Skip to main content Accessibility help
×
Hostname: page-component-7bb8b95d7b-fmk2r Total loading time: 0 Render date: 2024-09-05T11:22:52.765Z Has data issue: false hasContentIssue false

References

Published online by Cambridge University Press:  08 August 2009

Jay Schulkin
Affiliation:
Georgetown University, Washington DC
Get access
Type
Chapter
Information
Cognitive Adaptation
A Pragmatist Perspective
, pp. 155 - 190
Publisher: Cambridge University Press
Print publication year: 2008

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

Abrous, D. N., Koehl, M., & Moal, M. (2005). Adult neurogenesis: From precursors to network and physiology. Physiological Reviews, 85, 523–569.CrossRefGoogle ScholarPubMed
Adolphs, R. (1999). Social cognition and the human brain. Trends in Cognitive Sciences, 3, 469–479.CrossRefGoogle ScholarPubMed
Adolphs, R., Denburg, N. L., & Tranel, D. (2001). The amygdala's role in long-term declarative memory for gist and detail. Behavioral Neuroscience, 112, 983–992.CrossRefGoogle Scholar
Adolphs, R., & Spezio, M. (2006). Role of the amygdala in processing visual social stimuli. Progress in Brain Research, 156, 363–378.CrossRefGoogle ScholarPubMed
Aggleton, J. (1992/2000). The amygdala. Oxford: Oxford University Press.Google ScholarPubMed
Aiello, L. C., & Dunbar, R. I. B. (1993). Neocortex size, group size and the evolution of language. Current Anthropology, 34, 184–192.CrossRefGoogle Scholar
Allison, T. (2001). Neuroscience and morality. Neuroscientist, 7, 360–364.CrossRefGoogle ScholarPubMed
Alloway, T. P., Corley, M., & Ramscar, M. (2006). Seeing ahead: Experience and language in spatial perspective. Memory and Cognition, 34, 38–86.CrossRefGoogle ScholarPubMed
Altman, J. (1966). Autoradiographic and histological studies of postnatal neurogenesis. Journal of Comparative Neurology, 124, 431–474.CrossRefGoogle Scholar
Amaral, D. G., Bauman, M. D., & Schumann, C. M. (2003). The amygdala and autism: Implication from non-human primate studies. Genes, Brain and Behavior, 2, 295–302.CrossRefGoogle Scholar
Anderson, D. (1997). The radical enlightenment of Benjamin Franklin. Baltimore: Johns Hopkins University Press.Google Scholar
Anderson, J. R. (1990). The adaptive character of thought. Hillsdale, NJ: Lawrence Erlbaum Associates.Google Scholar
Aristotle, . (1962). The Nicomachean ethics. New York: Macmillan.Google Scholar
Aristotle, . (1968). De anima. Oxford: Oxford University Press.Google Scholar
Ashwin, C., Baron-Cohen, S., Wheelwright, S., O'Riordan, M., & Bullmore, E. T. (2007). Differential activation of the amygdala and the social brain during fearful face-processing in Asperger syndrome. Neuropsychologia, 45, 2–14.CrossRefGoogle ScholarPubMed
Atran, S. (1990/1996). Cognitive foundations of natural history. New York: Cambridge University Press.Google Scholar
Atran, S. (2002). In gods we trust. Oxford: Oxford University Press.Google Scholar
Atran, S., & Medin, D. (2008). The native mind and the cultural construction of nature. Boston, MA: MIT Press.Google Scholar
Atran, A., Medin, D. L., & Ross, N. O. (2005). The cultural mind. Psychological Review, 112, 744–766.CrossRefGoogle ScholarPubMed
Atran, S., Medin, D., Ross, N., Lynch, E., Coley, J., Ucan Ek, E., & Vapnarsky, V. (1999). Folkecology and commons management in the Maya lowlands. Proceedings of the National Academy of Sciences, 96, 7598–7603.CrossRefGoogle ScholarPubMed
Atran, S., & Norenzayan, A. (2004). Religion's evolutionary landscape. Behavioral Brain Sciences, 27, 713–770.CrossRefGoogle ScholarPubMed
Augustine, . (1949). The confessions (E. B. Pusey, Trans.). New York: Random House.Google Scholar
Austin, J. H. (2006). Zen-brain reflections. Boston, MA: MIT Press.Google Scholar
Bachofen, J. J. (1926/1967). Myth, religion and Mother Right: Selected writings of J. J. Bachofen (R. Manheim, Trans.). Princeton, NJ: Princeton University Press.Google Scholar
Bacon, F. (1605/1974). The advancement of learning. Oxford: Oxford University Press.Google Scholar
Baddley, A., Conway, M., & Aggleton, J. (2002). Episodic memory. Oxford: Oxford University Press.Google Scholar
Bandura, A. (2006). Toward a psychology of human agency. Perspectives on Psychological Science, 1, 164–180.CrossRefGoogle Scholar
Barkow, J. H., Cosmides, L., & Tooby, J. (1992). The adapted mind. Oxford: Oxford University Press.Google Scholar
Baron, J. (1985). Rationality and intelligence. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Baron, J. (1988/2008). Thinking and deciding. Cambridge: Cambridge University Press.Google Scholar
Baron-Cohen, S. (1995/2000). Mindblindness. Cambridge, MA: MIT Press.Google Scholar
Baron-Cohen, S. (2008). Autism, hypersystemizing, and truth. Quarterly Journal of Experimental Psychology, 61, 64–75.CrossRefGoogle ScholarPubMed
Baron-Cohen, S., Knickmeyer, R. C., & Belmonte, M. K. (2005). Sex differences in the brain: Implications for explaining autism. Science, 310, 819–822.CrossRefGoogle ScholarPubMed
Baron-Cohen, S., Ring, H. A., Bullmore, E. T., Wheelwright, S., Ashwin, C., & Williams, S. C. R. (2000). The amygdala theory of autism. Neuroscience and Biobehavioral Review, 24, 355–364.CrossRefGoogle ScholarPubMed
Baron-Cohen, S., Ring, H. A., Wheelwright, S., Bullmore, E. T., Brammer, M. J., Simmons, A., & Williams, S. C. R. (1999). Social intelligence in the normal and autistic brain: An fMRI study. European Journal of Neuroscience, 11, 1–8.CrossRefGoogle Scholar
Baron-Cohen, S., Tager-Flushberg, H., & Cohen, D. J. (1993/2000). Understanding other minds. Oxford: Oxford University Press.Google Scholar
Barrett, J. L. (2000). Exploring the natural foundations of religion. Trends in Cognitive Sciences, 4, 29–33.CrossRefGoogle Scholar
Barrett, J. L., & Keil, F. C. (1996). Conceptualizing a non-natural entity: Anthropomorphism in God concepts. Cognitive Psychology, 31, 219–247.CrossRefGoogle Scholar
Barsalou, L. W. (1999). Perceptual symbol systems. Behavioral and Brain Sciences, 22, 577–660.Google ScholarPubMed
Barsalou, L. W. (2003). Abstraction in perceptual symbol systems. Philosophical Transactions of the Royal Society B, 358, 1177–1187.CrossRefGoogle ScholarPubMed
Barsalou, L. W. (2008). Grounded cognition. Annual Review of Psychology, 59, 617–645.CrossRefGoogle ScholarPubMed
Barsalou, L. W., Barbey, A. K., Simmons, W. K., & Satos, A. (2005). Embodiment in religious knowledge. Journal of Cognition and Culture, 5, 14–57.CrossRefGoogle Scholar
Barton, R. A. (2004). Binocularity and brain evolution in primates. Proceedings of the National Academy of Sciences, 101, 10113–10115.CrossRefGoogle ScholarPubMed
Barton, R. A. (2006). Primate brain evolution: Integrating comparative neurophysiological and ethological data. Evolutionary Anthropology, 15, 224–236.CrossRefGoogle Scholar
Bartsch, K., & Wellman, H. M. (1995). Children's talk about the mind. Oxford: Oxford University Press.Google Scholar
Barzun, J. (1989). The culture we deserve. Middletown, CT: Wesleyan University Press.Google Scholar
Bergson, H. (1908/1991). Matter and memory. New York: Zone Books.Google Scholar
Bergson, H. (1919/1946). The creative mind (M. Andison, Trans.). New York: Citadel Press.Google Scholar
Berlin, I. (1976). Vico and Herder. New York: Vintage Press.Google Scholar
Berlin, I. (1991). The crooked timer of humanity. New York: Knopf.Google Scholar
Bernard, C. (1865/1957). An introduction to the study of experimental medicine. New York: Dover Press.Google Scholar
Berridge, K. C. (2007). The debate over dopamine's role in reward: The case for incentive salience. Physiology and Behavior, 191, 391–431.Google ScholarPubMed
Berridge, K. C., & Robinson, T. E. (1998). What is the role of dopamine in reward: Hedonic impact, reward learning, or incentive salience?Brain Research Reviews, 28, 309–369.CrossRefGoogle ScholarPubMed
Berthoz, A. (2000). The brain's sense of movement. Cambridge, MA: Harvard University Press.Google Scholar
Bird-David, N. (1999). Animism revisited. Current Anthropology, 40, 567–591.Google Scholar
Black, J. E., Issacs, K. R., Anderson, B. J., Alcantara, A. A., & Greenough, W. T. (1990). Learning causes synaptogenesis, whereas motor activity causes angiogenesis, in cerebellar cortex of adult rats. Proceedings of the National Academy of Sciences, 87, 5568–5572.CrossRefGoogle ScholarPubMed
Blackbun, S. (1998). Ruling passion. Oxford, England: Clarendon.Google Scholar
Blakemore, S. J., Boyer, P., Pachot-Clouard, M., Meltzoff, A., Segebarth, C., & Decety, J. (2003). The detection of contingency and animacy from simple animations in the brain. Cerebral Cortex, 13, 837–844.CrossRefGoogle Scholar
Blakemore, S. J., & Decety, J. (2001). From the perception of action to the understanding of intention. Nature Reviews, 2, 561–567.Google ScholarPubMed
Bloch, M. (1953). The historian's craft. New York: Knopf.Google Scholar
Bloom, P. (2007). Religion is natural. Developmental Science, 10, 147–151.CrossRefGoogle ScholarPubMed
Boroditsky, L. (2000). Metaphoric structuring: Understanding time through spatial metaphors. Cognition, 75, 1–28.CrossRefGoogle ScholarPubMed
Boroditsky, L., & Ramscar, M. (2002). The roles of body and mind in abstract thought. Psychological Science, 13, 185–189.CrossRefGoogle ScholarPubMed
Bourdieu, P. (1980/1990). The logic of practice. Palo Alto, CA: Stanford University Press.Google Scholar
Boyd, R., & Richerson, P. (1985). Culture and evolutionary process. Chicago: University of Chicago Press.Google Scholar
Boyer, P. (1990). Tradition as truth and communication. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Boyer, P. (1994). The naturalness of religious idea: A cognitive theory of religion. Berkeley: University of California Press.Google Scholar
Boyer, P. (2001). Religion explained. New York: Basic Books.Google Scholar
Boyer, P. (2002). Religious thought and behavior as by-products of brain function. Trends in Cognitive Science, 7, 119–124.CrossRefGoogle Scholar
Boyer, P., & Ramble, C. (2001). Cognitive templates for religious concepts: Cross-cultural evidence for recall of counter-intuitive representations. Cognitive Science, 25, 535–564.CrossRefGoogle Scholar
Brehier, E. (1931/1965). The Helenistic and Roman age. Chicago: University of Chicago Press.Google Scholar
Brook, A., & Akins, K. (Eds.). (2005). Cognition and the brain. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Buber, M. (1937/1970). I and thou. New York: Harper & Row.Google Scholar
Buffon, G.-L. (1749–67). Histoire naturelle générale et particulière (Vols. 1–15). Paris: Imprimerie Royale.Google Scholar
Bunning, E. (1963). Die Physiologische Uhr, 2nd ed. Berlin: Springer.Google Scholar
Burckhardt, J. (1929/1958). The civilization of the Renaissance in Italy. New York: Colophon Books.Google Scholar
Burkhardt, R. W. (1977/1995). Lamarck and evolutionary biology. Cambridge, MA: Harvard University Press.Google Scholar
Bury, J. B. (1933/1960). The idea of progress. New York: Dover Press.Google Scholar
Butterfield, H. (1981). The origins of history. New York: Basic Books.Google Scholar
Buzsaki, G. (2006). Rhythms of the brain. Oxford: Oxford University Press.CrossRefGoogle Scholar
Byrne, R. W. (1995). The thinking ape: Evolutionary origins of intelligence. Oxford: Oxford University Press.CrossRefGoogle Scholar
