide to the 21 Species. Jefferson, GA: McFarland.Google Scholar
Navarrete, A. F., Reader, S. M., Street, S. E., Whalen, A., & Laland, K. N. (2016). The coevolution of innovation and technical intelligence in primates. Philosophical Transactions of the Royal Society B: Biological Sciences, 371(1690), 20150186. https://doi.org/10.1098/rstb.2015.0186Google Scholar
O’Hara, M., Huber, L., & Gajdon, G. K. (2015). The advantage of objects over images in discrimination and reversal learning by kea, Nestor notabilis. Animal Behaviour, 101, 5160. https://doi.org/10.1016/j.anbehav.2014.12.022Google Scholar
O’Hara, M., Mioduszewska, B., von Bayern, A., Auersperg, A., Bugnyar, T., Wilkinson, A., Huber, L., & Gajdon, G. K. (2017). The temporal dependence of exploration on neotic style in birds. Scientific Reports, 7(1). https://doi.org/10.1038/s41598-017-04751-0Google Scholar
O’Hara, M., Mioduszewska, B., Haryoko, T., Prawiradilaga, D. M., Huber, L., & Auersperg, A. (2018). Extraction without tooling around: The first comprehensive description of the foraging- and socio-ecology of wild Goffin’s cockatoos (Cacatua goffiniana). Behaviour, 156(5–8), 661690) https://doi.org/10.1163/1568539X-00003523Google Scholar
Olkowicz, S., Kocourek, M., Lučan, R. K., Porteš, M., Fitch, W. T., Herculano-Houzel, S., & Němec, P. (2016). Birds have primate-like numbers of neurons in the forebrain. Proceedings of the National Academy of Sciences, 113(26), 72557260. https://doi.org/10.1073/pnas.1517131113Google Scholar
O’Neill, L., Picaud, A., Maehner, J., Gahr, M., & von Bayern, A. M. (2018). Two macaw species can learn to solve an optimised two-trap problem, but without functional causal understanding. Behaviour, 156(5–8), 691720. https://doi.org/10.1163/1568539X-00003521Google Scholar
Osuna-Mascaró, A. J. & Auersperg, A. M. I. (2018). On the brink of tool use? Could object combinations during foraging in a feral Goffin’s cockatoo (Cacatua goffiniana) result in tool innovations? Animal Behavior and Cognition, 5(2), 229234. https://doi.org/10.26451/abc.05.02.05.2018Google Scholar
Parker, S. T. & Gibson, K. R. (1977). Object manipulation, tool use and sensorimotor intelligence as feeding adaptations in Cebus monkeys and great apes. Journal of Human Evolution, 6(7), 623641.Google Scholar
Pepperberg, I. M. (2002). The Alex Studies: Cognitive and Communicative Abilities of Grey Parrots. Cambridge, MA: Harvard University Press.Google Scholar
Power, T. G. (1999). Play and Exploration in Children and Animals. Hove, UK: Psychology Press.Google Scholar
Ramsey, G., Bastian, M. L., & van Schaik, C. (2007). Animal innovation defined and operationalized. Behavioral and Brain Sciences, 30(04). https://doi.org/10.1017/S0140525X07002373Google Scholar
Reader, S. M. & Laland, K. N. (2001). Primate innovation: Sex, age and social rank differences. International Journal of Primatology, 19.Google Scholar
Reader, S. M. & Laland, K. N. (2002). Social intelligence, innovation, and enhanced brain size in primates. Proceedings of the National Academy of Sciences, 22(5), 787-805. https://doi.org/10.1073/pnas.062041299Google Scholar
Reader, S. M. & Laland, K. N. (Eds.) (2003). In Reader, S. M. & Laland, K. N. (Eds.), Animal Innovation (pp. 3961). New York: Oxford University Press.Google Scholar
Rössler, T., Mioduszewska, B., O’Hara, M., Huber, L., Prawiradilaga, D. M., & Auersperg, A. M. I. (2020). Using an innovation arena to compare wild-caught and laboratory Goffin´s cockatoosScientific Reports, 10(1), 8681. https://doi.org/10.1038/s41598-020-65223-6Google Scholar
Russell, P. (1973). Relationships between exploratory behaviour and fear: A review. British Journal of Psychology, 64(3), 417433.Google Scholar
Sayol, F., Downing, P. A., Iwaniuk, A. N., Maspons, J., & Sol, D. (2018). Predictable evolution towards larger brains in birds colonizing oceanic islands. Nature Communications, 9(1). https://doi.org/10.1038/s41467-018-05280-8Google Scholar
Schneider, L., Serbena, A. L., & Guedes, N. M. (2002). Manipulação de frutos de acuri e bocaiúva por araras-azuis no Pantanal Sul, XX; Encontro Anual de Etologia, 378.Google Scholar
Schuck-Paim, C., Alonso, W. J., & Ottoni, E. B. (2008). Cognition in an ever-changing world: Climatic variability is associated with brain size in neotropical parrots. Brain, Behavior and Evolution, 71(3), 200215. https://doi.org/10.1159/000119710Google Scholar
Schwing, R., Weber, S., & Bugnyar, T. (2017). Kea (Nestor notabilis) decide early when to wait in food exchange task. Journal of Comparative Psychology, 131(4), 269276. https://doi.org/10.1037/com0000086Google Scholar
Seibert, L. M. (2006). Social Behavior of Psittacine Birds. In Luescher, A. U. (Ed.), Manual of Parrot Behavior (1st ed.) (pp. 4348). Ames, IA: Blackwell.Google Scholar
Shumaker, R. W., Walkup, K. R., & Beck, B. B. (2011). Animal Tool Behavior: The Use and Manufacture of Tools by Animals. Baltimore, MD: Johns Hopkins University Press.Google Scholar
Szabo, B., Bugnyar, T., & Auersperg, A. M. I. (2017). Within-group relationships and lack of social enhancement during object manipulation in captive Goffin’s cockatoos (Cacatua goffiniana). Learning & Behavior, 45(1), 719. https://doi.org/10.3758/s13420-016-0235-0Google Scholar
Tebbich, S., Griffin, A. S., Peschl, M. F., & Sterelny, K. (2016). From mechanisms to function: An integrated framework of animal innovation. Philosophical Transactions of the Royal Society B: Biological Sciences, 371(1690), 20150195. https://doi.org/10.1098/rstb.2015.0195Google Scholar
Timmermans, S., Lefebvre, L., Boire, D., & Basu, P. (2000). Relative size of the hyperstriatum ventrale is the best predictor of feeding innovation rate in birds. Brain, Behavior and Evolution, 56(4), 196203.Google Scholar
Tokita, M. (2003). The skull development of parrots with special reference to the emergence of a morphologically unique cranio-facial hinge. Zoological Science, 20(6), 749758.Google Scholar
van Horik, J. O. & Madden, J. R. (2016). A problem with problem solving: Motivational traits, but not cognition, predict success on novel operant foraging tasks. Animal Behaviour, 114, 189198. https://doi.org/10.1016/j.anbehav.2016.02.006Google Scholar
van Horik, J. O. & Emery, N. J. (2018). Serial reversal learning and cognitive flexibility in two species of Neotropical parrots (Diopsittaca nobilis and Pionites melanocephala). Behavioural Processes, 157, 664672. https://doi.org/10.1016/j.beproc.2018.04.002Google Scholar
van Schaik, C. P., Burkart, J., Damerius, L., Forss, S. I. F., Koops, K., van Noordwijk, M. A., & Schuppli, C. (2016). The reluctant innovator: Orangutans and the phylogeny of creativity. Philosophical Transactions of the Royal Society B: Biological Sciences, 371(1690), 20150183. https://doi.org/10.1098/rstb.2015.0183Google Scholar
Vick, S.-J., Bovet, D., & Anderson, J. R. (2010). How do African grey parrots (Psittacus erithacus) perform on a delay of gratification task? Animal Cognition, 13(2), 351358.Google Scholar
Wallace, A. R. (2015). The Malay Archipelago: The land of the orang-utan, and the bird of paradise: A narrative of travel, with studies of man and nature (3rd ed.). Oxford: John Beaufoy Publishing.Google Scholar
Wood, G. (1984). Tool use by the palm cockatoo Probosciger aterrimus during display. Corella, 8(4), 9495.Google Scholar
Wood, G. (1988). Further field observations of the palm cockatoo Probosciger aterrimus in the Cape York Peninsula, Queensland. Corella, 12, 4852.Google Scholar

References

Abe, K. (2010). Interaction between body and environment in creative thinkingCognitive Studies: Bulletin of the Japanese Cognitive Science Society17, 599610.Google Scholar
Amabile, T. M. (1996). Creativity in Context: Update to “The Social Psychology of Creativity.” Boulder, CO: Westview Press.Google Scholar
Auersperg, A. M. I., von Bayern, A. M. I., Weber, S., Szabadvari, A., Bugnyar, T., & Kacelnik, A. (2014). Social transmission of tool use and tool manufacture in Goffin cockatoos (Cacatua goffini). Proceedings of the Royal Society B: Biological Sciences, 281(1793), 2014097220140972. https://doi.org/10.1098/rspb.2014.0972Google Scholar
Cantor, M. & Whitehead, H. (2013). The interplay between social networks and culture: Theoretically and among whales and dolphins. Philosophical Transactions of the Royal Society of London. Series B, Biological Sciences, 368(1618), 20120340. https://doi.org/10.1098/rstb.2012.0340Google Scholar
Ciardelli, L. E., Weiss, A., Powell, D. M., & Reiss, D. (2017). Personality dimensions of the captive California sea lion (Zalophus californianus). Journal of Comparative Psychology, 131(1), 5058. https://doi.org/10.1037/com0000054Google Scholar
Costa, P. T. & McCrae, R. R. (1992). Normal personality assessment in clinical practice: The neo personality inventory. Psychological Assessment, 4(1), 513. https://doi.org/10.1037/1040-3590.4.1.5Google Scholar
Goodall, J. (1986). The Chimpanzees of Gombe. Cambridge, MA: Harvard University Press.Google Scholar
Hall, K. R. L. & Schaller, G. B. (1964). Tool-using behavior of the California sea otter. Journal of Mammalogy, 45(2), 287298. https://doi.org/10.2307/1376994Google Scholar
Herrmann, E., Hernández-Lloreda, M. V., Call, J., Hare, B. A., & Tomasello, M. (2010). The structure of individual differences in the cognitive abilities of children and chimpanzees. Psychological Science, 21(1), 102110. https://doi.org/10.1177/0956797609356511Google Scholar