Byrne, R. W., & Corp, N. (2004). Neocortex size predicts deception rate in primates. Proceedings of the Royal Society, 271, 1693–1699.CrossRefGoogle ScholarPubMed
Cajal, S. R. Y. (1897/1999). Advice for a young investigator. Cambridge, MA: MIT Press.Google Scholar
Cangelosi, A., & Riga, T. (2006). An embodied model for sensorimotor grounding and grounding transfer: Experiments with epigenetic robots. Cognitive Science, 30, 673–689.CrossRefGoogle ScholarPubMed
Cannon, W. B. (1932/1963). The wisdom of the body. New York: W. W. Norton.Google Scholar
Cannon, W. B. (1945). The way of an investigator. New York: W. W. Norton.Google Scholar
Caramazza, A., & Mahon, B. Z. (2005). The organization of conceptual knowledge in the brain: The future's past and some future directions. Cognitive Neuropsychology, 22, 1–25.Google Scholar
Caramazza, A., & Shelton, J. R. (1998). Domain specific knowledge systems in the brain. Journal of Cognitive Science, 10, 1–34.Google Scholar
Carey, S. (1985/1987). Conceptual change in childhood. Cambridge, MA: MIT Press.Google Scholar
Carey, S. (2001). Cognitive foundations of arithmetic: Evolution and ontogenesis. Mind and Language, 16, 37–55.CrossRefGoogle Scholar
Carey, S. (2003). Science education as conceptual change. Journal of Applied Developmental Psychology, 21, 13–19.CrossRefGoogle Scholar
Carey, S. (2004). Bootstrapping and the origins of concepts. Daedalus (Winter), 59–68.CrossRefGoogle Scholar
Carey, S., & Gelman, R. (1991). The epigenesis of mind: Essays on biology and cognition. Hillsdale, NJ: Lawrence Erlbaum Associates.Google Scholar
Carey, S., & Markman, E. M. (2000). Cognitive development. In B. M. Bly et al. (Eds.), Cognitive science: Handbook of perception and cognition (2d ed.). San Diego, CA: Academic Press, pp. 202–254.Google Scholar
Carey, S., & Smith, C. (1993). On understanding the nature of scientific knowledge. Educational Psychologist, 28, 235–251.CrossRefGoogle Scholar
Carruthers, P., Stitch, S., & Siegal, M. (2002). The cognitive basis of science. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Carter, C. S. (2007). Sex differences in oxytocin and vasopressin: Implications for autism spectrum disorders?Behavioural Brain Research, 176, 170–186.CrossRefGoogle ScholarPubMed
Carter, C. S., Lederhendler, I. L., & Kirkpatrick, B. (1997/1999). The integrative neurobiology of affiliation. Cambridge, MA: MIT Press.Google ScholarPubMed
Casasanto, D., & Boroditsky, L. (2008). Time in the mind: Using space to think about time. Cognition, 106, 579–593.CrossRefGoogle ScholarPubMed
Casebeer, W. D. (2003). Natural ethical facts. Cambridge, MA: MIT Press.Google Scholar
Cassirer, E. (1944/1978). An essay on man. New Haven, CT: Yale University Press.Google Scholar
Cassirer, E. (1946). Language and myth. New York: Harper & Row.Google Scholar
Cassirer, E. (1953/1957). Philosophy of symbolic forms (Vols. 1–3). New Haven, CT: Yale University Press.Google Scholar
Castelli, F., Happe, F., Frith, U., & Frith, C. (2000). Movement and mind: A functional imaging study of perception and interpretation of complex intentional movement patterns. Neuroimage, 12, 314–325.CrossRefGoogle ScholarPubMed
Chaminade, T., Meltzoff, A. N., & Decety, J. (2002). Does the end justify the means? A PET exploration of the mechanisms involved in human imitation. Neuroimage, 15, 318–328.CrossRefGoogle Scholar
Chan, W. J. (1963). A source book in Chinese philosophy. Princeton, NJ: Princeton University Press.Google Scholar
Changeux, J. P. (2002). The physiology of truth. Cambridge, MA: Harvard University Press.Google Scholar
Cheney, D. L., & Seyfarth, R. M. (1990). How monkeys see the world. Chicago:University of Chicago Press.Google Scholar
Cheney, D. L., & Seyfarth, R. M. (2007). Baboon metaphysics. Chicago: University of Chicago Press.CrossRefGoogle Scholar
Cherniak, C. (1986). Minimal rationality. Cambridge, MA: MIT Press.Google Scholar
Chomsky, N. (1965). Aspects of the theory of syntax. Cambridge, MA: MIT Press.Google Scholar
Chomsky, N. (1972). Language and mind. New York: Harcourt Brace Jovanovich.Google Scholar
Clark, A. (1997). Being there. Cambridge, MA: MIT Press.Google Scholar
Clark, A. (1999). An embodied cognitive science?Trends in Cognitive Sciences, 3, 345–351.CrossRefGoogle Scholar
Clark, A. (2003). Natural-born cyborgs. Oxford: Oxford University Press.Google Scholar
Clark, W. (2006). Academic charisma and the origins of the research university. Chicago: University of Chicago Press.Google Scholar
Comery, T. A., Stamoudis, C. X., Irwin, S. A., & Greenough, W. T. (1996). Increased density of multiple-head dendritic spines on medium-sized spiny neurons of the striatum in rats reared in a complex environment. Neurobiology of Learning and Memory, 66, 93–96.CrossRefGoogle Scholar
Cohen, M. R. (1931/1959). Reason and nature. New York: Dover Press.Google Scholar
Collingwood, R. G. (1929/1978). An autobiography. Oxford: Oxford University Press.Google Scholar
Collingwood, R. G. (1945/1976). The idea of nature. Oxford: Oxford University Press.Google Scholar
Collingwood, R. G. (1946/1956). The idea of history. Oxford: Oxford University Press.Google Scholar
Collingwood, R. G. (2001). The principles of history (W. H. Dray & W. J. Van Der Dussen, Eds.). Oxford: Oxford University Press.Google Scholar
Comte, A. C. (1842/1975). A general view of positivism (J. Bridges, Trans.). New York: Robert Speller & Sons.Google Scholar
Conze, E. (1951/1975). Buddhism: Its essence and development. New York: Harper & Row.Google Scholar
Conze, E. (1959/1984). Buddhistic scriptures. New York: Penguin Books.Google Scholar
Corballis, M. C. (2002). From hand to mouth. Princeton, NJ: Princeton University Press.Google Scholar
Corballis, M. C., & Lea, S. E. G. (Eds.). (1999). The descent of mind. New York: Oxford University Press.Google Scholar
Craig, W. (1918). Appetites and aversions as constituents of instinct. Biological Bulletin, 34, 91–107.CrossRefGoogle Scholar
Critchley, H. D. (2005). Neural mechanisms of autonomic, affective, and cognitive integration. Journal of Comparative Neurology, 493, 154–166.CrossRefGoogle ScholarPubMed
Critchley, H. D., Daly, E. M., Bullmore, E. T., Williams, S. C. R., Amelsvoort, T., Robertson, D. M., et al. (2000). The functional neuroanatomy of social behaviour: Changes in cerebral blood flow when people with autistic disorder process facial expressions. Brain, 123, 2203–2212.CrossRefGoogle ScholarPubMed
Critchley, H. D., Mathias, C. J., Josephs, O., O'Doherty, J., Zanini, S., Dewar, B. K., et al. (2003). Human cingulate cortex and autonomic control: Converging neuroimaging and clinical evidence. Brain, 126, 2139–2152.CrossRefGoogle ScholarPubMed
Croce, B. (1941). History as the story of liberty. Chicago: Henry Regnery Co.Google Scholar
Dallman, M. F. (2007). Modulation of stress responses: How we cope with excess glucocorticoids. Experimental Neurology, 206, 179–182.CrossRefGoogle ScholarPubMed
Damasio, A. R. (1996). The somatic marker hypothesis and the possible functions of the prefrontal cortex. Philosophical Transactions of the Royal Society, 351, 1413–1420.CrossRefGoogle ScholarPubMed
Damasio, A. R. (1999). The feeling of what happens. New York: Harcourt.Google Scholar
Darwin, C. (1859/1958). The origin of species. New York: Mentor Books.Google Scholar
Darwin, C. (1871/1874). The descent of man. New York: Rand McNally.Google Scholar
Darwin, C. (1872/1965). The expression of emotions in man and animals. Chicago: University of Chicago Press.CrossRefGoogle Scholar
Darwin, E. (1789/1991). The book of plants: Vol. 2. Oxford: Woodstock Books.Google Scholar
Darwin, E. (1794/1801). Zoonomia, or the law of organic life (3d ed.). London: J. Johnson.Google Scholar
Dasser, V., Ulbaek, I., & Premack, D. (1989). The perception of intention. Science, 243, 365–367.CrossRefGoogle ScholarPubMed
Davidson, R. J., & Harrington, A. (2002). Visions of compassion. Oxford: Oxford University Press.Google Scholar
Davidson, R. J., Kabat-Zinn, J., Schumacher, J., Rosenkranz, B. A., Muller, M., Saki, F., et al. (2003). Alterations in brain and immune function produced by mindfulness meditation. Psychosomatic Medicine, 65, 564–570.CrossRefGoogle ScholarPubMed
Dawkins, R. (2006). The God delusion. Boston: Houghton Mifflin.Google Scholar
Dayes, R. M. (2001). Everyday irrationality. Boulder, CO: Westview Press.Google Scholar
Deacon, T. W. (1997). The symbolic species. New York: W. W. Norton.Google Scholar
Dear, P. (1995). Discipline and experience. Chicago: University of Chicago Press.CrossRefGoogle Scholar
Dear, P. (2006). The intelligibility of nature. Chicago: University of Chicago Press.CrossRefGoogle Scholar
Decety, J. (1996). Do imagined and executed actions share the same neural substrate?Cognitive Brain Research, 3, 87–93.CrossRefGoogle ScholarPubMed
Decety, J., & Jackson, P. W. (2006). A social neuroscience perspective on empathy. Current Directions in Psychological Science, 15, 54–58.CrossRefGoogle Scholar
Decety, J., Perani, D., & Jeannerod, M. (1994). Mapping motor representations with positron emission tomography. Nature, 371, 600–602.CrossRefGoogle ScholarPubMed
Deheane, S., Izard, V., Pica, P., & Spelke, E. (2006). Core knowledge of geometry in an Amazonian Indigene group. Science, 311, 381–384.CrossRefGoogle Scholar
Delza, S. (1961/1985). Tai-chi ch'uan. Albany: State University of New York Press.Google Scholar
Dennett, D. (1987). The intentional stance. Cambridge, MA: MIT Press.Google Scholar
Dennett, D. (2006). Breaking the spell. New York: Viking Press.Google Scholar
Derrida, J. (1972/1981). Dissemination. Chicago: University of Chicago Press.Google Scholar
Descartes, R. (1663/1967). Discourse on method and meditations. Indianapolis: Bobbs-Mettill Co.Google Scholar
Waal, F. B. M. (2000). Primates: A natural heritage of conflict resolution. Science, 289, 586–590.CrossRefGoogle ScholarPubMed
Waal, F., & Lanting, F. (1997). Bonobo: The forgotten ape. Berkeley: University of California Press.Google Scholar
Dewey, J. (1896). The reflex arc concept in psychology. Psychological Review, 3, 357–370.CrossRefGoogle Scholar
Dewey, J. (1902/1974). The child and the curriculum and the school and society. Chicago: University of Chicago Press.Google Scholar
Dewey, J. (1908/1960). Theory of moral life. New York: Holt, Rinehart & Winston.Google Scholar
Dewey, J. (1909/1975). Moral principles in education. Carbondale: Southern Illinois University Press.Google Scholar
Dewey, J. (1910/1965). The influence of Darwin on philosophy. Bloomington: Indiana University Press.Google Scholar
Dewey, J. (1916). Essays in experimental logic. Chicago: University of Chicago Press.CrossRefGoogle Scholar
Dewey, J. (1920/1948). Reconstruction in philosophy. Boston: Beacon Press.CrossRefGoogle Scholar
Dewey, J. (1925/1989). Experience and nature. La Salle, IL: Open Court.Google Scholar
Dewey, J. (1929/1960). The quest for certainty. New York: Capricorn Books.Google Scholar
Dewey, J. (1934/1970). A common faith. New Haven, CT: Yale University Press.Google Scholar
Dewey, J. (1938). Logic: The theory of inquiry. New York: Henry Holt and Co., Inc.Google Scholar
Dewey, J. (1938/1972). Experience and education. New York: Collier-MacMillan.Google Scholar