Herrmann, E. & Call, J. (2012). Are there geniuses among the apes? Philosophical Transactions of the Royal Society of London. Series B, Biological Sciences, 367(1603), 27532761. https://doi.org/10.1098/rstb.2012.0191Google Scholar
Highfill, L. E. & Kuczaj, S. A. (2007). Do bottlenose dolphins (Tursiops truncatus) have distinct and stable personalities? Aquatic Mammals, 33(3), 380389.Google Scholar
Higuchi, H. (1988). Individual differences in bait‐fishing by the green‐backed heron Ardeola striata associated with territory quality. Ibis, 130(1), 3944. https://doi.org/10.1111/j.1474-919X.1988.tb00953.xGoogle Scholar
Hill, H. M., Kahn, M. S., Brilliott, L. J., Roberts, B. M., Gutierrez, C., & Artz, S. (2011). Beluga (Delphinapterus leucas) bubble bursts: Surprise, protection, or play? International Journal of Comparative Psychology, 24, 235243. Retrieved from http://comparativepsychology.org/ijcp-2011-2/05.Hill_PDF.pdfGoogle Scholar
Huffman, M. & Hirata, S. (2003). Biological and Ecological Foundations of Primate Behavioral Traditions. The Biology of Traditions: Models and Evidence. In Fragaszy, D. M. & Perry, S (Eds.), The Biology of Traditions (pp. 267296). Cambridge, England: Cambridge University Press.Google Scholar
Jones, B. L. & Kuczaj, S. A. (2014). Beluga (Delphinapterus leucas) novel bubble helix play behavior. Animal Behavior and Cognition, 2(2), 206. https://doi.org/10.12966/abc.05.10.2014Google Scholar
Kako, E. (1999). Elements of syntax in the systems of three language-trained animals. Animal Learning & Behavior, 27(1), 114.Google Scholar
Kaufman, J. C. (2016). Creativity 101 (2nd ed.). New York: Springer.Google Scholar
Kaufman, J. C., Baer, J., Cole, J. C., & Sexton, J. D. (2008). A comparison of expert and nonexpert raters using the consensual assessment technique. Creativity Research Journal, 20(2), 171178. https://doi.org/10.1080/10400410802059929Google Scholar
Kaufman, A. B. & Rosenthal, R. (2009). Can you believe my eyes? The importance of interobserver reliability statistics in observations of animal behaviour. Animal Behaviour, 78(6), 14871491. https://doi.org/10.1016/j.anbehav.2009.09.014Google Scholar
Kaufman, J. C., Baer, J., & Cole, J. C. (2009). Expertise, domains, and the consensual assessment technique. The Journal of Creative Behavior, 43(4), 223233. https://doi.org/10.1002/j.2162-6057.2009.tb01316.xGoogle Scholar
Kaufman, J. C. & Baer, J. (2012). Beyond new and appropriate: Who decides what is creative? Creativity Research Journal, 24(1), 8391. https://doi.org/10.1080/10400419.2012.649237Google Scholar
Kaufman, J. C., Baer, J., Cropley, D. H., Reiter-Palmon, R., & Sinnett, S. (2013). Furious activity vs. understanding: How much expertise is needed to evaluate creative work? Psychology of Aesthetics, Creativity, and the Arts, 7, 332340.Google Scholar
Kaufman, A. B., Reynolds, M. R., & Kaufman, A. S. (2019). The structure of ape (Hominoidea) intelligence. Journal of Comparative Psychology, 133(1), 92105. https://doi.org/10.1037/com0000136Google Scholar
Kopps, A. M., Ackermann, C. Y., Sherwin, W. B., Allen, S. J., Bejder, L., & Krützen, M. (2014). Cultural transmission of tool use combined with habitat specializations leads to fine-scale genetic structure in bottlenose dolphins. Proceedings of the Royal Society B: Biological Sciences, 281(1782), 2013324520133245. https://doi.org/10.1098/rspb.2013.3245Google Scholar
Krützen, M., Mann, J., Heithaus, M. R., Connor, R. C., Bejder, L., … Sherwin, W. B. (2005). Cultural transmission of tool use in bottlenose dolphins. Proceedings of the National Academy of Sciences, 102(25), 89398943. https://doi.org/10.1073/pnas.0500232102Google Scholar
Krützen, M., Kreicker, S., Macleod, C. D., Learmonth, J., Kopps, A. M., Walsham, P., & Allen, S. J. (2014). Cultural transmission of tool use by Indo-Pacific bottlenose dolphins (Tursiops sp.) provides access to a novel foraging niche. Proceedings. Biological Sciences/The Royal Society, 281(1784), 20140374. https://doi.org/10.1098/rspb.2014.0374Google Scholar
Kuczaj, S. A. & Highfill, L. E. (2005). Dolphin play: Evidence for cooperation and culture? Behavioral and Brain Sciences, 28(5), 705706.Google Scholar
Kuczaj, S. A., Paulos, R. D. R. D., & Ramos, J. A. (2005). Imitation in Apes, Children, and Dolphins: Implications for the Ontogeny and Phylogeny of Symbolic Representation. In Namy, L. A (Ed.), Symbol Use and Symbolic Representation: Developmental and Comparative Perspectives (p. 221). Retrieved from http://scholar.google.com/scholar?hl=en&btnG=Search&q=intitle:Imitation+in+apes,+children,+and+dolphins:+Implications+for+the+Ontogeny+and+Phylogeny+of+Symbolic+Representation#0Google Scholar
Kuczaj, S. A., Makecha, R., Trone, M., Paulos, R. D., & Ramos, J. A. (2006). Role of peers in cultural innovation and cultural transmission: Evidence from the play of dolphin calves. International Journal of Comparative Psychology, 19(2), 223240. Retrieved from http://escholarship.org/uc/item/4pn1t50s.pdfGoogle Scholar
Kuczaj, S. A. & Makecha, R. N. (2008). The Role of Play in the Evolution and Ontogeny of Contextually Flexible Communication. In Oller, D. K. & Griebel, U. (Eds.), Evolution of Communicative Flexibility: Complexity, Creativity, and Adaptability in Human and Animal Communication (pp. 253277). Cambridge, MA: MIT Press.Google Scholar
Lawrence, M. K., Borger-Turner, J. L., Turner, T. N., & Eskelinen, H. C. (2016). Investigating the effects of applied learning principles on the “Create” response in Atlantic bottlenose dolphins (Tursiops truncatus). International Journal of Comparative Psychology, 29(1), 111.Google Scholar
Leung, A. K. Y., Kim, S., Polman, E., Ong, L. S., Qiu, L., Goncalo, J. A., & Sanchez-Burks, J. (2012). Embodied metaphors and creative “acts.” Psychological Science, 23, 502509.Google Scholar
McCowan, B., Marino, L., Vance, E., Walke, L., & Reiss, D. (2000). Bubble ring play of bottlenose dolphins (Tursiops truncatus): Implications for cognition. Journal of Comparative Psychology. https://doi.org/10.1037/0735-7036.114.1.98Google Scholar
McCrae, R. R. & Costa, P. T. (1997). Personality trait structure as a human universal. American Psychologist, 52(5), 509516. https://doi.org/10.1037/0003-066X.52.5.509Google Scholar
Montgomery, S. H. (2014). The relationship between play, brain growth and behavioural flexibility in primates. Animal Behaviour, 90, 281286. https://doi.org/10.1016/j.anbehav.2014.02.004Google Scholar
Paulos, R. D., Trone, M., & Kuczaj, S. A. (2010). Play in wild and captive cetaceans. International Journal of Comparative Psychology, 23(4), (701722).Google Scholar
Perals, D., Griffin, A. S., Bartomeus, I., & Sol, D. (2017). Revisiting the open-field test: What does it really tell us about animal personality? Animal Behaviour, 123, 6979. https://doi.org/10.1016/j.anbehav.2016.10.006Google Scholar
Pryor, K. W. (2014). Historical perspectives: A dolphin journey. Aquatic Mammals, 40(1), 104114.Google Scholar
Pryor, K. W., Haag, R., & O’Reilly, J. (1969). The creative porpoise: Training for novel behavior. Journal of the Experimental Analysis of Behavior, 12(4), 653. Retrieved from www.ncbi.nlm.nih.gov/pmc/articles/PMC1338662/Google Scholar
Ramsey, G., Bastian, M. L., & van Schaik, C. (2007). Animal innovation defined and operationalized. The Behavioral and Brain Sciences, 30(4), 393407; discussion 407–432. https://doi.org/10.1017/S0140525X07002373Google Scholar
Rendell, L. E. & Whitehead, H. (2001). Culture in whales and dolphins. Behavioral and Brain Sciences, 24(2), 309382.Google Scholar
Rendell, L. E. & Whitehead, H. (2003). Comparing repertoires of sperm whale codas: A multiple methods approach. Bioacoustics, 14(1), 6181.Google Scholar
Rendell, L. E., Mesnick, S. L., Dalebout, M. L., Burtenshaw, J., & Whitehead, H. (2011). Can genetic differences explain vocal dialect variation in sperm whales, Physeter macrocephalus? Behavior Genetics, 42(2), 332343. https://doi.org/10.1007/s10519-011-9513-yGoogle Scholar
Savage-Rumbaugh, E. S. S., Murphy, J., Sevcik, R. A. A., Brakke, K. E., Williams, S. L. L., Rumbaugh, D. M. M., & Bates, E. (1993). Language comprehension in ape and child. Monographs of the Society for Research in Child Development, 58(3/4), 1252. Retrieved from www.jstor.org/stable/1166068Google Scholar
Soto, C. J., & John, O. P. (2016). The next big five inventory (BFI-2): Developing and assessing a hierarchical model with 15 facets to enhance bandwidth, fidelity, and predictive power. Journal of Personality and Social Psychology, 113, 117143. https://doi.org/10.1037/pspp0000096Google Scholar
Sternberg, R. J. (1985). Implicit theories of intelligence, creativity, and wisdom. Journal of Personality and Social Psychology, 49, 607627.Google Scholar
Torrance, E. P. (1974). Torrance Tests of Creative Thinking. Lexington, MA: Ginn.Google Scholar
Torrance, E. P. (2008). Torrance Tests of Creative Thinking: Norms-Technical Manual, Verbal Forms A and B. Bensenville, IL: Scholastic Testing Service.Google Scholar
Walsh, R. N. & Cummins, R. A. (1976). The open-field test: A critical review. Psychological Bulletin, 83, 482504. https://doi.org/10.1037/0033-2909.83.3.482Google Scholar

References

Barrett, B. J., Monteza-Moreno, C. M., Dogandzic, T., Zwyns, N., Ibáñez, A., & Crofoot, M. C. (2018). Habitual stone-tool-aided extractive foraging in white-faced capuchins, Cebus capucinus. Royal Society Open Science, 5, 181002. doi: 10.1098/rsos.181002.Google Scholar
Cardoso, R. M. & Ottoni, E.B. (2016). The effects of tradition on problem solving by two wild populations of bearded capuchin monkeys in a probing task. Biology Letters, 12, 20160604. doi: 10.1098/rsbl.2016.0604Google Scholar
Coelho, C. G., Falótico, T., Izar, P., Mannu, M., Resende, B. D., Siqueira, J. O., & Ottoni, E. B. (2015). Social learning strategies for nut-cracking by tufted capuchin monkeys (Sapajus spp.). Animal Cognition, 18, 911919. doi: 10.1007/s10071-015-0861-5Google Scholar
Falótico, T. & Ottoni, E. B. (2013). Stone throwing as a sexual display in wild female bearded capuchin monkeys, Sapajus libidinosus. PLoS ONE, 8, e79535. doi: 10.1371/journal.pone.0079535Google Scholar
Falótico, T. & Ottoni, E. B. (2014). Sexual bias in probe tool manufacture and use by wild bearded capuchin monkeys. Behavioural Processes, 108, 117122. doi: 10.1016/j.beproc.2014.09.036Google Scholar
Falótico, T., Proffitt, T., Ottoni, E. B., Staff, R. A., & Haslam, M. (2019) Three thousand years of wild capuchin stone tool use. Nature Ecology & Evolution, 3, 10341038. doi. org/10.1038/s41559–019-0904-4Google Scholar
Fragaszy, D. M. (2003). Making space for traditions. Evolutionary Anthropology, 12, 6170. doi: 10.1002/evan.10104Google Scholar
Fragaszy, D. M., Visalberghi, E., & Fedigan, L. M. (2004a). The Complete Capuchin: The Biology of the Genus Cebus. Cambridge, UK: Cambridge University Press.Google Scholar
Fragaszy, D. M., Izar, P., Visalberghi, E., Ottoni, E. B., & Oliveira, M. (2004b). Wild capuchin monkeys use anvils and stone pounding tools. American Journal of Primatology, 64, 359366. doi: 10.1002/ajp.20085Google Scholar
Fragaszy, D. M., Biro, D., Eshchar, Y., Humle, T., Izar, P., Resende, B., & Visalberghi, E. (2013). The fourth dimension of tool use: Temporally enduring artefacts aid primates learning to use tools. Philosophical Transactions of the Royal Society B: Biological Sciences, 368, 20120410. doi: 10.1098/rstb.2012.0410Google Scholar
Galef, B. (2003). Social Learning: Promoter or Inhibitor of Innovation? In Reader, S. M. & Laland, K. N. (Eds.), Animal Innovation, ch.6 (pp. 137152). Oxford, UK: Oxford University Press.Google Scholar
Haslam, M., Luncz, L. V., Staff, R. A., Bradshaw, F., Ottoni, E. B., & Falótico, T. (2016). Pre-Columbian Monkey Tools. Current Biology, 26, R515R522. doi: 10.1016/j.cub.2016.05.046Google Scholar
Itani, J. & Nishimura, A. (1973). The Study of Infrahuman Culture in Japan: A Review. In Menzel, E. W. (Ed.), Precultural Primate Behavior (pp. 2650). Basel, Germany: Karger.Google Scholar
Kummer, H. & Goodall, J. (1985). Conditions of innovative behaviour in primates. Philosophical Transactions of the Royal Society B: Biological Sciences, 308, 203214. doi: 10.1098/rstb.1985.0020Google Scholar
Kendal, R., Coelho, C., Corat, C., & Ottoni, E. B. (2017). Experimental field investigations of cultural capacities in the tool-using bearded capuchin (Sapajus libidinosus). Abstracts of the 7th European Federation for Primatology Meeting. Folia Primatologica, 88, 128. doi: 10.1159/000479094Google Scholar
Laland, K. N. & Reader, S. M. (2010). Innovation in Animals. In Breed, M. D. & Moore, J. (Eds.), Encyclopaedia of Animal Behaviour (pp. 150154). Oxford, UK: Elsevier.Google Scholar
Lee, P. C. & Moura, A. C. A. (2015). Necessity, Unpredictability and Opportunity: An Exploration of Ecological and Social Drivers of Behavioral Innovation. In Kaufman, A. B. & Kaufman, J. C. (Eds.), Animal Creativity and Innovation (pp. 217233). San Diego, CA: Academic Press. doi: 10.1016/B978-0-12-800648-1.00011-5Google Scholar
Lynch Alfaro, J. W., Boubli, J.P., Olson, L.E., Di Fiore, A., Wilson, B., Gutierrez-Espeleta, G. A. … & Alfaro, M. E. (2012). Explosive Pleistocene range expansion leads to widespread Amazonian sympatry between robust and gracile capuchin monkeys. Journal of Biogeography, 39, 272288. doi: 10.1111/j.1365-2699.2011.02609.xGoogle Scholar
Mannu, M. & Ottoni, E. B. (2009). The enhanced tool-kit of two groups of wild bearded capuchin monkeys in the Caatinga: Tool making, associative use, and secondary tools. American Journal of Primatology, 71, 242251. doi: 10.1002/ajp.20642Google Scholar
Matsuzawa, T., Humle, T., & Sugiyama, Y. (Eds.) (2011). The Chimpanzees of Bossou and Nimba. Tokyo: Springer.Google Scholar
Ottoni, E. B., Resende, B. D., & Izar, P. (2005). Watching the best nutcrackers: What capuchin monkeys (Cebus apella) know about others´ tool using skills. Animal Cognition, 8, 215219. doi: 10.1007/s10071-004-0245-8Google Scholar
Ottoni, E. B. & Izar, P. (2008). Capuchin monkey tool use: Overview and implications. Evolutionary Anthropology, 17, 171178. doi: 10.1002/evan.20185Google Scholar
Perry, S. & Manson, J. H. (2008). Manipulative Monkeys: The Capuchins of Lomas Barbudal. Cambridge, MA: Harvard University Press.Google Scholar
Ramsey, G., Bastian, M. L., & van Schaik, C. (2007). Animal innovation defined and operationalized. Behavioral and Brain Sciences, 30, 393437. doi: 10.1017/S0140525X07002373.Google Scholar
Shumaker, R. W., Walkup, K. R., & Beck, B. B. (2011). Animal Tool Behavior. Baltimore, MD: Johns Hopkins University Press.Google Scholar
Souto, A., Bione, C. B. C., Bastos, M., Bezerra, B. M., Fragaszy, D., & Schiel, N. (2011). Critically endangered blonde capuchins fish for termites and use new techniques to accomplish the task. Biology Letters, 7, 532535. doi: 10.1098/rsbl.2011.0034.Google Scholar
Spagnoletti, N., Visalberghi, E., Verderane, M.P., Ottoni, E., Izar, P., & Fragaszy, D. (2012). Stone tool use in wild bearded capuchin monkeys, Cebus libidinosus. Is it a strategy to overcome food scarcity? Animal Behaviour, 83, 12851294. doi: 10.1016/j.anbehav.2012.03.002Google Scholar
Tebbich, S., Griffin, A.S., Peschl, M.F., & Sterelny, K. (2016). From mechanisms to function: An integrated framework of animal innovation. Philosophical Transactions of the Royal Society B, 371, 20150195. doi: 10.1098/rstb.2015.0195.Google Scholar
Visalberghi, E. & Limongelli, L. (1994). Lack of comprehension of cause-effect relations in tool-using capuchin monkeys (Cebus apella). Journal of Comparative Psychology, 108, 1522. doi: 10.1037/0735-7036.108.1.15Google Scholar
Visalberghi, E., Fragaszy, D. M., Izar, P., & Ottoni, E. B. (2005). Terrestriality and tool use. Science, 308, 951. doi: 10.1126/science.308.5724.951cGoogle Scholar
Westergaard, G. C. & Fragaszy, D. M. (1987). The manufacture and use of tools by capuchin monkeys (Cebus apella). Journal of Comparative Psychology, 101, 159168. doi: 10.1037/0735-7036.101.2.159Google Scholar
Whiten, A., Goodall, J., McGrew, W. C., Nishida, T., Reynolds, V., Sugiyama, Y. … & Boesch, C. (1999). Cultures in chimpanzees. Nature, 399, 682685. doi: 10.1038/21415Google Scholar

References

Bastian, M. L., van Noordwijk, M. A., & van Schaik, C. P. (2012). Innovative behaviors in wild Bornean orangutans revealed by targeted population comparison. Behaviour, 149(2012) 275297.Google Scholar
Bebko, A. O. (2012). Route-Based Travel Networks in Wild Orangutans. MA Thesis, Psychology, York University, Toronto, CanadaGoogle Scholar
Boesch, C. & Boesch-Achermann, H. (2000). The Chimpanzees of the Taï Forest: Behavioural Ecology and Evolution. Oxford: Oxford University Press.Google Scholar
Bertinetto, A. (2012). Performing the unexpected: Improvisation and artistic creativity. Revista Internacional de Filosofıa, 57, 117135.Google Scholar
Byrne, R. W. (1999). Imitation without intentionality: Using string parsing to copy the organization of behavior. Animal Cognition, 2, 6372.Google Scholar
Byrne, R. W. (2003). Imitation as behaviour parsing. Philosophical Transactions: Biological Sciences, 358, 529536.Google Scholar
Byrne, R. W. (2004). The Manual Skills and Cognition that Lie Behind Hominid Tool Use. In Russon, A. E. & Begun, D. R. (Eds.), The Evolution of Thought: Evolutionary Origins of Great Ape Intelligence (pp. 3144). Cambridge, UK: Cambridge University Press,.Google Scholar
Byrne, R. W. & Byrne, J. M. E. (1993). Complex leaf-gathering skills of mountain gorillas (Gorilla g. beringei): Variability and standardization. American Journal of Primatology, 31, 241261.Google Scholar
Byrne, R. W. & Russon, A. E. (1998). Learning by imitation: A hierarchical approach. Behavioral and Brain Sciences, 21, 667721.Google Scholar
Byrne, R. W., Corps, N., & Byrne, J. M. E. (2001a). Estimating the complexity of animal behavior: How mountain gorillas eat thistles. Behaviour, 138, 525557.Google Scholar
Byrne, R. W., Corp, N., & Byrne, J. M. E. (2001b). Manual dexterity in the gorilla: Bimanual and digit role differentiation in a natural task. Animal Cognition, 4, 347361.Google Scholar
Call, J. (2015). Conservatism versus Innovation: The Great Ape Story. In Kaufman, A. B. & Kaufman, J. C. (Eds.), Animal Creativity and Innovation (pp. 397414). San Diego, CA: Academic Press.Google Scholar
Cropley, D. H. (2015). Commentary on Chapter 15: Tools for the Trees: Orangutan Arboreal Tool Use and Creativity. In Kaufman, A. B. & Kaufman, J. C. (Eds.), Animal Creativity and Innovation (pp. 456458). San Diego, CA: Academic Press.Google Scholar
Damerius, L. A., Graber, S. M., Willems, E. P., & van Schaik, C. P. (2017a). Curiosity boosts orang-utan problem-solving ability. Animal Behaviour, 134, 5770. https://doi.org/10.1016/j.anbehav.2017.10.005Google Scholar
Damerius, L. A., Forss, S. I. F., Kosonen, Z. K., Willems, E. P., Burkart, J. M., Call, J., Galdikas, B. M. F, Liebal, K., Haun, D. B. M, & van Schaik, D. P. (2017b). Orientation to humans predicts cognitive performance in orangutans. Scientific Reports, 7, 40052. doi: 10.1038/srep40052Google Scholar