Dickinson, E. (1960). Selected poems. New York: Dell.Google Scholar
Dilthey, W. (1926/1961). Pattern and meaning in history (H. Richman, Trans.). New York: HarperTorch.Google Scholar
di Pellegrino, G., Fadiga, L., Fogassi, L., Gallese, V., & Rizzolatti, G. (1992). Understanding motor events: A neurophysiological study. Experimental Brain Research, 91, 176–180.CrossRefGoogle ScholarPubMed
Dobzhansky, T. C. (1962). Mankind evolving. New Haven, CT: Yale University Press.Google Scholar
Dolan, R. J. (2007). The human amygdala and orbital prefrontal cortex in behavioural regulation. Philosophical Transactions of the Royal Society B, 362, 787–799.CrossRefGoogle ScholarPubMed
Dolgin, K. G., & Behrend, D. A. (1984). Children's knowledge about animates and inanimates. Child Development, 35, 1645–1650.Google Scholar
Donald, M. (1991). Origins of modern man. Cambridge, MA: Harvard University Press.Google Scholar
Donald, M. (2004). The virtues of rigorous interdisciplinarity. In J. M. Luraciello, J. A. Hudson, R. Fivush, & P. J. Baver (Eds.), The development of the mediated mind. Mahwah, NJ: Lawrence Erlbaum Associates, pp. 245–256.Google Scholar
Duchaine, B., Cosmides, L., & Tooby, H. (2001). Evolutionary psychology and the brain. Current Opinion in Neurobiology, 11, 225–250.CrossRefGoogle Scholar
Dunbar, R. I. M. (1992). Neocortex size as a constraint on group size in primates. Journal of Human Evolution, 22, 469–493.CrossRefGoogle Scholar
Dunbar, R. I. M. (1996). Grooming, gossip and the evolution of language. London: Faber & Faber.Google Scholar
Dunbar, R. I. M. (2003). The social brain. Annual Review of Anthropology, 32, 163–181.CrossRefGoogle Scholar
Dunbar, R. I. M., & Shultz, S. (2007). Understanding primate evolution. Philosophical Transactions of the Royal Society, 362, 649–658.CrossRefGoogle Scholar
Dupré, J. (1993). The disorder of things. Cambridge, MA: Harvard University Press.Google Scholar
Dupré, J. (2003). Darwin's legacy. Oxford: Oxford University Press.Google Scholar
Durkheim, E. (1915/1965). The elementary forms of religious life. New York: Free Press.Google Scholar
Durkheim, E. (1974). Sociology and philosophy (D. F. Pocock, Trans.). New York: Free Press.Google Scholar
Edwards, J. (1754/1969). Freedom of the will (A. S. Kaufman & W. K. Frankena, Eds.). New York: Bobbs-Merrill.Google Scholar
Eichenbaum, H., & Cohen, N. J. (2001). From conditioning to conscious recollection. Oxford: Oxford University Press.Google Scholar
Eiseley, L. (1959/1961). Darwin's century. New York: Anchor Books.Google Scholar
Ekman, P. (1972). Universals and cultural differences in facial expressions of emotion. In Cole, J. (Ed.), Nebraska symposium on motivation, 1971. Lincoln: University of Nebraska Press, pp. 207–283.Google Scholar
Ekman, P., Davidson, R. J., Ricard, M., & Wallace, B. A. (2005). Buddhist and psychological perspectives on emotion and well-being. Current Directions in Psychological Science, 14, 59–63.CrossRefGoogle Scholar
Eliade, M. (1959). The sacred and the profane; the nature of religion. New York: Harcourt.Google Scholar
Eldridge, N. (1985). Unfinished synthesis. Oxford: Oxford University Press.Google Scholar
Eldridge, N. (1999). The pattern of evolution. New York: W. H. Freeman.Google Scholar
Elster, J. (1979/1988). Ulysses and the Sirens. Cambridge: Cambridge University Press.Google Scholar
Elster, J. (1983/1988). Explaining technical change. Cambridge: Cambridge University Press.Google Scholar
Emerson, Ralph Waldo (1855/1883). Nature, addresses and lectures. Cambridge: Riverside Press.Google Scholar
Emery, N. J., & Amaral, D. G. (2000). The role of the amygdala in primate social cognition. In R. D. Lane & L. Nadel (Eds.), Cognitive neuroscience of emotion. New York: Oxford University Press, pp. 156–191.Google Scholar
Emery, N. J., Lorincz, E. N., Perrett, D. I., Oram, M. W., & Baker, C. I. (1997). Gaze following and joint attention in Rhesus monkeys (Macaca mulatta). Journal of Comparative Psychology, 111, 286–293.CrossRefGoogle Scholar
Erickson, K., Drevets, W. C., & Schulkin, J. (2003). Glucocorticoid regulation of diverse cognitive functions in normal and pathological emotional states. Neuroscience and Biobehavioral Reviews, 27, 233–246.CrossRefGoogle ScholarPubMed
Erickson, K., Mah, L., Schulkin, J., Charney, W. S., & Drevets, W. C. (2005). Effects of hydrocorticosterone infusion on affectively valenced autobiographical memory. Neuroscience Abstracts.Google Scholar
Fauconnier, G., & Turner, M. (2003). The way we think. New York: Basic BooksGoogle Scholar
Fesmire, S. (2003). John Dewey and moral imagination: Pragmatism in ethics. Bloomington: Indiana University Press.Google Scholar
Finley, M. I. (1971). The use and abuse of history. New York: Penguin Books.Google Scholar
Fiorillo, C. D., Tobler, P. N., & Schultz, W. (2003). Discrete coding of reward probability and uncertainty by dopamine neurons. Science, 299, 1898–1902.CrossRefGoogle ScholarPubMed
Fiske, A. P. (1991). Structures of social life. New York: Free Press.Google Scholar
Flanagan, O. (2007). The really hard problems. Cambridge, MA: MIT Press.Google Scholar
Fleagle, J. G. (1988). Primate adaptation and evolution. San Diego: Academic Press.Google Scholar
Flower, E., & Murphy, M. G. (1977). A history of philosophy in America (Vols. 1–2). New York: Capricorn Books.Google Scholar
Fodor, J. (1983). The modularity of mind. Cambridge, MA: MIT Press.Google Scholar
Fogassi, L., & Luppino, G. (2005). Motor functions of the parietal lobe. Current Opinion in Neurobiology, 15, 626–631.CrossRefGoogle ScholarPubMed
Fogassi, L., Ferrari, P. F., Gesierich, B., Rozzi, S., Chersi, F., & Rizzolatti, G. (2005). Parietal lobe: From action organization to intention understanding. Science, 308, 662–666.CrossRefGoogle ScholarPubMed
Foley, R. (1995). Humans before humanity. Oxford, England: Blackwell.Google Scholar
Foley, R. (1996). An evolutionary and chronological framework for human social behaviour. Proceedings of the British Academy, 88, 95–117.Google Scholar
Foley, R. (2001). In the shadow of the modern synthesis?Evolutionary Anthropology, 10, 5–14.3.0.CO;2-Y>CrossRefGoogle Scholar
Foley, R., & Lahr, M. M. (2003). On stony ground: Lithic technology, human evolution and the emergence of culture. Evolutionary Anthropology, 12, 109–122.CrossRefGoogle Scholar
Foley, R., & Lahr, M. M. (2004). Human evolution writ small. Nature, 431, 1043–1044.Google Scholar
Fonberg, E. (1974). Amygdala function within the alimentary system. Aeta Neurobiologae Experimentalis, 22, 51–57.Google Scholar
Foner, E. (2002). Who owns history?New York: Hill & Wang.Google Scholar
Foster, R. G., & Kreitzman, L. (2004). Rhythms of life. New Haven, CT: Yale University Press.Google Scholar
Franklin, B. (1726, 1757/1987). Essays, articles, bagatelles, and letters: Poor Richard's almanac. New York: Library Classics.Google Scholar
Frazer, J. H. (1921/2000). The new golden bough. New York: Bartleby.Google Scholar
Freud, S. (1924/1960). A general introduction to psychoanalysis. New York: Washington Square Press.Google Scholar
Freud, S. (1927/1964). The future of an illusion. New York: Doubleday.Google Scholar
Frith, C. D. (2007). The social brain?Philosophical Transactions of the Royal Society B, 362, 671–678.CrossRefGoogle ScholarPubMed
Frith, C. D., & Frith, U. (1999). Interacting minds: A biological basis. Science, 286, 1692–1694.CrossRefGoogle ScholarPubMed
Frith, C., & Wolpert, D. (2003). The neuroscience of social interaction. Oxford: Oxford University Press.Google Scholar
Fromm, E. (1947). Man for himself. New York: Rinehart.Google Scholar
Fromm, E. (1950/1972). Psychoanalysis and religion. New York: Doubleday.Google Scholar
Fromm, E., Suzuki, D. T., & Martino, R. (1960/1970). Zen Buddhism and psychoanalysis. New York: Harper & Row.Google Scholar
Fuller, S. (1988). Social epistemology. Bloomington: Indiana University Press.Google Scholar
Gage, F. H. (1998). Stem-cells of the central nervous system. Current Opinion in Neurobiology, 8, 671–676.CrossRefGoogle ScholarPubMed
Galileo, G. (1957). Discoveries and opinions. New York: Anchor Books.Google Scholar
Galison, P. (1988). History, philosophy and the central metaphor. Science in Context, 3, 197–212.Google Scholar
Galison, P. (1999). Objectivity is romantic. In The Humanities and the Sciences. American Council of Learned Societies. Occasional Paper No 47, 15–43.Google Scholar
Gallagher, S. (2005). How the body shapes the mind. Oxford: Oxford University Press.CrossRefGoogle Scholar
Gallagher, S., & Meltzoff, A. N. (1996). The earliest sense of self and others: Merleau-Ponty and recent developmental studies. Philosophical Psychology, 9, 211–233.CrossRefGoogle Scholar
Gallese, V. (2007). Before and below ‘theory of mind’: Embodied simulation and the neural correlates of social cognition. Philosophical Transactions of the Royal Society B, 362, 659–669.CrossRefGoogle ScholarPubMed
Gallistel, C. R. (1990). The organization of learning. Cambridge, MA: MIT Press.Google Scholar
Gandhi, M. (1964). On non-violence (T. Merton, Ed.). New York: New Directions.Google Scholar
Gardenfors, P. (2003). How Homo became sapiens: On the evolution of thinking. Oxford: Oxford University Press.Google Scholar
Gardner, H. (1985). The mind's new science. New York: Basic Books.Google Scholar
Gazzaniga, M. S. (1995/2000). The new cognitive neurosciences. Cambridge, MA: MIT Press.Google Scholar
Gazzaniga, M. S. (1998). The mind's past. Berkeley: University of California Press.Google Scholar
Gelman, R., Durgin, F., & Kaufman, L. (1995). Distinguishing between animates and inanimates: Not by motion alone. In D. Sperber, D. Premack, & A. Premack (Eds.), Causal cognition. Oxford, England: Clarendon Press, pp. 150–184.Google Scholar
Gelman, R., Spelke, E. S., & Meck, E. (1983). What preschoolers know about animate and inanimate objects. In D. Rogers & J. A. Sloboda (Eds.), The acquisition of symbolic skills. New York: Plenum, pp. 297–326.Google Scholar
Gelman, S. A. (2003). The essential child. Oxford: Oxford University Press.CrossRefGoogle Scholar
Gelman, S. A., & Markman, E. M. (1987). Young children's inductions from natural kinds. Child Development, 58, 1532–1541.CrossRefGoogle ScholarPubMed
Geschwind, N. (1974). Selected papers on language and the brain. Boston: Reidel Publishing.CrossRefGoogle Scholar
Gibbon, E. (1788/1952). The decline and fall of the Roman Empire. New York: Penguin Books.Google Scholar
Gibbs, R. W. (2006). Embodiment and cognitive science. Cambridge: Cambridge University Press.Google Scholar
Gibson, J. J. (1966). The senses considered as perceptual systems. New York: Houghton Mifflin.Google Scholar
Gibson, K. R., & Ingold, T. (Eds.). (1993). Tools, language and cognition in human evolution. Cambridge: Cambridge University Press.Google Scholar
Giere, R. N. (2006). Scientific perspectivism. Chicago: University of Chicago Press.CrossRefGoogle Scholar
Gigerenzer, G. (2007). Gut feelings. New York: Viking Press.Google Scholar
Gigerenzer, G., & Selten, R. (2001). Bounded rationality. Cambridge, MA: MIT Press.Google Scholar
Gigerenzer, G. (2000). Adaptive thinking: Rationality in the real world. New York: Oxford University Press.Google Scholar
Glenberg, A. M. (1997). What memory is for. Behavioral and Brain Sciences, 20, 1–55.Google ScholarPubMed
Glenberg, A. M., Gutierrez, T., Levin, J. R., Japuntich, S., & Kaschak, M. P. (2004). Activity and imagined activity can enhance young children's reading comprehension. Journal of Educational Psychology, 96, 424–436.CrossRefGoogle Scholar
Glenberg, A. M., & Kaschak, M. P. (2002). Grounding language in action. Psychonomic Bulletin & Review, 9, 558–565.CrossRefGoogle ScholarPubMed
Gloor, P. (1997). The temporal lobe and limbic system. New York: Oxford University Press.Google Scholar
Godfrey-Smith, P. (1996). Complexity and the function of mind. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Godfrey-Smith, P. (2002). Dewey on naturalism, realism and science. Philosophy of Science, 69, S1–S11.CrossRefGoogle Scholar
Goffman, E. (1971). Relations in public. New York: Harper & Row.Google Scholar
Goldman, A. I. (1999). Knowledge in a social world. Oxford: Oxford University Press.CrossRefGoogle Scholar
Goldsmith, R. (1940/1982). The material basis of evolution. New Haven, CT: Yale University Press.Google Scholar
Gopnik, A., Glymour, C., Sobel, D., Schulz, L., Kushnir, T., & Danks, D. (2004). A theory of causal learning in children: Causal maps and Bayes nets. Psychological Review, 111, 1–31.CrossRefGoogle ScholarPubMed
Gopnik, A., & Meltzoff, A. N. (1997). Words, thoughts and theories. Cambridge, MA: MIT Press.Google Scholar
Gould, E., Beylin, A., Tanapat, P., Reeves, A., & Shors, T. J. (1999). Learning enhances adult neurogenesis in the hippocampal formation. Nature Neuroscience, 2, 260–265.CrossRefGoogle ScholarPubMed
Gould, E., & McEwen, B. (1993). Neuronal birth and death. Current Opinion in Neurobiology, 3, 676–682.CrossRefGoogle ScholarPubMed
Gould, E., Reeves, A. J., Craziano, M. S. A., & Gross, C. G. (1999). Neurogenesis in the neocortex of adult primates. Science, 286, 548–552.CrossRefGoogle ScholarPubMed
Gould, E., Vall, N., Wagers, M., & Gross, C. G. (2001). Adult generated hippocampal and neocortical neurons in macaques have a transient existence. Proceedings of the National Academy of Sciences, 98, 10910–10916.CrossRefGoogle ScholarPubMed
Gould, S. J. (1977). Ontogeny and phylogeny. Cambridge, MA: Harvard University Press.Google Scholar
Gould, S. J. (2002). The structure of evolutionary theory. Cambridge, MA: Harvard University Press.Google Scholar
Gould, S. J., & Eldridge, N. (1977). Punctuated equilibria: The tempo and mode of evolution reconsidered. Paleobiology, 3, 115–151.CrossRefGoogle Scholar
Gould, S. J., & Lewontin, R. C. (1979). The spandrels of San Marco and the Panglossian paradigm: A critique of the adaptationist programme. Proceedings of the Royal Society B, 205(1161), 581–598.CrossRefGoogle Scholar
Green, L., Fein, D., Modahl, C., Feinstein, C., Waterhouse, L., & Morris, M. (2001). Oxytocin and autistic disorder: Alterations in peptide forms. Biological Psychiatry, 50, 609–613.CrossRefGoogle ScholarPubMed
Greene, J. D., Sommerville, R. B., Nystrom, L. E., Darley, J. M., & Cohen, J. D. (2001). An fMRI investigation of emotional engagement in moral judgment. Science, 293, 2105–2108.CrossRefGoogle ScholarPubMed
Greenough, W. T., & Volkmar, F. R. (1973). Pattern of dendritic branching in occipital cortex of rats reared in complex environments. Experimental Neurology, 40, 491–504.CrossRefGoogle ScholarPubMed
Grene, M. (1995). A philosophical testament. Chicago: Open Court.Google Scholar
Grice, P. (1957). Meaning. The Philosophical Review, 3, 377–388.CrossRefGoogle Scholar
Guthrie, R. D. (2005). The nature of Paleolithic art. Chicago: University of Chicago Press.Google Scholar
Habermas, J. (1967/1988). On the logic of the social sciences (S. W. Nicholson & J. A. Stark, Trans.). Cambridge, MA: MIT Press.Google Scholar
Hacking, I. (1975). The emergence of probability. Cambridge: Cambridge University Press.Google Scholar
Hacking, I. (1990). The taming of chance. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Hacking, I. (1999). The social construction of what?Cambridge, MA: Harvard University Press.Google Scholar
Haidt, J. (2006). The happiness hypothesis: Finding modern truth in ancient wisdom. New York: Basic Books.Google Scholar
Hanson, N. R. (1958/1972). Patterns of discovery. Cambridge: Cambridge University Press.Google Scholar
Hari, R., Forss, N., Avikainen, S., Kirverskari, E., Salenius, S., & Rizzolatti, G. (1998). Activation of human primary cortex during action observation: A neuromagnetic study. Proceedings of the National Academy of Sciences, 95, 15061–15065.CrossRefGoogle ScholarPubMed
Harrington, A., & Zajonc, A. (2006). The Dalai Lama. Cambridge, MA: MIT Press.Google Scholar
Harvey, W. (1651/1965). Anatomical exercises on the generation of animals. New York: Johnson Reprint Corp.Google Scholar
Haskell, T. L. (1998). Objectivity is not neutrality. Baltimore: Johns Hopkins University Press.Google Scholar
Haskins, C. H. (1923/1957). The rise of the universities. Ithaca, NY: Cornell University Press.Google Scholar
Hauk, O., Johnsrude, I., & Pulvermuller, F. (2004). Somatotopic representation of action words in human motor and premotor cortex. Neuron, 41, 301–307.CrossRefGoogle ScholarPubMed
Hauser, M. D. (2000). Wild minds. New York: Henry Holt.Google Scholar
Hauser, M. D., Chomsky, N., & Fitch, W. T. (2002). The faculty of language: What is it, who has it and how did it evolve?Science, 298, 1569–1579.CrossRefGoogle ScholarPubMed
Heelan, P. A. (1983). Space perception and the philosophy of science. Berkeley: University of California Press.Google Scholar
Heelan, P. A. (1994). Galileo, Luther, and the hermeneutics of natural science: The question of hermeneutics (T. J. Stapleton, Ed.). Netherlands: Kluwer Academic Publishers.CrossRefGoogle Scholar
Heelan, P. A., & Schulkin, J. (1998). Hermeneutical philosophy and pragmatism: A philosophy of the science. Synthese, 115, 269–302.CrossRefGoogle Scholar
Heidegger, M. (1927/1962). Being and time (J. Macquarrie & E. Robinson, Trans.). New York: Harper & Row.Google Scholar
Hendel, C. W. (1959). John Dewey and the experimental spirit in philosophy. New Haven, CT: Yale University Press.Google Scholar
Herbert, J., & Schulkin, J. (2002). Neurochemical coding of adaptive responses in the limbic system. In Pfaff, D. (Ed.), Hormones, brain and behavior. New York: Elsevier, pp. 659–689.Google Scholar
Herodotus, (1954). The histories (A. de Selincourt, Trans.). Baltimore: Penguin Books.Google Scholar
Himmelfarb, G. (1959/1962). Darwin and the Darwinian revolution. New York: W. W. Norton.Google Scholar
Hirschfield, L. A., & Gelman, S. A. (1994). Mapping the mind. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Hirshman, A. O. (1982). Shifting involvements. Princeton, NJ: Princeton University Press.Google Scholar
Hollander, E., Bartz, J., Chaplin, W., Phillips, A., Sumner, J., Soorya, L., et al. (2006). Oxytocin increases retention of social cognition in autism. Biological Psychiatry, 61, 498–503.CrossRefGoogle ScholarPubMed
Hollander, E., Novotny, S., Hanratty, M., Yaffe, R., DeCaria, C. M., Aronowitz, B. R., et al. (2003). Oxytocin infusion reduces repetitive behaviors in adults with autistic and Asperger's disorders. Neuropsychopharmacology, 28, 193–198.CrossRefGoogle ScholarPubMed
Houser, N., Eller, J. R., Lewis, A. C., Tienne, A., Clark, C. L., & Bront Davis, D. (1998). The essential Peirce: Vol. 2 (1893–1913). Bloomington: Indiana University Press.Google Scholar
Houser, N., & Kloesel, C. (1992/1998). The essential Peirce: Vol. 1 (1867–1893). Bloomington: Indiana University Press.Google Scholar
Houser, N., Roberts, D. D. & Evra, , J. (Eds.). (1997). Studies in the logic of Charles Sanders Peirce. Indianapolis: University of Indiana Press.Google Scholar
Howells, W. (1963). Back of history. New York: Anchor Books.Google Scholar
Hull, D. L. (1988). Science as a process. Chicago: University of Chicago Press.CrossRefGoogle Scholar
Humbolt, W. V. (1836/1971). Linguistic and intellectual development. Philadelphia: University of Pennsylvania Press.Google Scholar
Hume, D. (1757/1957). The natural history of religion. Stanford, CA: Stanford University Press.Google Scholar
Humphrey, N. (1976). The social function of intellect. In P. P. G. Bateson & R. A. Hinde (Eds.), Growing points in ethology. Cambridge: Cambridge University Press, pp. 307–317.Google Scholar
Humphrey, N. (2000). Cave art, autism and the human mind. Journal of Consciousness Studies, 6, 116–123.Google Scholar
Huxley, T. H. (1863). Man's place in nature. London: Macmillan.Google Scholar
Huxley, T. H. (1909). Autobiography and selected essays. Cambridge, MA: Houghton Mifflin.Google Scholar
Iacoboni, M. (2005). Neural mechanisms of imitation. Current Opinions in Neurobiology, 15(6), 632–637.CrossRefGoogle ScholarPubMed
Iacoboni, M., Lieberman, M. D., Knowlton, B. J., Molnar-Szakacs, I., Moritz, M., Throop, J., & Fiske, A. P. (2004). Watching social interactions produces dorsomedial prefrontal and medial parietal BOLD fMRI signal increases compared to a resting baseline. NeuroImage, 21, 1167–1173.CrossRefGoogle ScholarPubMed
Iacoboni, M., Woods, R. P., Brass, M., Bekkering, H., Mazziotta, J. C., & Rizzolatti, G. (1999). Cortical mechanisms of imitation. Science, 286, 2526–2528.CrossRefGoogle ScholarPubMed
Insel, T. R., & Fernald, R. D. (2004). How the brain processes social information. Annual Review of Neuroscience, 27, 697–722.CrossRefGoogle ScholarPubMed
Insel, T. R., O'Brien, D. J., & Leckman, J. F. (1999). Oxytocin, vasopressin, and autism: Is there a connection?Biological Psychiatry, 45, 145–157.CrossRefGoogle Scholar
Israel, J. I. (2001). Radical enlightenment. Oxford: Oxford University Press.CrossRefGoogle Scholar
Jackendoff, R. (1992). Language of the mind. Cambridge, MA: MIT Press.Google Scholar
Jackson, J. H. (1884/1958). Evolution and dissolution of the nervous system. In Taylor, J. (Ed.), Collected works of John Hughlings Jackson: Vol. 11. London: Staples Press, pp. 45–118.Google Scholar
Jackson, P. L., & Decety, J. (2004). Motor cognition. Current Opinion in Neurobiology, 14, 259–263.CrossRefGoogle ScholarPubMed
Jacob, P., & Jeannerod, M. (2003). Ways of seeing. Oxford: Oxford University Press.CrossRefGoogle Scholar
Jacob, P., & Jeannerod, M. (2005). The motor theory of social cognition: A critique. Trends in Cognitive Science, 9, 21–25.CrossRefGoogle ScholarPubMed
James, W. (1887). Some human instincts. Popular Science Monthly, 31, 160–176.Google Scholar
James, W. (1890/1917). Principles of psychology. New York: Henry Holt.Google Scholar
James, W. (1896/1956). The will to believe, human immortality. New York: Dover Press.Google Scholar
James, W. (1902/1974). The varieties of religious experience: A study in human nature. New York: Collier Books, Macmillan.CrossRefGoogle Scholar
James, W. (1910/1970). Pragmatism and other essays. New York: Washington Square Press.Google Scholar
Jaspers, K. (1913/1997). General psychopathology. Baltimore: Johns Hopkins University Press.Google Scholar
Jaspers, K. (1949/1968). The origin and goal of history (M. Bullock, Trans.). New Haven, CT: Yale University.Google Scholar
Jaspers, K. (1951/1954). Way to wisdom. New Haven, CT: Yale University Press.Google Scholar