Deaner, R. O., van Schaik, C. P., & Johnson, V. (2006). Do some taxa have better domain-general cognition than others? A meta-analysis of nonhuman primate studies. Evolutionary Psychology, 4, 149196.Google Scholar
Delgado, R. A. & van Schaik, C. P. (2000). The behavioral ecology and conservation of the orangutan (Pongo pygmaeus): A tale of two islands. Evolutionary Anthropology, 9 (5), 201-218.Google Scholar
Forss, S. I. F., Schuppli, C., Haiden, D., Zweifel, N., & van Schaik, C. P. (2015). Contrasting responses to novelty by wild and captive orangutans. American Journal of Primatology, 77, 11091121. (doi: 10.1002/ajp.22445)Google Scholar
Fox, E. A., Sitompul, A. F., & van Schaik, C. P. (1999). Intelligent Tool Use in Wild Sumatran Orangutans. In Parker, S. T., Mitchell, R. W., & Miles, H. L. (Eds.), The Mentalities of Gorillas and Orangutans (pp. 99116). Cambridge UK: Cambridge University Press.Google Scholar
Gibson, K. R. (1993). Tool Use, Language and Social Behaviour in Relationship to Information Processing Capacities. In Gibson, K. R. & Ingold, T. (Eds.), Tools, Language and Cognition in Human Evolution (pp. 251269). Cambridge, UK: Cambridge University Press,.Google Scholar
Gottlieb, J., Oudeyer, P.-Y., Lopes, M., & Barnes, A. (2013). Information-seeking, curiosity, and attention: Computational and neural mechanisms. Trends in Cognitive Science, 17(11). https://doi.org/10.1016/j.tics.2013.09.001Google Scholar
Griffin, A. S. (2016). Innovation as an emergent property: A new alignment of comparative and experimental research on animal innovation. Philosophical Transactions of the Royal Society B, 371, 20150544. http://dx.doi.org/10.1098/rstb.2015.0544Google Scholar
Gruber, T., Singleton, I., & van Schaik, C. (2012). Sumatran orangutans differ in their cultural knowledge but not their cognitive abilities. Current Biology, 22, 22312235. http://dx.doi.org/10.1016/j.cub.2012.09.041Google Scholar
Herzfeld, C. (2016). Wattana: An Orangutan in Paris. Translated from French by Martin, O. Y. & Martin, R. D.. Chicago: University of Chicago Press.Google Scholar
Huffman, M. (1997). Current evidence for self-medication in primates: A multidisciplinary approach. Yearbook of Physical Anthropology, 40, 171200.Google Scholar
Huffman, M & Hirata, S (2003). Biological and Ecological Foundations of Primate Behavioral Traditions. In Fragaszy, D. M. & Perry, S. (Eds.) The Biology of Traditions: Models and Evidence (pp. 267296). Cambridge, UK: Cambridge University Press.Google Scholar
Kummer, H. & Goodall, J. (1985). Conditions of innovative behaviour in primates. Philosophical Transactions of the Royal Society London B, 308, 203214. doi: 10.1098/rstb.1985.0020Google Scholar
Lehner, S. R., Burkart, J. M., & van Schaik, C. P. (2010). An evaluation of the geographic method for recognizing innovations in nature, using zoo orangutans. Primates, 51, 101118.Google Scholar
Lehner, S. R., Burkart, J. M., & van Schaik, C. P. (2011). Can captive orangutans (Pongo pygmaeus abelii) be coaxed into cumulative build-up of techniques? Journal of Comparative Psychology, 125, 446455. doi: 10.1037/a0024413Google Scholar
Manrique, H. M., Volter, C. J., & Call, J. (2013). Repeated innovation in great apes. Animal Behaviour, 85, 195202.Google Scholar
Marshall, A. J., Ancrenaz, M., Brearley, F. Q., Fredriksson, G. M., Ghaffar, N., Heydon, M., Husson, S. J., Leighton, M., McConkey, K. R., Morrogh-Bernard, H. C., Proctor, J., van Schaik, C. P., Yeager, C. P., & Wich, S. A. (2009). The Effects of Forest Phenology and Floristics on Populations of Bornean and Sumatran Orangutans. In Wich, S. A., Utami, S. A., Mitra, S. T., van Schaik, P. (Eds.), Orangutans: Geographic Variation in Behavioral Ecology and Conservation (pp. 97117). Oxford, UK: Oxford University Press.Google Scholar
Matsuzawa, T. (1996). Chimpanzee Intelligence in Nature and Captivity: Isomorphism of Symbol Use and Tool Use. In McGrew, W. C., Marchant, L. F., & Nishida, T. (Eds.), Great Ape Societies (pp. 196209). Cambridge, UK: Cambridge University Press.Google Scholar
McGrew, W. (1992). The Intelligent Use of Tools: Twenty Propositions. In Gibson, K. R. & Ingold, T. (Eds.), Tools, Language and Intelligence: Evolutionary Implications (pp. 151170). Cambridge: Cambridge University Press.Google Scholar
Morrogh-Bernard, H. (2008). Fur-rubbing as a form of self-medication in Pongo pygmaeus. International Journal of Primatology. doi: 10.1007/s10764–008-9266-5Google Scholar
Nater, A., Mattle-Greminger, M. P., Nurcahyo, A., Nowak, M. G., et al. (2017). Morphometric, behavioral, and genomic evidence for a new orangutan species. Current Biology. doi: 10.1016/j.cub.2017.09Google Scholar
Parker, S. T. & McKinney, M. L. (1999). Origins of Intelligence: The Evolution of Cognitive Development in Monkeys, Apes, and Humans. Baltimore, MD: Johns Hopkins University Press.Google Scholar
Povinelli, D. J. & Cant, J. G. H. (1995). Arboreal clambering and the evolution of self conception. Quarterly Review of Biology, 70, 393421.Google Scholar
Ramsey, G., Bastian, M. L., & van Schaik, C. P. (2007). Animal innovation defined and operationalized. Behavioral and Brain Sciences, 30, 393437.Google Scholar
Reader, S. M. (2016). Animal and human innovation: Novel problems and novel solutions. Philosophical Transactions of the Royal Society B, 371, 20150182. http://dx.doi.org/10.1098/rstb.2015.0182Google Scholar
Reader, S. M. & Laland, K. N. (2001). Primate innovation: Sex, age and social rank differences. International Journal of Primatology, 22(5), 787805.Google Scholar
Reader, S. M. & Laland, K. N. (2002) Social intelligence, innovation and enhanced brain size in primates. Procedures of the National Academy of Science USA, 99, 44364441. doi: 10. 1073/pnas.062041299Google Scholar
Reader, S. M. & Laland, K. N. (Eds.) (2003) Animal Innovation. New York: Oxford University Press.Google Scholar
Russon, A. E. (1998). The nature and evolution of intelligence in orangutans (Pongo pygmaeus). Primates, 39(4), 485503.Google Scholar
Russon, A. E. (2004). Great Ape Cognitive Systems. In Russon, A. E. & Begun, D. R. (Eds.), The Evolution of Thought: Evolutionary Origins of Great Ape Intelligence (pp. 76100). Cambridge, UK: Cambridge University Press.Google Scholar
Russon, A. E., van Schaik, C. P., Kuncoro, P., Ferisa, A., Handayani, D. P., van Noordwijk, M. A. (2009). Innovation and Intelligence in Orangutans. In Witch, S. A., Utami, S. A., Mitra, S. T., & van Schaik, P. (Eds.), Orangutans: Geographic Variation in Behavioral Ecology and Conservation (pp. 279298). Oxford, UK: Oxford University Press.Google Scholar
Russon, A. E., Kuncoro, P., Ferisa, A., & Handayani, D. P. (2010). How orangutans (Pongo pygmaeus) innovate for water. Journal of Comparative Psychology, 124(1), 1428.Google Scholar
Russon, A. E., Kuncoro, P., & Ferisa, A. (2015). Tools for the Trees: Orangutan Arboreal Tool Use and Creativity. In Kaufman, A. B. & Kaufman, J. C. (Eds.), Animal Creativity and Innovation (pp. 419455). San Diego, CA: Academic Press/Elsevier.Google Scholar
Schuppli, C., Forss, S., Meulman, E., Utami Atmoko, S., van Noordwijk, M., & van Schaik, C. (2017). The effects of sociability on exploratory tendency and innovation repertoires in wild Sumatran and Bornean orangutans. Scientific Reports, 7, 15464. doi: 10.1038/s41598–017-15640-xGoogle Scholar
Shettleworth, S. (2010). Cognition, Evolution, and Behavior (2nd ed.). New York: Oxford University Press.Google Scholar
Shumaker, R. W., Walkup, K. R., & Beck, B. (2011). Animal Tool Behavior: The Use and Manufacture of Tools by Animals (Revised and updated edition). Baltimore, MD: John’s Hopkins University Press.Google Scholar
Singleton, I., Knott, C. D., Morrogh-Bernard, H.C., Wich, S. A., & van Schaik, C. P. (2009). Ranging Behavior of Orangutan Females and Social Organization. In Wich, S. A et al. (Eds.), Orangutans: Geographic Variation in Behavioral Ecology and Conservation (pp. 205213). New York: Oxford University Press.Google Scholar
Sol, D. (2015). The Evolution of Innovativeness: Exaptation or Specialized Adaptation? In Kaufman, A. B. & Kaufman, J. C. (Eds.), Animal Creativity and Innovation (pp. 163182). San Diego, CA: Academic Press/Elsevier.Google Scholar
Stokes, E. & Byrne, R. W. (2001). Cognitive abilities for behavioural flexibility in wild chimpanzees (Pan troglodytes): The effect of snare injury on complex manual food processing. Animal Cognition, 4, 1128.Google Scholar
Sugardjito, J. & van Hooff, J. A. R. A. M. (1986). Age-sex class differences in the positional behavior of the Sumatran orangutan (Pongo pygmaeus abelii) in the Gunung Leuser National Park, Indonesia. Folia Primatologica, 47, 1425.Google Scholar
Thorpe, S. K. S, Crompton, R. H., & Alexander, R. M. (2007). Orangutans use compliant branches to lower the energetic cost of locomotion. Biology Letters, 3, 253256. doi: 10.1098/rsbl.2007.0049Google Scholar
Tomasello, M., Kruger, A. C., & Ratner, H. H. (1993). Cultural learning. Behavioral and Brain Sciences, 16, 495511.Google Scholar
van Schaik, C. P., Deaner, R. O., & Merrill, M. Y. (1999). The conditions for tool use in primates: Implications for the evolution of material culture. Journal of Human Evolution, 36, 719741.Google Scholar
van Schaik, C. P. & Knott, C. D. (2001). Geographic variation in tool use on Neesia fruit in orangutans. American Journal of Physical Anthropology, 114, 331342.Google Scholar
van Schaik, C. P., Ancrenaz, M., Borgen, G., Galdikas, B., Knott, C. D., Singleton, I., Suzuki, A., Utami, S. S., & Merrill, M. (2003). Orangutan cultures and the evolution of material culture. Science, 299, 102105.Google Scholar