Jeannerod, M. (1999). To act or not to act: Perspectives on the representation of action. Quarterly Journal of Experimental Psychology, 52, 1–29.CrossRefGoogle ScholarPubMed
Jevning, R., Wilson, A. F., & VanderLaan, E. F. (1978). Plasma prolactin and growth hormone during meditation. Psychosomatic Medicine, 40, 329–333.CrossRefGoogle ScholarPubMed
Johanson, D. C., & Edey, M. (1981). Lucy: The beginnings of humankind. New York: Simon & Schuster.Google Scholar
Johnson, M. (1987/1990). The body in the mind. Chicago: University of Chicago Press.Google Scholar
Johnson, M. (1993). Moral imagination. Chicago: University of Chicago Press.Google Scholar
Johnson, M. (2007). The meaning of the body. Chicago: University of Chicago Press.CrossRefGoogle Scholar
Johnson, M., & Rohrer, T. (2007). We are live creatures: Embodiment, American pragmatism and the cognitive organism. In Body, language and mind: Vol. 1. Berlin: Mouton de Gruyter, pp. 17–54.Google Scholar
Johnson-Laird, P. N. (2001). Mental models and deduction. Trends in Cognitive Science, 5, 434–442.CrossRefGoogle ScholarPubMed
Johnson-Laird, P. N. (2002). Peirce, logic diagrams, and the elementary operations of reasoning. Thinking and Reasoning, 8, 69–95.CrossRefGoogle Scholar
Jolly, A. (1966). Lemur social behavior and primate intelligence. Science, 153, 501–506.CrossRefGoogle ScholarPubMed
Jolly, A. (1999). Lucy's legacy. Cambridge, MA: Harvard University Press.Google Scholar
Kagan, J. (1984). The nature of the child. New York: Basic Books.Google Scholar
Kagan, J. (2002). Surprise, uncertainty and mental structure. Cambridge, MA: Harvard University Press.Google Scholar
Kahneman, D., Slovic, P., & Tversky, A. (Eds.). (1982). Judgment under uncertainty: Heuristics and biases. New York: Cambridge University Press.CrossRefGoogle Scholar
Kahneman, D., & Tversky, A. (1973). On the psychology of prediction. Psychological Review, 80, 237–251.CrossRefGoogle Scholar
Kant, I. (1787/1965). Critique of pure reason (L. W. Beck, Trans.). New York: St. Martin's Press.Google Scholar
Kant, I. (1788/1956). Critique of practical reason (L. W. Beck, Trans.). New York: Bobbs-Merrill.Google Scholar
Kant, I. (1792/1951). Critique of judgment. New York: Hafner Press.Google Scholar
Keil, F. (1979). Semantic and conceptual development: An ontological perspective. Cambridge, MA: Harvard University Press.CrossRefGoogle Scholar
Keil, F. (1983). On the emergence of semantic and conceptual distinctions. Journal of Experimental Psychology, 112, 357–385.CrossRefGoogle Scholar
Keil, F. C. (2007). Biology and beyond: Domain specificity in a broader developmental context. Human Development, 50, 31–38.CrossRefGoogle Scholar
Keil, F. C., & Wilson, R. A. (2000). Explanation and cognition. Cambridge, MA: MIT Press.Google Scholar
Kelley, A. E. (1999). Neural integrative activities of nucleus accumbens subregions in relation to learning and motivation. Psychobiology, 27, 198–213.Google Scholar
Keltner, D., & Haidt, J. (2003). Approaching awe, a moral, spiritual, and aesthetic emotion. Cognition and Emotion, 17, 297–314.CrossRefGoogle ScholarPubMed
Kempermann, G. (2006). Adult neurogenesis. Oxford: Oxford University Press.CrossRefGoogle Scholar
Kempermann, G., Kuhn, H. G., & Gage, F. H. (1999). Experience-induced neurogenesis in the senescent dentate gyrus. Journal of Neuroscience, 18, 3206–3212.CrossRefGoogle Scholar
Kempermann, G., Wiskott, L., & Gage, F. H. (2004). Functional significance of adult neurogenesis. Current Opinion in Neurobiology, 14, 186–191.CrossRefGoogle ScholarPubMed
Kierkegaard, S. (1844/1980). Fear and trembling and the sickness unto death. Princeton, NJ: Princeton University Press.Google Scholar
King, B. J. (2007). Evolving God. New York: Doubleday.Google Scholar
King-Hele, D. (1986). Erasmus Darwin and the romantic poets. New York: Macmillan.CrossRefGoogle Scholar
Kirk, G. S., & Raven, J. E. (1957). The presocratic philosophers. Cambridge: Cambridge University Press.Google Scholar
Kitcher, P. (1990). Kant's transcendental psychology. Oxford: Oxford University Press.Google Scholar
Kitcher, P. (1993). The advancement of science. Oxford: Oxford University Press.Google Scholar
Kitcher, P. (1996). The lives to come. New York: Simon & Schuster.Google Scholar
Kitcher, P. (2007). Living with Darwin. Oxford: Oxford University Press.Google Scholar
Klintsova, A. Y., & Greenough, W. T. (1999). Synaptic plasticity in cortical systems. Current Opinion in Neurobiology, 9, 203–208.CrossRefGoogle ScholarPubMed
Knowlton, B., Mangels, J., & Squire, L. (1996). A neostriatal habit learning system in humans. Science, 273, 1399–1402.CrossRefGoogle ScholarPubMed
Kornblith, H. (1993). Inductive inference and its natural ground. Cambridge, MA: MIT Press.Google Scholar
Kosslyn, S. M., Alpert, N. M., & Thompson, W. L. (1993). Visual mental imagery and visual perception: PET studies. In Functional MRI of the brain. Arlington, VA: Society for Magnetic Resonance Imaging, pp. 183–190.Google Scholar
Koyre, A. (1961). Renaissance thought: The classic, scholastic, and humanist strains. New York: HarperTorch.Google Scholar
Koyre, A. (1968). Metaphysics and measurement: Essays in scientific revolution. Cambridge, MA: Harvard University Press.Google Scholar
Kripke, S. (1980). Naming and necessity. Cambridge, MA: Harvard University Press.Google Scholar
Krieckhaus, E. E. (1970). Innate recognition aids rats in sodium regulation. Journal of Comparative and Physiological Psychology, 73, 117–122.CrossRefGoogle ScholarPubMed
Krieckhaus, E. E., & Wolf, G. (1968). Interaction of innate mechanisms and latent learning. Journal of Comparative and Physiological Psychology, 65, 197–201.CrossRefGoogle ScholarPubMed
Kuhn, T. S. (1962). The structure of scientific revolution. Chicago: University of Chicago Press.Google Scholar
Kuhn, T. S. (1971). The relations between history and the history of science. Daedalus, 100, 271–304.Google Scholar
Kuhn, T. S. (2000). The road since structure. Chicago: University of Chicago Press.Google Scholar
Kuklick, B. (2001). A history of philosophy in America. Oxford: Oxford University Press.Google Scholar
Lahr, M. M., & Foley, R. (2004). Human evolution writ small. Nature, 431, 1043–1044.CrossRefGoogle Scholar
Lakoff, G., & Johnson, M. (1999). Philosophy in the flesh: The embodied mind and its challenge to Western thought. New York: Basic Books.Google Scholar
Lamarck, J. B. (1809/1984). Zoological philosophy (H. Elliot, Trans.). Chicago: University of Chicago Press.Google Scholar
Lamm, C., Batson, C. D., & Decety, J. (2007). The neural substrate of human empathy: Effects of perspective-taking and cognitive appraisal. Journal of Cognitive Neuroscience, 19, 42–58.CrossRefGoogle ScholarPubMed
Langer, S. (1937). Philosophy in a new key. Cambridge, MA: Harvard University Press.Google Scholar
Lashley, K. S. (1951). The problem of serial order in behavior. In Jeffress, L. A. (Ed.), Cerebral mechanisms in behavior. New York: Wiley and Sons, pp. 112–136.Google Scholar
Latour, B. (1999). Pandora's hope. Cambridge, MA: Harvard University Press.Google Scholar
Laudan, L. (1977). Progress and its problems. Berkeley: University of California Press.Google Scholar
Lawick-Goodall, J. V. (1971). In the shadow of man. London: Collins.Google Scholar
Lawrence, C., & Shapin, S. (Eds.). (1998). Science incarnate. Chicago: University of Chicago Press.Google Scholar
Lawson, E. T., & McCauley, R. N. (1990). Rethinking religion: Connecting cognition and culture. Cambridge: Cambridge University Press.Google Scholar
Leakey, L. S. B. (1934/1954). Adam's ancestors. New York: HarperTorch.Google Scholar
Leakey, R. F., & Lewin, R. (1977). Origins. New York: E. P. Dutton.Google Scholar
Doux, J. E. (1996). The emotional brain. New York: Simon & Schuster.Google Scholar
Lehrman, D. (1958). Induction of broodiness by participation in courtship and nest-building in the ring dove. Journal of Comparative Physiology and Psychology, 51, 32–36.CrossRefGoogle Scholar
Lennox, J. G. (2001). Aristotle's philosophy of biology. Cambridge: Cambridge University Press.Google Scholar
Leslie, A. M. (1987). Pretense and representation: The origins of “theory of mind.”Psychological Review, 94, 412–426.CrossRefGoogle Scholar
Levinson, S. (1996). Language and space. Annual Review of Anthropology, 25, 353–382.CrossRefGoogle Scholar
Levinson, S. (2003). Space in language and cognition. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Levinson, S. (2006). Cognition at the heart of human interaction. Discourse Studies, 8, 85–93.CrossRefGoogle Scholar
Levinson, S., & Jaisson, P. (Eds.). (2006). Evolution and culture. Cambridge, MA: MIT Press.Google Scholar
Lewis, B. (2004). What went wrong?Oxford: Oxford University Press.Google Scholar
Lim, M. M., Bielsky, I. F., & Young, L. J. (2005). Neuropeptides and the social brain. International Journal of Developmental Neuroscience, 23, 235–243.CrossRefGoogle ScholarPubMed
Linas, R. R. (2001). I of the vortex. Cambridge, MA: MIT Press.Google Scholar
Linnaeus, C. (1735). Systema naturae. Leiden: Haak.Google Scholar
Locke, J. (1692–1704/1955). A letter concerning toleration. Indianapolis: Bobbs-Merrill.Google Scholar
Loewenstein, G. (1994). The psychology of curiosity. Psychological Bulletin, 116, 75–98.CrossRefGoogle Scholar
Loewenstein, G. (2006). The pleasures and pains of information. Science, 312, 704–706.CrossRefGoogle Scholar
Lovejoy, A. O. (1936/1978). The great chain of being. Cambridge, MA: Harvard University Press.Google Scholar
Lovejoy, A. O. (1955). Essays in the history of ideas. New York: George Braziller.Google Scholar
Lovejoy, C. O. (1978). The origin of man. Science, 211, 341–350.CrossRefGoogle Scholar
Lowith, K. (1949). Meaning in history. Chicago: University of Chicago Press.Google Scholar
Lukas, J. (1968/1997). Historical consciousness. New Brunswick, NJ: Transactions.Google Scholar
Lupien, S. J., Maheu, F., Tu, M., Fiocco, A., & Schramek, T. E. (2007). The effects of stress and stress hormones on human cognition: Implications for the field of brain and cognition. Brain and Cognition, 65, 209–237.CrossRefGoogle ScholarPubMed
Lupien, S. J., & McEwen, B. S. (1997). The acute effects of corticosteroids on cognition: Integration of animal and human models studies. Brain Research, 24, 1–27.CrossRefGoogle Scholar
Lutz, A., Greischar, L. L., Rawling, N. B., Ricard, M., & Davidson, R. J. (2004). Long-term meditators self-induce high-amplitude gamma synchrony during mental practice. Proceedings of the National Academy of Sciences, 101, 16369–16373.CrossRefGoogle ScholarPubMed
Luzatti, F., Marchis, S., Fasolo, A., & Peretto, P. (2006). Neurogenesis in the caudate nucleus of the adult rabbit. Journal of Neuroscience, 28, 609–621.CrossRefGoogle Scholar
Lynch, M. P. (1998). Truth in context. Cambridge, MA: MIT Press.Google Scholar
Machiavelli, N. (1525/1988). Florentine histories. (Banfield, L. & Mansfield, Jr. H. C., Trans.). Princeton, NJ: Princeton University Press.Google Scholar
Maess, B., Koelsch, S., Gunter, T. C., & Frederici, A. D. (2001). Musical syntax is processed by Broca's area. Nature Neuroscience, 4, 540–545.CrossRefGoogle ScholarPubMed
Malinowski, B. (1948). Magic, science and religion. New York: Doubleday.Google Scholar
Malthus, T. R. (1798/1970). An essay on the principle of population. Baltimore: Penguin Books.Google Scholar