van Schaik, C. P., van Noordwijk, M. A., & Wich, S. A. (2006). Innovation in wild Bornean orangutans (Pongo pygmaeus wurmbii). Behaviour, 143, 839–76. doi: 10.1163/156853906778017944Google Scholar
van Schaik, C. P., Burkart, J., Damerius, L., Forss, S. I. F., Koops, K., van Noordwijk, M. A., & Schuppli, C. (2016). The reluctant innovator: Orangutans and phylogeny of creativity. Phil. Trans. Royal Society B, 371, 20150183. http://dx.doi.org/10.1098/rstb.2015.0183Google Scholar
Whiten, A., Goodall, J., McGrew, W. C., Nishida, T., Reynolds, V., Sugiyama, Y., Tutin, C. E. G., Wrangham, R. W. & Boesch, C. (1999). Cultures in chimpanzees. Nature, 399, 682685.Google Scholar
Wich, S. A., Utami-Atmoko, S. S., Mitra-Setia, T., & van Schaik, C. P. (Eds.) (2009). Orangutans: Geographic Variation in Behavioral Ecology and Conservation. New York: Oxford University Press.Google Scholar
Yamakoshi, G. (2004). Evolution of Complex Feeding Techniques in Primates: Is This the Origin of Great-Ape Intelligence? In Russon, A. E. & Begun, D. R. (Eds.), The Evolution of Thought: Evolutionary Origins of Great Ape Intelligence (pp. 140171). Cambridge, UK: Cambridge University Press.Google Scholar

References

Basile, B. M., Hampton, R. R., Suomi, S. J., & Murray, E. A. (2009) An assessment of memory awareness in tufted capuchin monkeys (Cebus apella). Animal Cognition 12, 169180.Google Scholar
Basile, B. M., Schroeder, G. R., Brown, E. K., Templer, V. L., & Hampton, R. R. (2015). Evaluation of seven hypotheses for metamemory performance in rhesus monkeys. Journal of Experimental Psychology: General, 144(1), 85.Google Scholar
Beran, M. J., Smith, J. D., Redford, J. S., & Washburn, D. A. (2006). Rhesus macaques (Macaca mulatta) monitor uncertainty during numerosity judgments. Journal of Experimental Psychology: Animal Behavior Processes, 32(2), 111.Google Scholar
Beran, M. J., Smith, J. D., Coutinho, M. V. C., Couchman, J. C., & Boomer, J. (2009). The psychological organization of ‘‘uncertainty’’ responses and ‘‘middle’’ responses: A dissociation in capuchin monkeys (Cebus apella). Journal of Experimental Psychology: Animal Behavior Processes, 35, 371381.Google Scholar
Beran, M. J. & Smith, J. D. (2011). Information seeking by rhesus monkeys (Macaca mulatta) and capuchin monkeys (Cebus apella). Cognition, 120, 90105.Google Scholar
Beran, M. J., Smith, J. D., & Perdue, B. M. (2013). Language-trained chimpanzees (Pan troglodytes) name what they have seen but look first at what they have not seen. Psychological Science, 24, 660666.Google Scholar
Beran, M. J., Perdue, B. M., & Smith, J. D. (2014). What are my chances? Closing the gap in uncertainty monitoring between rhesus monkeys (Macaca mulatta) and capuchin monkeys (Cebus apella). Journal of Experimental Psychology: Animal Learning & Cognition. doi: 10.1037/xan0000020Google Scholar
Beran, M. J., Perdue, B. M., Futch, S. E., Smith, J. D., Evans, T. A., & Parrish, A. E. (2015). Go when you know: Chimpanzees’ confidence movements reflect their responses in a computerized memory task. Cognition, 142, 236246.Google Scholar
Bermúdez, J. L. (2017). Can Nonlinguistic Animals Think about Thinking? In Andrews, K. & Bec, J.,. The Routledge Handbook of Philosophy of Animal Minds (pp. 119130). Abingdon, UK: Routledge.Google Scholar
Bohn, M., Allritz, M., Call, J., & Völter, C. J. (2017). Information seeking about tool properties in great apes. Scientific Reports, 7(1), 10923.Google Scholar
Bosc, M., Bioulac, B., Langbour, N., Nguyen, T. H., Goillandeau, M., Dehay, B., & Michelet, T. (2017). Checking behavior in rhesus monkeys is related to anxiety and frontal activity. Scientific Reports, 7, 45267.Google Scholar
Brown, E. K., Templer, V. L., & Hampton, R. R. (2017). An assessment of domain-general metacognitive responding in rhesus monkeys. Behavioural Processes, 135, 132144.Google Scholar
Brown, R. & McNeill, D. (1966). The “tip of the tongue” phenomenon. Journal of Verbal Learning and Verbal Behavior, 5(4), 325337.Google Scholar
Browne, D. (2004). Do dolphins know their own minds? Biology and Philosophy, 19, 633653.Google Scholar
Call, J. (2004). Inferences about the location of food in the great apes (Pan paniscus, Pan troglodytes, Gorilla gorilla, and Pongo pygmaeus)Journal of Comparative Psychology118(2), 232.Google Scholar
Call, J. (2010). Do apes know that they could be wrong? Animal Cognition, 13, 689700.Google Scholar
Call, J. (2012). Seeking Information in Non-Human Animals: Weaving a Metacognitive Web. In Beran, M. J., Brandl, J., Perner, J., & Proust, J. (Eds.), Foundations of Metacognition (pp. 6275). Oxford: Oxford University Press.Google Scholar
Call, J. & Carpenter, M. (2001). Do apes and children know what they have seen? Animal Cognition, 4, 20072220.Google Scholar
Carruthers, P. (2008). Meta-cognition in animals: A skeptical look. Mind Language 23, 5889.Google Scholar
Couchman, J. J., Coutinho, M. V., Beran, M. J., & Smith, J. D. (2010). Beyond stimulus cues and reinforcement signals: A new approach to animal metacognition. Journal of Comparative Psychology, 124(4), 356.Google Scholar
Davis, H. & Memmott, J. (1982). Counting behavior in animals: A critical evaluation. Psychological Bulletin, 92, 547571.Google Scholar
Davis, H. & Perusse, R. (1988). Numerical competence: From backwater to mainstream of comparative psychology. Behavioral Brain Science, 11, 602615.Google Scholar
Ferrigno, S., Kornell, N., & Cantlon, J. F. (2017). A metacognitive illusion in monkeys. Proceedings of the Royal Society B: Biological Sciences, 284(1862), 20171541.Google Scholar
Fujita, K. (2009). Metamemory in tufted capuchin monkeys (Cebus apella). Animal Cognition, 12, 575585.Google Scholar
Hampton, R. R. (2001). Rhesus monkeys know when they remember. Proceedings of the National Academy of Science, 98, 53595362.Google Scholar
Hampton, R. R. (2009). Multiple demonstrations of metacognition in nonhumans: Converging evidence or multiple mechanisms? Comparative Cognition & Behavior Reviews 4, 1728.Google Scholar
Hampton, R. R. & Hampstead, B. M. (2006). Spontaneous behavior of a rhesus monkey (Macaca mulatta) during memory tests suggests memory awareness. Behavioural Processes, 72(2), 184189.Google Scholar
Hampton, R. R., Zivin, A., & Murray, E. A. (2004). Rhesus monkeys (Macaca mulatta) discriminate between knowing and not knowing and collect information as needed before acting. Animal Cognition 7, 239246.Google Scholar
Hart, J. T. (1965). Memory and the feeling-of-knowing experience. Journal of Educational Psychology, 56(4), 208.Google Scholar
Kiani, R. & Shadlen, M. N. (2009). Representation of confidence associated with a decision by neurons in the parietal cortex. Science, 324, 759764.Google Scholar
Kornell, N, Son, L. K., & Terrace, H. S. (2007). Transfer of metacognitive skills and hint seeking in monkeys. Psychological Science, 18, 6471.Google Scholar
Malassis, R., Gheusi, G., & Fagot, J. (2015). Assessment of metacognitive monitoring and control in baboons (Papio papio). Animal Cognition, 18(6), 13471362.Google Scholar
Marsh, H. L. (2014). Metacognitive-like information seeking in lion-tailed macaques: A generalized search response after all? Animal Cognition, 17(6), 13131328.Google Scholar
Marsh, H. L. & MacDonald, S. E. (2012a) Information seeking by orangutans: A generalized search strategy? Animal Cognition, 15, 293304.Google Scholar
Marsh, H. L. & MacDonald, S. E. (2012b). Orangutans (Pongo abelii) ‘‘play the odds’’: Information-seeking strategies in relation to cost, risk, and benefit. Journal of Comparative Psychology, 126, 263278.Google Scholar
Marsh, H. L., Vining, A. Q., Levendoski, E. K., & Judge, P. G. (2015). Inference by exclusion in lion-macaques (Macaca silenus), a hamadryas baboon (Papio hamadryas), capuchins (Sapajus apella), and squirrel monkeys (Saimiri sciureus). Journal of Comparative Psychology, 129(3), 256.Google Scholar
Middlebrooks, P. G. & Sommer, M. A. (2011). Metacognition in monkeys during an oculomotor task. Journal of Experimental Psychology: Learning Memory, and Cognition, 37, 325337.Google Scholar
Morales, J. (2016). When behavior is not enough: Reading metacognition from the brain. Proceedings of the Rotman Institute of Philosophy Annual Conference “Rethinking the taxonomy of psychology workshop,” London, Canada.Google Scholar
Morgan, G., Kornell, N., Kornblum, T., & Terrace, H. S. (2014). Retrospective and prospective metacognitive judgments in rhesus macaques (Macaca mulatta). Animal Cognition, 17(2), 249257.Google Scholar
Mulcahy, N. J. (2016). Orangutans (Pongo abelii) seek information about tool functionality in a metacognition tubes task. Journal of Comparative Psychology, 130(4), 391.Google Scholar
Neldner, K., Collier-Baker, E., & Nielsen, M. (2015). Chimpanzees (Pan troglodytes) and human children (Homo sapiens) know when they are ignorant about the location of food. Animal Cognition, 18(3), 683699.Google Scholar
Paukner, A., Anderson, J. R., & Fujita, K. (2006) Redundant food searches by capuchin monkeys (Cebus apella): A failure of metacognition? Animal Cognition, 9, 110117.Google Scholar
Perdue, B. M., Evans, T. A., & Beran, M. J. (2018). Chimpanzees show some evidence of selectively acquiring information by using tools, making inferences, and evaluating possible outcomes. PloS One, 13(4), e0193229.Google Scholar
Perner, J. (2012). MiniMeta: In Search of Minimal Criteria for Metacognition. In Beran, M. J., Brandl, J., Perner, J., & Proust, J., (Eds.), Foundations of Metacognition (pp. 94116). Oxford: Oxford University Press.Google Scholar
Premack, D. & Woodruff, G. (1978). Does the chimpanzee have a theory of mind? Behavioral and Brain Sciences, 1(4), 515526.Google Scholar
Redford, J. S. (2010). Evidence of metacognitive control by humans and monkeys in a perceptual categorization task. Journal of Experimental Psychology: Learning, Memory, & Cognition, 36, 248254.Google Scholar
Rosati, A. G. & Santos, L. R. (2016). Spontaneous metacognition in rhesus monkeys. Psychological Science, 27(9), 11811191.Google Scholar
Sayers, K., Evans, T. A., Menzel, E., Smith, J. D., & Beran, M. J. (2015). The misbehaviour of a metacognitive monkey. Behaviour, 152(6), 727756.Google Scholar
Shields, W. E., Smith, J. D., & Washburn, D. A. (1997). Uncertain responses by humans and rhesus monkeys (Macaca mulatta) in a psychophysical same-different task. Journal of Experimental Psychology: General, 126, 147164.Google Scholar
Smith, J. D. (2009). The study of animal metacognition. Trends in Cognitive Science, 13, 389396.Google Scholar
Smith, J. D. (2010). Inaugurating the study of animal metacognition. International Journal of Comparative Psychology, 23(3), 401.Google Scholar
Smith, J. D., Schull, J., Strote, J., McGee, K., Egnor, R., & Erb, L. (1995). The uncertain response in the bottlenosed dolphin (Tursiops truncates). Journal of Experimental Psychology: General, 124, 391408.Google Scholar
Smith, J. D., Shields, W. E., Schull, J., & Washburn, D. A. (1997). The uncertain response in humans and animals. Cognition, 62, 7597.Google Scholar
Smith, J. D., Shields, W. E., Allendoerfer, K. R., & Washburn, D. A. (1998). Memory monitoring by animals and humans. Journal of Experimental Psychology: General, 127, 227250.Google Scholar
Smith, J. D., Beran, M. J., Couchman, J. J., & Coutinho, M. V. (2008). The comparative study of metacognition: Sharper paradigms, safer inferences. Psychonomic Bulletin & Review, 15(4), 679691.Google Scholar
Smith, J. D., Redford, J. S., Beran, M. J., & Washburn, D. A. (2010). Rhesus monkeys (Macaca mulatta) adaptively monitor uncertainty while multi-tasking. Animal cognition, 13(1), 93.Google Scholar
Smith, T. R., Smith, J. D., & Beran, M. J. (2018). Not knowing what one knows: A meaningful failure of metacognition in capuchin monkeys. Animal Behavior and Cognition, 5(1), 5567.Google Scholar
Staddon, J. E. R., Jozefowiez, J., & Cerutti, D. T. (2007). Metacognition: A problem, not a process. PsyCrit, 13(1), 15.Google Scholar
Suda, C. & Call, J. (2006). What does an intermediate success rate mean? An analysis of a Piagetian liquid conservation task in the great apesCognition99(1), 5371.Google Scholar
Suda-King, C. (2008). Do orangutans (Pongo pygmaeus) know when they do not remember? Animal Cognition, 11, 2142.Google Scholar
Suda-King, C., Bania, A. E., Stromberg, E. E., & Subiaul, F. (2013). Gorillas’ use of the escape response in object choice memory tests. Animal Cognition, 16(1), 6584.Google Scholar
Takagi, S. & Fujita, K. (2018). Do capuchin monkeys (Sapajus apella) know the contents of memory traces? A study of metamemory for compound stimuli. Journal of Comparative Psychology, 132(1), 88.Google Scholar
Templer, V. L. & Hampton, R. R. (2012). Rhesus monkeys (Macaca mulatta) show robust evidence for memory awareness across multiple generalization tests. Animal Cognition, 15(3), 409419.Google Scholar
Tu, H. W., Pani, A. A., & Hampton, R. R. (2015). Rhesus monkeys (Macaca mulatta) adaptively adjust information seeking in response to information accumulated. Journal of Comparative Psychology, 129(4), 347.Google Scholar
Vining, A. Q. & Marsh, H. L. (2015). Information seeking in capuchins (Cebus apella): A rudimentary form of metacognition? Animal cognition, 18(3), 667681.Google Scholar
Washburn, D. A., Smith, J. D., & Shields, W. E. (2006). Rhesus monkeys (Macaca mulatta) immediately generalize the uncertain response. Journal of Experimental Psychology: Animal Behavior Processes, 32(2), 185.Google Scholar

References

Addessi, E., Paglieri, F., & Focaroli, V. (2011). The ecological rationality of delay tolerance: Insights from capuchin monkeysCognition119(1), 142147.Google Scholar
Addessi, E., Paglieri, F., Beran, M. J., Evans, T. A., Macchitella, L., De Petrillo, F., & Focaroli, V. (2013). Delay choice versus delay maintenance: Different measures of delayed gratification in capuchin monkeys (Cebus apella). J. Comp. Psychol., 127(4), 392.Google Scholar
Ainslie, G. (2001). Breakdown of Will. Cambridge, UK: Cambridge University Press.Google Scholar
Ainslie, G. & Herrnstein, R. J. (1981). Preference reversal and delayed reinforcement. Anim. Learn. Behav., 9(4), 476482.Google Scholar
Anderson, J. R., Kuroshima, H., & Fujita, K. (2010). Delay of gratification in capuchin monkeys (Cebus apella) and squirrel monkeys (Saimiri sciureus). J. Comp. Psychol., 124(2), 205.Google Scholar
Auersperg, A. M. I., Laumer, I. B., & Bugnyar, T. (2013). Goffin cockatoos wait for qualitative and quantitative gains but prefer “better” to “more.” Biol. Letts, 9(3), 20121092.Google Scholar
Barberis, N. C. (2013). Thirty years of prospect theory in economics: A review and assessment. J. Econ. Perspect., 27(1), 173196.Google Scholar
Baron, J. (2000). Thinking and Deciding. Cambridge, UK: Cambridge University Press.Google Scholar
Bateson, M. (2002). Recent advances in our understanding of risk-sensitive foraging preferences. Proc. Nutrit. Soc., 61(4), 509516.Google Scholar
Beran, M. J. (2002). Maintenance of self-imposed delay of gratification by four chimpanzees (Pan troglodytes) and an orangutan (Pongo pygmaeus). J. Gen. Psychol., 129(1), 4966.Google Scholar
Beran, M. J., Savage‐Rumbaugh, E. S., Pate, J. L., & Rumbaugh, D. M. (1999). Delay of gratification in chimpanzees (Pan troglodytes). Dev. Psychobiol., 34(2), 119127Google Scholar
Beran, M. J. & Evans, T. A. (2006). Maintenance of delay of gratification by four chimpanzees (Pan troglodytes): The effects of delayed reward visibility, experimenter presence, and extended delay intervals. Behav. Proc., 73(3), 315324.Google Scholar
Bernoulli, D. (1954). [1738] Exposition of a new theory on the measurement of risk. Econometrica, 22, 2336.Google Scholar
Brewer, S. M. & McGrew, W. C. (1990). Chimpanzee use of a tool-set to get honey. Folia Primatol., 54(1-2), 100104.Google Scholar
Brosnan, S. F., Jones, O. D., Lambeth, S. P., Mareno, M. C., Richardson, A. S., & Schapiro, S. J. (2007). Endowment effects in chimpanzees. Curr. Biol., 17(19), 17041707.Google Scholar
Brosnan, S. F., Jones, O. D., Gardner, M., Lambeth, S. P., & Schapiro, S. J. (2012). Evolution and the expression of biases: Situational value changes the endowment effect in chimpanzees. Evol. Hum. Behav., 33(4), 378386.Google Scholar
Camerer, C., Loewenstein, G., & Rabin, M. (2004). Advances in Behavioral Economics. New York: Russell Sage Foundation.Google Scholar
Caraco, T., Martindale, S., & Whittam, T. S. (1980). An empirical demonstration of risk-sensitive foraging preferences. Anim. Behav., 28, 820830.Google Scholar
Carruthers, P. (2002). The cognitive functions of language. Behav. Brain Sci., 25(6), 657674.Google Scholar
Charnov, E. L. (1976). Optimal foraging, the marginal value theorem. Theor. Popul. Biol., 9(2), 129136.Google Scholar
Chen, M. K., Lakshminarayanan, V., & Santos, L. R. (2006). How basic are behavioral biases? Evidence from capuchin monkey trading behavior. J. Polit. Econ., 114(3), 517537.Google Scholar
Clutton-Brock, T. H. & Harvey, P. H. (1979). Comparison and adaptation. Proc. R. Soc. Lond. B., 205(1161), 547565.Google Scholar
Csibra, G. & Gergely, G. (2009). Natural pedagogy. Trends Cogn. Sci., 13(4), 148153.Google Scholar
Darwin, C. (1854). Journal of Researches into the Natural History and Geology of the Countries Visited During the Voyage of H.M.S. Beagle round the World, under the Command of Capt. Fitz Roy, R.N. (2nd ed.). London: John Murray.Google Scholar
De Petrillo, F., Ventricelli, M., Ponsi, G., & Addessi, E. (2015). Do tufted capuchin monkeys play the odds? Flexible risk preferences in Sapajus sppAnim. Cogn., 18(1), 119130.Google Scholar
De Petrillo, F. & Rosati, A. G. (2019). Ecological rationality: Convergent decision-making in apes and capuchins. Behav. Proc., 164, 201213.Google Scholar
Drayton, L. A., Brosnan, S. F., Carrigan, J., & Stoinski, T. S. (2013). Endowment effects in gorillas (Gorilla gorilla). J. Comp. Psychol., 127(4), 365.Google Scholar
Dufour, V., Wascher, C. A., Braun, A., Miller, R., & Bugnyar, T. (2012). Corvids can decide if a future exchange is worth waiting for. Biol. Letters, 8(2), 201204.Google Scholar
Eckert, J., Rakoczy, H., & Call, J. (2017). Are great apes able to reason from multi‐item samples to populations of food items? Am. J. Primatol., 79(10), e22693.Google Scholar
Eckert, J., Call, J., Hermes, J., Herrmann, E., & Rakoczy, H. (2018a). Intuitive statistical inferences in chimpanzees and humans follow Weber’s law. Cognition, 180, 99107.Google Scholar
Eckert, J., Rakoczy, H., Call, J., Herrmann, E., & Hanus, D. (2018b). Chimpanzees consider humans’ psychological states when drawing statistical inferences. Curr. Biol., 28(12), 19591963.Google Scholar
Evans, T. A. & Beran, M. J. (2007). Chimpanzees use self-distraction to cope with impulsivity. Biol. Letts, 3(6), 599602.Google Scholar
Evans, T. A., Beran, M. J., Paglieri, F., & Addessi, E. (2012). Delaying gratification for food and tokens in capuchin monkeys (Cebus apella) and chimpanzees (Pan troglodytes): When quantity is salient, symbolic stimuli do not improve performance. Anim. Cogn., 15(4), 539548.Google Scholar