Mandler, J. M. (1992). How to build a baby: II, Conceptual primitives. Psychological Review, 99, 587–603.CrossRefGoogle ScholarPubMed
Mandler, J. M. (2004). The foundations of mind. Oxford: Oxford University Press.Google Scholar
Margolis, J. (2002). Reinventing pragmatism: American philosophy at the end of the twentieth century. Ithaca, NY: Cornell University Press.Google Scholar
Marler, P. (1961). The logical analysis of animal communication. Journal of Theoretical Biology, 1, 295–317.CrossRefGoogle ScholarPubMed
Marler, P. (2000). On innateness. In M. D. Hauser & M. Kinishi (Eds.), The design of animal communication. Cambridge, MA: MIT Press, pp. 293–318.Google Scholar
Martin, A., (2007). The representation of object concepts in the brain. Annual Review of Psychology, 58, 25–45.CrossRefGoogle Scholar
Martin, A., & Caramazza, A. (2003). Neuropsychological and neurimaging perspectives on conceptual knowledge. Cognitive Neuroscience, 30, 195–212.Google Scholar
Martin, A., & Weisberg, J. (2003). Neural foundations for understanding social and mechanical concepts. Cognitive Neuropsychology, 20 (3/4/5/6), 575–587.CrossRefGoogle ScholarPubMed
Martin, A., Ungerleider, L. G., & Haxby, J. V. (2000). Category specificity and the brain. In Gazzaniga, M. S. (Ed.), The new cognitive neurosciences. Cambridge, MA: MIT Press, pp. 1023–1036.Google Scholar
Matlock, T. (2004). Fictive motion as cognitive stimulation. Memory & Cognition, 32, 1389–1400.CrossRefGoogle Scholar
Mayr, E. (1942/1982). Systemics and the origin of species. New York: Columbia University Press.Google Scholar
Mayr, E. (1963). Animal species and evolution. Cambridge, MA: Harvard University Press.CrossRefGoogle Scholar
Mayr, E. (1991). One long argument. Cambridge, MA: Harvard University Press.Google Scholar
McCauley, R. N. (2000). The naturalness of religion and the unnaturalness of science. In F. C. Keil & R. A. Wilson (Eds.), Explanation and cognition. Cambridge, MA: MIT Press, pp. 61–85.Google Scholar
McCauley, R. N., & Lawson, E. T. (2002). Bringing ritual to mind. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
McGaugh, J. L. (2000). Memory: A century of consolidation. Science, 287, 248–251.CrossRefGoogle ScholarPubMed
McGaugh, J. L. (2003). Memory and emotion. New York: Columbia University Press.Google Scholar
McGinn, C. (1997/1999). Ethics, evolution and fiction. Oxford: Oxford University Press.Google Scholar
Mead, G. H. (1934/1972). Mind, self and society. Chicago: University of Chicago Press.Google Scholar
Mead, M. (1928). Coming of age in Samoa. New York: W. Morrow.Google Scholar
Mead, M. (1964). Mind in cultural evolution. New Haven, CT: Yale University Press.Google Scholar
Medin, D. L., & Atran, S. (1999). Folkbiology. Cambridge, MA: MIT Press.Google Scholar
Medin, D. L., & Atran, S. (2004). The naïve mind: Biological categorization and reasoning in development and across culture. Psychological Reviews, 111, 960–983.CrossRefGoogle Scholar
Medvedev, R. (1989). Let history judge. New York: Columbia University Press.Google Scholar
Mele, A. R. (2003). Motivation and agency. Oxford: Oxford University Press.CrossRefGoogle Scholar
Mellars, P. (1996). Grooming, gossip and the evolution of language. London: Faber & Faber.Google Scholar
Mellars, P. (2006). Why did modern human populations disperse from Africa ca. 60,000 years ago?Proceedings of the National Academy of Sciences, 103, 9381–9386.CrossRefGoogle ScholarPubMed
Meltzoff, A. N. (2004). The case for developmental cognitive science: Theories of people and things. In G. Bremmer & A. Slater (Eds.), Theories of infant development. Oxford, England: Blackwell, pp. 145–173.Google Scholar
Meltzoff, A. N. (2007). “Like me”: A foundation for social cognition. Developmental Science, 10, 126–134.CrossRefGoogle Scholar
Meltzoff, A. N., & Moore, M. K. (1977). Imitation of facial and manual gestures by human neonates. Science, 198, 75–78.CrossRefGoogle ScholarPubMed
Menand, L. (2001). The metaphysical club. New York: Farrar, Straus & Giroux.Google Scholar
Merleau-Ponty, M. (1942/1967). The structure of behavior. Boston: Beacon Press.Google Scholar
Midgley, M. (1979/1995). Beast and man. London: Routledge.Google Scholar
Mill, J. S. (1843/1873). A system of logic. London: Longmans, Green, Rader & Dyer.Google Scholar
Ming, G. L., & Song, H. (2005). Adult neurogenesis in the mammalian central nervous system. Annual Review of Neuroscience, 28, 223–250.CrossRefGoogle ScholarPubMed
Mishkin, M., Malamut, B., & Bachevalier, J. (1984). Memories and habits: Two neural systems. In Lynch, G., McGaugh, J. L., & Weinberger, N. M. (Eds.), Neurobiology of learning and memory. New York: Gulliford, pp. 65–77.Google Scholar
Mishkin, M., Suzuki, W. A., Gadian, D. G., & Vargha-Khadem, F. (1997). Hierarchical organization of cognitive memory. Philosophical Transactions of the Royal Society of London, 352, 1461–1467.CrossRefGoogle ScholarPubMed
Mistlberger, R. E. (1994). Circadian food anticipatory activity: Formal models and physiological mechanisms. Neuroscience & Biobehavioral Reviews, 18, 171–195.CrossRefGoogle ScholarPubMed
Mithen, S. (1996). The prehistory of the mind: The cognitive origins of art and science. London: Thames & Hudson.Google Scholar
Mithen, S. (2006). The singing Neanderthal. Cambridge, MA: Harvard University Press.Google Scholar
Modahl, C., Green, L., Fein, D., Morris, M., Waterhouse, L., Feinstein, C., et al. (1998). Plasma oxytocin levels in autistic children. Biological Psychiatry, 43, 270–277.CrossRefGoogle ScholarPubMed
Moll, J., Oliveira-Souza, R., & Eslinger, P. J. (2003). Morals and the human brain: A working model. NeuroReport, 14, 299–305.CrossRefGoogle ScholarPubMed
Moore, J. A. (1993). Science as a way of knowing. Cambridge, MA: Harvard University Press.Google Scholar
Moore-Ede, M. C., Sulzman, F. M., & Fuller, C. A. (1992). The clocks that time us. Cambridge, MA: Harvard University Press.Google Scholar
Moreno, J. D. (2003). Neuroethics: An agenda for neuroscience and society. Nature Reviews Neuroscience, 4, 149–153.CrossRefGoogle ScholarPubMed
Muir, J. (1912/1962). The Yosemite. New York: Doubleday.Google Scholar
Muramoto, O. (2004). The role of the medial prefrontal cortex in human religious activity. Medical Hypotheses, 62, 479–485.CrossRefGoogle ScholarPubMed
Murphy, G. L. (2002). The big book of concepts. Cambridge, MA: MIT Press.Google Scholar
Nash, R. (1967). Wilderness and the American mind. New Haven, CT: Yale University Press.Google Scholar
Nelissen, K., Luppino, G., Vanduffel, W., Rizzolatti, G., & Orban, G. A. (2005). Observing others: Multiple action representation in the frontal lobe. Science, 310, 332–336.CrossRefGoogle ScholarPubMed
Nelson, R. J. (1995). An introduction to behavioral endocrinology. Sunderland, MA: Sinauer Associates.Google Scholar
Neville, R. C. (1974). The cosmology of freedom. New Haven, CT: Yale University Press.Google Scholar
Neville, R. C. (1992). Highroad around modernism. Albany: State University of New York Press.Google Scholar
Nichols, N., & Stitch, S. P. (2003). Mindreading. Oxford: Oxford University Press.CrossRefGoogle Scholar
Nunes, C. R., Pelz, K. M., Muecke, E. M., Holekamp, K. E., & Zucker, I. (2006). Plasma glucocorticoid concentrations and body mass in ground squirrels: Seasonal variation and circannual organization. General and Comparative Endocrinology, 146, 136–163.CrossRefGoogle ScholarPubMed
Nussbaum, M. C. (1997). Cultivating humanity. Cambridge, MA: Harvard University Press.Google Scholar
Nussbaum, M. C. (2004). Hiding from humanity. Princeton, NJ: Princeton University Press.Google Scholar
Nye, M. J. (1996). Before big science. Cambridge, MA: Harvard University Press.Google Scholar
Oakes, G. (1988). Weber and Rickert. Cambridge, MA: MIT Press.Google Scholar
Oelschlaeger, M. (1991). The idea of wilderness. New Haven, CT: Yale University Press.Google Scholar
Okamoto-Barth, S., Call, J., & Tomasello, M. (2007). Great apes' understanding of other individuals' line of sight. Psychological Science, 18, 462–468.CrossRefGoogle ScholarPubMed
Opfer, J. E., & Gelman, S. A. (2001). Children's and adult's models for predicting teleological action: The development of a biology-based model. Child Development, 72, 1367–1381.CrossRefGoogle Scholar
Otto, R. (1923/1975). The idea of the holy. Oxford: Oxford University Press.Google Scholar
Paabo, S. (2001). The human genome and our view of ourselves. Science, 291, 1219–1220.CrossRefGoogle ScholarPubMed
Panksepp, J. (1998). Affective neuroscience: The foundations of human and animal emotions. New York: Oxford University.Google Scholar
Parrott, W. G., & Schulkin, J. (1993). Neuropsychology and the cognitive nature of the emotions. Cognition and Emotion, 7, 43–59.CrossRefGoogle Scholar
Paton, J. A., & Nottebohm, F. N. (1984). Neurons generated in the adult brain are recruited into functional circuits. Science, 225, 1046–1048.CrossRefGoogle ScholarPubMed
Pauly, P. J. (1987). Controlling life. Oxford: Oxford University Press.Google Scholar
Paus, T. (2001). Primate anterior cingulate cortex: Where motor control, drive and cognition interface. Nature Reviews Neuroscience, 2, 417–424.CrossRefGoogle ScholarPubMed
Peirce, C. S. (1878). Deduction, induction and hypothesis. Popular Science Monthly, 13, 470–82.Google Scholar
Peirce, C. S. (1892). The architecture of theories. The Monist, 1, 61–76.Google Scholar
Peirce, C. S. (1893/1992). Evolutionary love. In N. Houser & C. Kloesel (Eds.), The essential Peirce: Vol. 1. Bloomington: Indiana University Press, pp. 352–362.Google Scholar
Peirce, C. S. (1893/1936). Religion and science: Collected works (E. Hartshorne & P. Weiss, Eds.). Cambridge, MA: Harvard University Press.Google Scholar
Peirce, C. S. (1899/1992). Reasoning and the logic of things (K. L. Ketner & H. Putnam, Eds.). Cambridge, MA: Harvard University Press.Google Scholar
Peirce, C. S. (1908/1998). A neglected argument for the reality of god. In N. Houser et al. (Eds.), The essential Peirce: Vol. 2. Bloomington: Indiana University Press, pp. 434–451.Google Scholar
Perani, D., Cappa, S. F., Bettinardi, V., Bressi, S., Gorno-Tempini, M.,Matarrese, M., et al. (1995). Different neural systems for the recognition of animals and man-made tools. Neuroreport, 6, 1636–1641.CrossRefGoogle ScholarPubMed
Perrett, D., Harries, M., Bevan, R., Thomas, S., Benson, P., Mistlin, A., et al. (1989). Frameworks of analysis for the neural representation of animate objects and actions. Journal of Experimental Biology, 146, 87–113.Google ScholarPubMed
Piaget, J. (1954). The construction of reality in the child. New York: Basic Books.CrossRefGoogle Scholar
Piaget, J. (1971/1975). Biology and knowledge. Chicago: University of Chicago Press.Google Scholar
Pinker, S. (1994). The language instinct. New York: William Morrow.CrossRefGoogle Scholar
Pinker, S. (1997). How the mind works. New York: W. W. Norton.Google Scholar
Pinker, S. (2007). The stuff of thought. New York: Viking.Google Scholar
Plato, (1941). The republic. Oxford: Oxford University Press.Google Scholar
Plotkin, H. (1993). Darwin machines. Cambridge, MA: Harvard University Press.Google Scholar
Polanyi, M. (1946/1964). Science, faith and society. Chicago: University of Chicago Press.Google Scholar