Fawcett, T. W., McNamara, J. M., & Houston, A. I. (2012). When is it adaptive to be patient? A general framework for evaluating delayed rewards. Behav. Processes, 89(2), 128136.Google Scholar
Flemming, T. M., Jones, O. D., & Mayo, L. (2012). The endowment effect in orangutans. J. Comp. Psychol, 25, 4.Google Scholar
Fragaszy, D. M., Visalberghi, E., & Fedigan, L. M. (2004). The Complete Capuchin: The Biology of the Genus Cebus. Cambridge, UK: Cambridge University Press.Google Scholar
Franciosi, R., Kujal, P., Michelitsch, R., Smith, V., & Deng, G. (1996). Experimental tests of the endowment effect. J. Econ. Behav. Organ., 30(2), 213226.Google Scholar
Frederick, S., Loewenstein, G., & O’Donoghue, T. (2002). Time discounting and time preference: A critical review. J. Econ. Lit., 40, 350401.Google Scholar
Furuichi, T., Sanz, C., Koops, K., Sakamaki, T., Ryu, H., Tokuyama, N., & Morgan, D. (2015). Why do wild bonobos not use tools like chimpanzees do? Behaviour152(3-4), 425460.Google Scholar
Garber, P. A. (1993). Feeding Ecology and Behaviour of the Genus Saguinus. In Rylands, A. B. (Ed.), Marmosets and Tamarins: Systematics, Behaviour, and Ecology (pp. 273295). Oxford: Oxford University Press.Google Scholar
Gilby, I. C. & Wrangham, R. W. (2007). Risk-prone hunting by chimpanzees (Pan troglodytes schweinfurthii) increases during periods of high diet quality. Behav. Ecol. Sociobio., 61(11), 17711779.Google Scholar
Gilby, I. C., Machanda, Z. P., O’Malley, R. C., Murray, C. M., Lonsdorf, E. V., Walker, K., … & Pusey, A. E. (2017). Predation by female chimpanzees: Toward an understanding of sex differences in meat acquisition in the last common ancestor of Pan and Homo. J. Hum. Evol., 110, 8294.Google Scholar
Green, L., Fristoe, N., & Myerson, J. (1994). Temporal discounting and preference reversals in choice between delayed outcomes. Psychon. B. Rev., 1(3), 383389.Google Scholar
Green, L., Myerson, J., Holt, D. D., Slevin, J. R., & Estle, S. J. (2004). Discounting of delayed food rewards in pigeons and rats: Is there a magnitude effect? J. Exp. Anal. Behav., 81(1), 39-50.Google Scholar
Green, L., Myerson, J., & Calvert, A. L. (2010). Pigeons’ discounting of probabilistic and delayed reinforcers. J. Exp. Anal. Behav., 94(2), 113123.Google Scholar
Haun, D. B., Nawroth, C., & Call, J. (2011). Great apes’ risk-taking strategies in a decision making task. PLoS One, 6, e28801e28801.Google Scholar
Hayden, B. Y. & Platt, M. L. (2007). Temporal discounting predicts risk sensitivity in rhesus macaques. Curr. Biol., 17, 4953.Google Scholar
Heilbronner, S. R., Rosati, A. G., Stevens, J. R., Hare, B., & Hauser, M. D. (2008). A fruit in the hand or two in the bush? Divergent risk preferences in chimpanzees and bonobos. Biol. Letters, 4, 246249.Google Scholar
Hillemann, F., Bugnyar, T., Kotrschal, K., & Wascher, C. A. (2014). Waiting for better, not for more: Corvids respond to quality in two delay maintenance tasks. Anim. Behav., 90, 110.Google Scholar
Houston, A. I. & McNamara, J. M. (1999). Models of Adaptive Behaviour: An Approach Based on State. Cambridge, UK: Cambridge University Press.Google Scholar
Houston, A. I., McNamara, J. M., & Steer, M. D. (2007). Do we expect natural selection to produce rational behaviour? Phil. Trans. R. Soc. Lond. B, 362(1485), 15311543.Google Scholar
Johnson, E. J., Hershey, J., Meszaros, J., & Kunreuther, H. (1993). Framing, probability distortions, and insurance decisions. J. Risk Uncert., 7(1), 3551.Google Scholar
Kacelnik, A. (2003). The Evolution of Patience. In Loewenstein, G, Read, G, & Baumeister, R. (Eds.), Time and Decision: Economic and Psychological Perspectives on Intertemporal Choice (pp. 115138). New York: Russell Sage Foundation.Google Scholar
Kacelnik, A. (2006). Meanings of Rationality. In Hurley, S. & Nudds, M (Eds.), Rational Animals? (87106), Oxford, UK: Oxford University Press.Google Scholar
Kacelnik, A. & Bateson, M. (1996). Risky theories: The effects of variance on foraging decisions. Am. Zool., 36, 402434.Google Scholar
Kacelnik, A. & Marsh, B. (2002). Cost can increase preference in starlings. Anim. Behav., 63(2), 245250.Google Scholar
Kacelnik, A. & El Mouden, C. (2013). Triumphs and trials of the risk paradigm. Anim. Behav., 86, 11171129.Google Scholar
Kahneman, D. & Tversky, A. (1979). Prospect theory: An analysis of decision under risk. Econometrica, 47, 263291.Google Scholar
Kahneman, D., Knetsch, J. L., & Thaler, R. H. (1990). Experimental tests of the endowment effect and the Coase theorem. J. Polit. Econ., 98(6), 13251348.Google Scholar
Kahneman, D., Knetsch, J. L., & Thaler, R. H. (1991). Anomalies: The endowment effect, loss aversion, and status quo bias. J. Econ. Perspect., 5(1), 193206.Google Scholar
Kahneman, D. & Tversky, A. (2000). Choices, Values, and Frames. Cambridge University Press, New York: Russell Sage Foundation.Google Scholar
Kahneman, D. & Thaler, R. H. (2006). Anomalies: Utility maximization and experienced utility. Journal of Economic Perspectives, 20(1), 221234.Google Scholar
Kanngiesser, P., Santos, L. R., Hood, B. M., & Call, J. (2011). The limits of endowment effects in great apes (Pan paniscus, Pan troglodytes, Gorilla gorilla, Pongo pygmaeus). J. Comp. Psychol., 125(4), 436.Google Scholar
Krebs, J. R. & Davies, N. B. (1978). Behavioural Ecology: An Evolutionary ApproachOxford, UK: Blackwell Sci.Google Scholar
Krupenye, C., Rosati, A. G., & Hare, B. (2015). Bonobos and chimpanzees exhibit human-like framing effects. Biol. Letters, 11. 20140527.Google Scholar
Lakshminaryanan, V. R., Chen, M. K., & Santos, L. R. (2008). Endowment effect in capuchin monkeys. Phil. Trans. R. Soc. Lond. B, 363(1511), 38373844.Google Scholar
Lakshminarayanan, V. R., Chen, M. K., & Santos, L. R. (2011). The evolution of decision-making under risk: Framing effects in monkey risk preferences. J. Exp. Soc. Psychol., 47(3), 689693.Google Scholar
Leonardi, R. J., Vick, S. J., & Dufour, V. (2012). Waiting for more: The performance of domestic dogs (Canis familiaris) on exchange tasks. Anim. Cogn., 15(1), 107120.Google Scholar
Lerner, J. S., Small, D. A., & Loewenstein, G. (2004). Heart strings and purse strings: Carryover effects of emotions on economic decisions. Psychol. Sci., 15(5), 337341.Google Scholar
MacArthur, R. H. & Pianka, E. R. (1966). On optimal use of a patchy environment. Am. Nat., 100(916), 603609.Google Scholar
MacLean, E. L., Mandalaywala, T. M., & Brannon, E. M. (2012). Variance-sensitive choice in lemurs: Constancy trumps quantity. Anim. Cogn., 15(1), 1525.Google Scholar
Malenky, R. K. & Wrangham, R. W. (1994). A quantitative comparison of terrestrial herbaceous food consumption by Pan paniscus in the Lomako Forest, Zaire, and Pan troglodytes in the Kibale Forest, Uganda. Am. J. Primatol., 32(1), 112.Google Scholar
Marsh, B. & Kacelnik, A. (2002). Framing effects and risky decisions in starlings. PNAS, 99(5), 33523355.Google Scholar
Marsh, B., Schuck-Paim, C., & Kacelnik, A. (2004). Energetic state during learning affects foraging choices in starlings. Behav. Ecol., 15(3), 396399.Google Scholar
Marshall-Pescini, S., Besserdich, I., Kratz, C., & Range, F. (2016). Exploring differences in dogs’ and wolves’ preference for risk in a foraging taskFront. Psychol., 7, 1241.Google Scholar
Mayr, E. (1982). The Growth of Biological Thought: Diversity, Evolution, and Inheritance. Cambridge, MA: Harvard University Press.Google Scholar
Mazur, J. E. (1987). An Adjusting Procedure for Studying Delayed Reinforcement. In Commons, M. L., Mazur, J. E., Nevin, J. A., Rachlin, H. (Eds.), Quantitative Analyses of Behavior: The Effect of Delay and of Intervening Events on Reinforcement Value (pp. 5573). Hillsdale, NJ: ErlbaumGoogle Scholar
McCoy, A. N. & Platt, M. L. (2005). Risk-sensitive neurons in macaque posterior cingulate cortex. Nat. Neurosci., 8, 12201227.Google Scholar
Modeling Animal Decisions Group, Fawcett, T. W., Fallenstein, B., Higginson, A. D., Houston, A. I, et al. (2014). The evolution of decision rules in complex environments. Trends Cogn. Sci., 18, 153161.Google Scholar
Morewedge, C. K., Shu, L. L., Gilbert, D. T., & Wilson, T. D. (2009). Bad riddance or good rubbish? Ownership and not loss aversion causes the endowment effect. J. Exp. Soc. Psychol., 45(4), 947951.Google Scholar
Morewedge, C. K. & Giblin, C. E. (2015). Explanations of the endowment effect: An integrative review. Trends Cogn. Sci., 19(6), 339348.Google Scholar
Mühlhoff, N., Stevens, J. R., & Reader, S. M. (2011). Spatial discounting of food and social rewards in guppies (Poecilia reticulata). Front. Psychol., 2, 68.Google Scholar
Parrish, A. E., Perdue, B. M., Stromberg, E. E., Bania, A. E., Evans, T. A., & Beran, M. J. (2014). Delay of gratification by orangutans (Pongo pygmaeus) in the accumulation task. J. Comp. Psychol., 128(2), 209.Google Scholar
Pelé, M., Dufour, V., Micheletta, J., & Thierry, B. (2010). Long-tailed macaques display unexpected waiting abilities in exchange tasks. Anim. Cogn., 13(2), 263271.Google Scholar
Penn, D. C., Holyoak, K. J., & Povinelli, D. J. (2008). Darwin’s mistake: Explaining the discontinuity between human and nonhuman minds. Behav. Brain Sci., 31(2), 109130.Google Scholar
Perry, S. & Rose, L. (1994). Begging and transfer of coati meat by white-faced capuchin monkeys, Cebus capucinusPrimates35(4), 409415.Google Scholar
Platt, M. L., Brannon, E. M., Briese, T. L., & French, J. A. (1996). Differences in feeding ecology predict differences in performance between golden lion tamarins (Leontopithecus rosalia) and Wied’s marmosets (Callithrix kuhli) on spatial and visual memory tasks. Anim. Learn. Behav., 24(4), 384393.Google Scholar
Platt, M. L. & Huettel, S. A. (2008). Risky business: The neuroeconomics of decision making under uncertainty. Nat. Neurosci., 11(4), 398.Google Scholar
Pompilio, L. & Kacelnik, A. (2005). State-dependent learning and suboptimal choice: When starlings prefer long over short delays to food. Anim. Behav., 70(3), 571578.Google Scholar
Pompilio, L., Kacelnik, A., & Behmer, S. T. (2006). State-dependent learned valuation drives choice in an invertebrate. Science, 311(5767), 16131615.Google Scholar
Premack, D. (2007). Human and animal cognition: Continuity and discontinuity. PNAS, 104(35), 1386113867.Google Scholar
Proctor, D., Williamson, R. A., Latzman, R. D., de Waal, F. B., & Brosnan, S. F. (2014). Gambling primates: Reactions to a modified Iowa Gambling Task in humans, chimpanzees and capuchin monkeys. Anim. Cogn., 17(4), 983995.Google Scholar
Rakoczy, H., Clüver, A., Saucke, L., Stoffregen, N., Gräbener, A., Migura, J., & Call, J. (2014). Apes are intuitive statisticiansCognition131(1), 6068.Google Scholar
Richards, J. B., Mitchell, S. H., De Wit, H., & Seiden, L. S. (1997). Determination of discount functions in rats with an adjusting‐amount procedure. J. Exp. Anal. Behav., 67(3), 353366.Google Scholar
Rosati, A. G. (2017a). The Evolution of Primate Executive Function: From Response Control to Strategic Decision-Making. In Kaas, J. & Krubitzer, L. (Eds.), Evolution of Nervous Systems Vol. 3 (2nd ed.) (pp. 423437). Amsterdam: Elsevier.Google Scholar
Rosati, A.G. (2017b). Decisions under Uncertainty: Preferences, Biases, and Choice. In Call, J. (Ed.), APA Handbook of Comparative Psychology, Vol. 2 (pp. 329357). Washington, DC: American Psychological Association.Google Scholar
Rosati, A. G. (2017c). Foraging cognition: Reviving the ecological intelligence hypothesis. Trends Cogn. Sci., 21(9), 691702.Google Scholar
Rosati, A. G. (2017d). Ecological Variation in Cognition: Insights from Bonobos and Chimpanzees. In Hare, B. & Yamamoto, S. (Eds.), Bonobos: Unique in Mind, Brain and Behavior (pp. 157170). Oxford: Oxford University Press.Google Scholar
Rosati, A. G., Stevens, J. R., & Hauser, M. D. (2006). The effect of handling time on temporal discounting in two New World primatesAnim. Behav., 71(6), 13791387.Google Scholar
Rosati, A. G., Stevensen, J. R., Hare, B. & Hauser, M. D. (2007). The evolutionary origins of human patience: Temporal preferences in chimpanzees, bonobos, and human adults. Curr. Biol., 17, 16631668.Google Scholar
Rosati, A. G. & Stevens, J. R. (2009). The Adaptive Nature of Context-Dependent Choice. In Watanabe, S., Young, A., Huber, L., Blaisdell, A. & Yamazaki, Y. (Eds.), Rational Animal, Irrational Human (pp. 101117). Tokyo: Keio University Press.Google Scholar
Rosati, A. G. & Hare, B. (2011). Chimpanzees and bonobos distinguish between risk and ambiguity. Biol. Lett., 7, 1518.Google Scholar
Rosati, A. G. & Hare, B. (2012). Decision making across social contexts: Competition increases preferences for risk in chimpanzees and bonobos. Anim. Behav., 84, 869879.Google Scholar
Rosati, A. G. & Hare, B. (2013). Chimpanzees and bonobos exhibit emotional responses to decision outcomes. PLoS One, 8, e63058.Google Scholar
Rosati, A. G. & Hare, B. (2016). Reward type modulates human risk preferences. Evol. Hum. Behav., 37, 159168.Google Scholar
Rossano, F., Rakoczy, H., & Tomasello, M. (2011). Young children’s understanding of violations of property rights. Cognition, 121(2), 219227.Google Scholar
Rylands, A. B. & de Faria, D. S. (1993). Habitats, Feeding Ecology and Range Size in the Genus Callithrix. In Rylands, A. B. (Ed.), Marmosets and Tamarins: Systematics, Behaviour, and Ecology (pp. 262272). Oxford: Oxford University Press.Google Scholar
Santos, L. R. & Rosati, A. G. (2015). The evolutionary roots of human decision making. Ann. Rev. Psychol., 66, 321347.Google Scholar
Schuck-Paim, C., Pompilio, L., & Kacelnik, A. (2004). State-dependent decisions cause apparent violations of rationality in animal choice. PLoS Biol, 2(12), e402.Google Scholar
Spagnoletti, N., Visalberghi, E., Verderane, M. P., Ottoni, E., Izar, P., & Fragaszy, D. (2012). Stone tool use in wild bearded capuchin monkeys, Cebus libidinosus. Is it a strategy to overcome food scarcity? Anim. Behav., 83, 12851294.Google Scholar
Spelke, E. S. (2002). Developing knowledge of space: Core systems and new combinations| 5. Lang. Bra., 6, 239.Google Scholar
Stanford, C. B. (1999). The Hunting Apes: Meat Eating and the Origins of Human Behavior. Princeton, NJ: Princeton University Press.Google Scholar
Stanford, C. B. & Wrangham, R. W. (1998). Chimpanzee and Red Colobus: The Ecology of Predator and Prey. Cambridge, MA: Harvard University Press.Google Scholar
Stephens, D. W. (1981). The logic of risk-sensitive foraging preferences. Anim. Behav., 29, 628629.Google Scholar
Stephens, D. W. (2002). Discrimination, discounting and impulsivity: A role for an informational constraint. Phil. Trans. R. Soc. Lond. B, 357(1427), 15271537.Google Scholar
Stephens, D. W. & Krebs, J. R. (1986). Foraging Theory, Vol. 1. Princeton, NJ: Princeton University Press.Google Scholar
Stephens, D. W., Kerr, B., & Fernández-Juricic, E. (2004). Impulsiveness without discounting: The ecological rationality hypothesisProc. R. Soc. Lond. B., 271(1556), 24592465.Google Scholar
Stevens, J. R. (2008). The Evolutionary Biology of Decision Making. In Platt, M. L & Ghazanfar, A. A (Eds.), Primate Neuroethology (pp. 96116). Oxford, UK: Oxford University Press.Google Scholar
Stevens, J. R. (2010). Rational Decision Making in Primates: The Bounded and the Ecological. In Platt, M. L. & Ghazanfar, A. A. (Eds.), Primate Neuroethology (pp. 96116). Oxford, UK: Oxford University Press.Google Scholar
Stevens, J. R. (2014). Evolutionary pressures on primate intertemporal choiceProc. R. Soc. Lond. B, 281(1786), 20140499.Google Scholar
Stevens, J. R., Hallinan, E. V., & Hauser, M. D. (2005a). The ecology and evolution of patience in two New World monkeysBiol. Letts1(2), 223226.Google Scholar
Stevens, J. R., Rosati, A. G., Ross, K. R., & Hauser, M. D. (2005b). Will travel for food: Spatial discounting in two new world monkeysCurr. Biol., 15(20), 18551860.Google Scholar
Stevens, J. R. & Stephens, D. W. (2008). PatienceCurr. Biol., 18(1), R11R12.Google Scholar
Stevens, J. R. & Stephens, D. W. (2010). The Adaptive Nature of Impulsivity. In Medden, G. J. & Bickel, W. K. (Eds.), Impulsivity: The Behavioral and Neurological Science of Discounting (pp. 361387). Washington, DC: American Psychological Association.Google Scholar
Stevens, J. R. & Mühlhoff, N. (2012). Intertemporal choice in lemurs. Behav. Proc., 89(2), 121127.Google Scholar
Surbeck, M. & Hohmann, G. (2008). Primate hunting by bonobos at LuiKotale, Salonga National ParkCurr. Biol., 18(19), R906R907.Google Scholar
Tecwyn, E. C., Denison, S., Messer, E. J., & Buchsbaum, D. (2017). Intuitive probabilistic inference in capuchin monkeysAnim. Cogn., 20(2), 243256.Google Scholar
Thaler, R. (1980). Toward a positive theory of consumer choice. J. Econ. Behav. Organ., 1(1), 3960.Google Scholar
Thaler, R. (1981). Some empirical evidence on dynamic inconsistency. Econ. Lett., 8(3), 201207.Google Scholar
Thom, J. M. & Clayton, N. S. (2014). No evidence of temporal preferences in caching by Western scrub-jays (Aphelocoma californica). Behav. Proc., 103, 173179.Google Scholar
Tobin, H. & Logue, A. W. (1994). Self-control across species. J. Comp. Psychol., 108, 126133.Google Scholar
Tobin, H., Logue, A. W., Chelonis, J. J., Ackerman, K. T., & May, J. G. (1996). Self-control in the monkey Macaca fascicularis. Anim. Learn. Behav., 24(2), 168174.Google Scholar
Tversky, A. & Kahneman, D. (1992). Advances in prospect theory: Cumulative representation of uncertainty. J. Risk Uncert., 5, 297323.Google Scholar
Vick, S. J., Bovet, D., & Anderson, J. R. (2010). How do African Grey parrots (Psittacus erithacus) perform on a delay of gratification task? Anim. Cogn., 13(2), 351358.Google Scholar
Völter, C. J. & Call, J. (2017). Causal and Inferential Reasoning in Animals. In Call, J., Burkhardt, G. M., Pepperberg, I. M, Snowdon, C. T., & Zentall, T (Eds.), APA Handbook of Comparative Psychology (pp. 643671). Washington, DC: American Psychological Association.Google Scholar
von Neumann, J. & Morgenstern, O. (1947). Theory of Games and Economic Behavior (2nd ed.). Princeton, NJ: Princeton University Press.Google Scholar
Weber, E. U. & Johnson, E. J. (2008). Decisions Under Uncertainty: Psychological, Economic, and Euroeconomic Explanations of Risk Preference. In Glimcher, P., Camerer, C., Fehr, E., & Poldrack, R. (Eds.), Neuroeconomics: Decision Making and the Brain (pp. 127144). New YorkElsevier.Google Scholar
Weiner, J. (2014). The Beak of the Finch: A Story of Evolution in Our Time. New York: Vintage.Google Scholar
Wrangham, R. W., Conklin, N. L., Chapman, C. A., & Hunt, K. D. (1991). The significance of fibrous foods for Kibale Forest chimpanzees. Phil. Trans. R. Soc. Lond. B., 334(1270), 171178.Google Scholar
Wrangham, R. W., Conklin, N. L., Etot, G., Obua, J., Hunt, K. D., Hauser, M. D., & Clark, A. P. (1993). The value of figs to chimpanzees. Int. J. Primatol., 14(2), 243256.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
×