Premack, D. (1990). The infant's theory of self-propelled objects. Cognition, 36, 1–16.CrossRefGoogle ScholarPubMed
Premack, D., & Premack, A. J. (1983). The mind of the ape. New York: W. W. Norton.Google Scholar
Prinz, J. J. (2004). Gut reactions. Oxford: Oxford University Press.Google Scholar
Prinz, J. J., & Barsalou, L. W. (2000). Steering a course for embodied representation. In E. Dietrich & A. B. Markman (Eds.), Cognitive dynamics: Conceptual change in humans and machines. Mahwah, NJ: Lawrence Erlbaum, pp. 51–77.Google Scholar
Pulvermuller, F., Shtyrov, Y., & Ilmoniemi, R. (2005). Brain signatures of meaning in action word recognition. Journal of Cognitive Neuroscience, 6, 884–892.CrossRefGoogle Scholar
Putnam, H. (1990). Realism with a human face. Cambridge, MA: Harvard University Press.Google Scholar
Quine, W. V. O. (1953/1961). From a logical point of view. New York: HarperTorch.Google Scholar
Quine, W. V. O. (1969). Epistemology naturalized. In Ontological relativity and other essays. New York: Columbia University Press, pp. 69–90.Google Scholar
Rakison, D. H., & Dubois, D. P. (2001). Developmental origins of the animate-inanimate distinction. Psychological Bulletin, 127, 209–228.CrossRefGoogle Scholar
Rakison, D. H., & Oakes, L. M. (2003). Early category and concept development. Oxford: Oxford University Press.Google Scholar
Reschler, N. (2000). Nature and understanding. Oxford, England: Clarendon Press.Google Scholar
Rescorla, R. A. & Wagner, A. R. (1972). A theory of Pavlovian conditioning; Variations in the effectiveness of reinforcement and nonreinforcement. In A. H. Black & W. F. Prokasy, (Eds.), Classical conditioning. New York: Appleton-Century-Crofts, pp. 64–99.Google Scholar
Richards, R. J. (1992/1995). The meaning of evolution. Chicago: University of Chicago Press.CrossRefGoogle Scholar
Richter, C. P. (1943). Total self-regulatory functions in animals and man. New York: Harvey Lecture Series.Google Scholar
Richter, C. P. (1965/1979). Biological clocks in medicine and psychiatry. Springfield, IL: Charles C. Thomas.Google Scholar
Rickert, H. (1929/1986). The limits of concept formation in natural science (G. Oakes, Trans.). Cambridge: Cambridge University Press.Google Scholar
Rizzolatti, G., & Luppino, G. (2001). The cortical motor system. Neuron, 31, 889–901.CrossRefGoogle ScholarPubMed
Rohrer, T. (2001). Pragmatism, ideology and embodiment: William James and the philosophical foundations of cognitive linguistics. In Language and ideology: Cognitive theoretical approaches. Amsterdam: John Benjamins, pp. 49–81.CrossRefGoogle Scholar
Rolls, E. T., & Treves, A. (1998). Neural networks and brain function. New York: Oxford University Press.Google Scholar
Romer, A. S. (1968). Vertebrate paleontology. Chicago: University of Chicago Press.Google Scholar
Roozendaal, B. (2000). Glucocorticoids and the regulation of memory consolidation. Psychoneuroendocrinology, 25, 213–238.CrossRefGoogle ScholarPubMed
Rorty, R. (1999). Philosophy and social hope. New York: Penguin Books.Google Scholar
Rose, S. (1998). Lifelines. Oxford: Oxford University Press.Google Scholar
Rosen, J. B., & Schulkin, J. (1998). From normal fear to pathological anxiety. Psychological Review, 105, 325–350.CrossRefGoogle ScholarPubMed
Rosenberg, A. (2006). Darwinian reductionism. Chicago: University of Chicago Press.CrossRefGoogle Scholar
Rosenwasser, A. M. (2003). Neurobiology of the mammalian circadian system: Oscillators, pacemakers and pathways. In S. J. Fluharty & H. J. Grill (Eds.), Progress in psychobiology and physiological psychology, Vol. 18. New York: Elsevier, pp. 1–38.Google Scholar
Rosenwasser, A. M., Schulkin, J., & Adler, N. T. (1988). Anticipatory appetitive behavior of adrenalectomized rats under circadian salt-access schedules. Animal Learning and Behavior, 16, 324–329.CrossRefGoogle Scholar
Rosenzweig, M. R. (1984). Experience, memory and the brain. American Psychology, 39, 365–376.CrossRefGoogle Scholar
Royce, J. (1912/1940). The sources of religious insight. New York: Charles Scribner's Sons.Google Scholar
Rozin, P. (1976). The evolution of intelligence and access to the cognitive unconscious. In J. Sprague & A. N. Epstein (Eds.), Progress in psychobiology and physiological psychology. New York: Academic Press, pp. 245–280.Google Scholar
Rozin, P. (1998). Evolution and development of brains and cultures. In M. S. Gazzaniga & J. S. Altman (Eds.), Brain and mind: Evolutionary perspectives. France: Human Frontiers Sciences Program, pp. 111–125.Google Scholar
Rozin, P. (2005). The meaning of natural: Process more important than content. Psychological Science, 16, 652–658.CrossRefGoogle ScholarPubMed
Rozin, P., & Fallon, A. E. (1987). A perspective on disgust. Psychological Review, 94, 23–41.CrossRefGoogle ScholarPubMed
Rozin, P., & Schulkin, J. (1990). Food selection. In Stricker, E. M. (Ed.), Handbook of behavioral biology, Volume 10: Food and fluid intake. New York: Plenum Press, pp. 297–328.Google Scholar
Rubin, D. C. (2005). A basic-systems approach to autobiographical memory. Current Directions in Psychological Science, 14, 79–83.CrossRefGoogle Scholar
Ruby, N. F., Dark, J., Burns, D. E., Heller, H. C., & Zucker, I. (2002). The suprachiasmatic nucleus is essential for circadian body temperature rhythms in hibernating ground squirrels. Journal of Neuroscience, 22, 357–364.CrossRefGoogle ScholarPubMed
Ruby, P., & Decety, J. (2001). Effects of subjective perspective during simulation of action: A PET investigation of agency. Nature Neuroscience, 4, 546–550.CrossRefGoogle Scholar
Rue, L. (2005). Religion is not about God. New Brunswick, NJ: Rutgers University Press.Google Scholar
Runciman, W. G., Smith, J. M., & Dunbar, R. I. M. (Eds.). (1996). Evolution of social behaviour patterns in primates and man. New York: Oxford University Press.Google Scholar
Rusak, B., & Zucker, I. (1979). Neural regulation of circadian rhythms. Physiological Review, 59, 499–526.CrossRefGoogle ScholarPubMed
Ruse, M. (2006). Darwinism and its discontents. Cambridge: Cambridge University Press.Google Scholar
Russell, R. J., Murphy, N., Meyering, T. C., & Arbib, M. A. (2002). Neuroscience and the person. Rome: Vatican Observatory Foundation.Google Scholar
Ryle, G. (1949). The concept of mind. London: Hutchinson.Google Scholar
Sabini, J., & Schulkin, J. (1994). Biological realism and social constructivism. Journal for the Theory of Social Behavior, 224, 207–217.CrossRefGoogle Scholar
Sabini, J., & Silver, M. (1982). Moralities of everyday life. Oxford: Oxford University Press.Google Scholar
Sagoff, M. (1988). The economy of the earth. Cambridge: Cambridge University Press.Google Scholar
Santayana, G. (1932/1967). Character and opinion in the United States. New York: W. W. Norton.Google Scholar
Sapolsky, R. M. (1992). Stress: The aging brain and the mechanisms of neuron death. Cambridge, MA: MIT Press.Google Scholar
Sarokin, D., & Schulkin, J. (1994). Co-evolution of rights and environmental justice. The Environmentalist, 14, 121–129.CrossRefGoogle Scholar
Savan, D. (1981). Peirce's semiotic theory of emotion. In K. L. Ketner et al. (Eds.), Proceedings of the Charles S. Peirce Bicentennial International Congress. Lubbock: Texas Tech Press, pp. 319–333.Google Scholar
Saver, J. L., & Rabin, J. (1997). The neural substrates of religious experience. Journal of Neuropsychiatry, 9, 498–510.Google ScholarPubMed
Saxe, R., Tzelnic, T., & Carey, S. (2006). Five-month-old infants know humans are solid, like inanimate objects. Cognition, 101, B1–B8.CrossRefGoogle ScholarPubMed
Schacter, D. L. (1996). Searching for memory. New York: Basic Books.Google Scholar
Schacter, D. L., & Addis, D. R. (2007). The cognitive neuroscience of constructive memory: Remembering the past and imagining the future. Philosophical Transactions of the Royal Society, 362, 773–786.CrossRefGoogle ScholarPubMed
Schacter, D. L., & Tulving, E. (1994). Memory systems. Cambridge, MA: MIT Press.Google Scholar
Schama, S. (1995). Landscapes and memory. New York: Vintage Books.Google Scholar
Schelling, F. W. J. (1797/1988). Ideas for a philosophy of nature (E. E. Harris & P. Heath, Eds.). Cambridge: Cambridge University Press.Google Scholar
Scheler, M. (1928/1976). Man's place in nature (H. Meyerhoff, Trans.). New York: Noonday Press.Google Scholar
Schliermacher, F. (1799/1958). On religion. New York: Harper & Row.Google Scholar
Schmidt, L. A., & Schulkin, J. (Eds.). (1999). Extreme fear, shyness, and social phobia: Origins, biological mechanisms, and clinical outcome. (Series in Affective Science). New York: Oxford University Press.CrossRefGoogle Scholar
Schneider, H. W. (1946/1963). A history of American philosophy. New York: Columbia University Press.Google Scholar
Schulkin, J. (1991). Sodium hunger. Cambridge: Cambridge University Press.Google Scholar
Schulkin, J. (1992). The pursuit of inquiry. Albany: State University of New York Press.Google Scholar
Schulkin, J. (1996). The delicate balance. Lanham, MD: University Press of America.Google Scholar
Schulkin, J. (1999). The neuroendocrine regulation of behavior. Cambridge: Cambridge University Press.Google Scholar
Schulkin, J. (2000). Roots of social sensibility and neural function. Cambridge, MA: MIT Press.Google Scholar
Schulkin, J. (2003). Rethinking homeostasis. Cambridge, MA: MIT Press.Google Scholar
Schulkin, J. (2004). Bodily sensibility: Intelligent action. Oxford: Oxford University Press.Google Scholar
Schulkin, J., (2005). Curt Richter: A life in the laboratory. Baltimore: Johns Hopkins University Press.Google Scholar
Schulkin, J. (2007a). Effort: A neurobiological perspective on the will. New York: Lawrence Erlbaum Associates.Google Scholar
Schulkin, J. (2007b). Autism and the amygdala: An endocrine hypothesis. Brain and Cognition, 65, 87–99.CrossRefGoogle Scholar
Schultz, W. (2002). Getting formal with dopamine and reward. Neuron, 36, 241–263.CrossRefGoogle ScholarPubMed
Schultz, W. (2004). Neural coding of basic reward terms of animal learning, game theory, microeconomics and behavioral ecology. Current Opinion in Neurobiology, 14, 139–147.CrossRefGoogle Scholar
Schumpeter, J. A. (1934). Theory of economic development. Cambridge, MA: Harvard University Press.Google Scholar
Schutz, A. (1932/1967). The phenomenology of the social world (G. Walsh & F. Lehnert, Eds.). Chicago: Northwestern University Press.Google Scholar
Sellars, W. (1962). Science, perception, and reality. New York: Routledge & Kegan.Google Scholar
Sellars, W. (1968). Science and metaphysics. New York: Humanities Press.Google Scholar
Seneca, (1969). Letters from a stoic. New York: Penguin Classics.Google Scholar
Shade, P. (2001). Habits of hope: A pragmatic theory. Nashville, TN: Vanderbilt University Press.Google Scholar
Shapin, S. (1995). A social history of truth. Chicago: University of Chicago Press.Google Scholar
Shapin, S. (1996). The scientific revolution. Chicago: University of Chicago Press.CrossRefGoogle Scholar
Shapiro, L. A. (2004). The mind incarnate. Cambridge, MA: MIT Press.Google Scholar
Shelley, M. (1817/1976). Frankenstein. New York: Pyram.Google Scholar
Shook, J. R. (Ed.). (2003). Pragmatic naturalism and realism. Amherst, NY: Prometheus Press.
Shors, T. J., Micseages, C., Beylin, A., Zhao, M., Rydel, T., & Gould, E. (2001). Neurogenesis in the adult is involved in the formation of trace memories. Nature, 410, 372–375.CrossRefGoogle ScholarPubMed
Simon, H. A. (1962). The architecture of complexity. Proceedings of the American Philosophical Society, 106, 470–473.Google Scholar
Simon, H. A. (1982). Models of bounded rationality. Cambridge, MA: MIT Press.Google Scholar
Simon, H. A. (1990). Invariants of human behavior. Annual Review of Psychology, 41, 1–20.CrossRefGoogle ScholarPubMed
Simpson, G. (1961). Principles of animal taxonomy. New York: Columbia University Press.Google Scholar
Simpson, G. G. (1949). The meaning of evolution. New Haven, CT: Yale University Press.Google Scholar
Simpson, G. G. (1980). Splendid isolation. New Haven, CT: Yale University Press.Google Scholar
Skrbina, D. (2005). Panpsychism in the West. Cambridge, MA: MIT Press.Google Scholar
Smith, J. E. (1970). Themes in American philosophy. New York: HarperTorch/Harper & Row.Google Scholar
Smith, J. E. (1978). Purpose and thought. New Haven, CT: Yale University Press.Google Scholar
Smith, J. E. (1985). Experience in Peirce, James and Dewey. Monist, 68, 538–554.CrossRefGoogle Scholar
Smith, L., & Gasser, M. (2005). The development of embodied cognition: Six lessons from babies. Artificial Life, 11, 13–29.CrossRefGoogle ScholarPubMed
Smith, W. J. (1977). The behavior of communicating: An ethological approach. Cambridge, MA: Harvard University Press.Google Scholar
Snow, C. P. (1961). Science and government. Cambridge, MA: Harvard University Press.Google Scholar
Solomon, R. C. (2002). Spirituality for the skeptic. Oxford: Oxford University Press.CrossRefGoogle Scholar
Spelke, E. S., Phillips, A., & Woodward, A. L. (1995). Infants' knowledge of object motion and human action. In D. Sperber, D. Premack, & A. J. Premack (Eds.), Causal cognition: A multidisciplinary debate. Oxford, England: Clarendon Press, pp. 44–77.Google Scholar
Sperber, D. (1975). Rethinking symbolism. Cambridge: Cambridge University Press.Google Scholar
Sperber, D. (1985). On anthropological knowledge. Cambridge: Cambridge University Press.Google Scholar
Spezio, M. L., Adolphs, R., Hurley, R. S., & Piven, J. (2007). Analysis of face gaze in autism using “bubbles.”Neuropsychologia, 45, 144–151.CrossRefGoogle ScholarPubMed
Spinoza, B. (1668/1955). On the improvement of the understanding (R. H. M. Elwes, Ed.). New York: Dover Press.Google Scholar
Squire, L. R. (1987). Memory and brain. New York: Oxford University Press.Google Scholar
Squire, L. R. (2004). Memory systems of the brain: A brief history and current perspective. Neurobiology of Learning and Memory, 82, 171–177.CrossRefGoogle ScholarPubMed
Squire, L. R., & Zola, S. M. (1998). Episodic memory, semantic memory, and amnesia. Hippocampus, 8, 205–211.3.0.CO;2-I>CrossRefGoogle ScholarPubMed
Sterelny, K. (2000). The evolution of agency and other essays. Cambridge: Cambridge University Press.Google Scholar
Sterelny, K. (2003). Thought in a hostile world. New York: Blackwell.Google Scholar
Sterelny, K. (2004). Genes, memes and human history. Mind and Language, 19, 249–57.CrossRefGoogle Scholar
Sterling, P. (2004). Principles of allostasis. In Schulkin, J. (Ed.), Allostasis, homeostasis and the costs of physiological adaptation. Cambridge: Cambridge University Press, pp. 17–64.Google Scholar
Sterling, P., & Eyer, J. (1988). Allostasis: A new paradigm to explain arousal pathology. In Fisher, S. & Reason, J. (Eds.), Handbook of life stress: Cognition and health. New York: John Wiley & Sons, pp. 629–648.Google Scholar
Suddendorf, T., & Corballis, M. C. (1997). Mental time travel and the evolution of the human mind. Genetic Social and General Psychology Monographs, 123, 133–167.
Suddendorf, T., & Corballis, M. C. (2007). The evolution of foresight. Behavioral and Brain Sciences, 30, 299–313.CrossRefGoogle ScholarPubMed
Swanson, L. W. (2000). Cerebral hemisphere regulation of motivated behavior. Brain Research, 886, 113–164.CrossRefGoogle ScholarPubMed
Swanson, L. W. (2003). Brain architecture. Oxford: Oxford University Press.Google Scholar
Tattersall, I. (1993). The human odyssey: Four million years of human evolution. New York: Prentice Hall.Google Scholar
Taupin, P., & Gage, F. H. (2002). Adult neurogenesis and neural stem cell of the central nervous system. Journal of Neuroscience Research, 69, 745–749.CrossRefGoogle ScholarPubMed
Taylor, C. (2002). Varieties of religion today. Cambridge, MA: Harvard University Press.Google Scholar
Taylor, C. (2007). A secular age. Cambridge, MA: Harvard University Press.Google Scholar
Thomas, E. (2001). Empathy and consciousness. Journal of Consciousness Studies, 8, 1–35.Google Scholar
Thoreau, H. D. (1971). Great short works. New York: Harper & Row.Google Scholar
Thucydides, (1989). The Peloponnesian wars (T. Hobbes, Trans.). Chicago: University of Chicago Press.CrossRefGoogle Scholar
Tillich, P. (1951/1967). Systematic theology (Vols. 1–2). Chicago: University of Chicago Press.Google Scholar
Todes, D. P. (1989). Darwin without Malthus. Oxford: Oxford University Press.Google Scholar
Todes, D. P. (1997). Pavlov's physiology factory. History of Science Society, 88, 205–46.Google ScholarPubMed
Tolman, E. C. (1949). Purposive behavior in animals and men. Berkeley: University of California Press.Google Scholar
Tomasello, M., & Call, J. (1997). Primate cognition. Oxford: Oxford University Press.Google Scholar
Tomasello, M., & Carpenter, M. (2007). Shared intentionality. Developmental Science, 10, 121–125.CrossRefGoogle ScholarPubMed
Tomasello, M., Carpenter, M., Call, J., Behne, T., & Moll, H. (2004). Understanding and sharing intentions: The origins of cultural cognition. Behavioral and Brain Sciences, 28, 675–735.Google Scholar
Tomasello, M., Savage-Rumbaugh, E. S., & Kruger, A. C. (1993). Imitative learning of actions on objects by children, chimpanzees, and enculturated chimpanzees. Child Development, 64, 688–705.CrossRefGoogle ScholarPubMed
Tooby, J., & Cosmides, L. (1992). The psychological foundations of culture. In Barkow, J. H., Cosmides, L., & Tooby, J. (Eds.), The adaptive mind. New York: Oxford University Press, pp. 19–136.Google Scholar
Toulmin, S. (1977). Human understanding. Princeton, NJ: Princeton University Press.Google Scholar
Toulmin, S. (2001). Return to reason. Cambridge, MA: Harvard University Press.Google Scholar
Tulving, E. (1983/1993). Elements of episodic memory. Oxford, England: Clarendon Press.Google Scholar
Tulving, E. (2002). Episodic memory: From mind to brain. Annual Review of Psychology, 53, 1–25.CrossRefGoogle Scholar
Tulving, E., & Craik, F. I. M. (2000). The Oxford handbook of memory. Oxford: Oxford University Press.Google Scholar
Tzu, C. (1962). The writing of Chuang Tzu. New York: Dover Press.Google Scholar
Ullman, M. T. (2001). A neurocognitive perspective on language: The declarative procedural model. Nature Neuroscience, 9, 266–286.Google Scholar
Ullman, M. T. (2004). Is Broca's area part of a basal ganglia thalamocortical circuit?Cognition, 92, 231–270.CrossRefGoogle Scholar
Prang, H., Christie, B. R., Sejnowski, T. J., & Gage, F. H. (1999). Running enhances neurogenesis, learning and long-term potentiation in mice. Proceedings of the National Academy of Sciences, 96, 13427–13431.Google Scholar
Varela, F. J., Thompson, E., & Rosch, E. (1991). The embodied mind. Cognitive Science and Human Experience Series. Cambridge, MA: MIT Press.Google Scholar
Vico, G. (1744/1970). The new sciences. Ithaca, NY: Cornell University Press.Google Scholar
Wang, A. T., Dapretto, M., Hariri, A. R., Sigman, M., & Bookheimer, S. Y. (2004). Neural correlates of facial affect processing in children and adolescents with autism spectrum disorder. Journal of the American Academy of Child and Adolescent Psychiatry, 43, 481–490.CrossRefGoogle ScholarPubMed
Warrington, E. K., & Shallice, T. (1984). Category-specific semantic impairment. Brain, 107, 829–854.CrossRefGoogle Scholar
Waxman, S. R. (1999). The dubbing ceremony revisited: Object naming and categorization in infancy and early childhood. In D. L. Medin & S. Atran (Eds.), Folkbiology. Cambridge, MA: MIT Press, pp. 232–284.Google Scholar
Waxman, S. R. (2007). Folkbiological reasoning from a cross-cultural developmental perspective: Early essentialist notions are shaped by cultural beliefs. Developmental Psychology, 43, 294–308.CrossRefGoogle ScholarPubMed
Weber, M. (1905/1958). The Protestant ethic and the spirit of capitalism (T. Parkson, Trans.). New York: Scribner's.Google Scholar
Weber, M. (1947). The theory of social and economic organization (A. Henderston & T. Parsons, Trans.). New York: Free Press.Google Scholar
Wehr, T. A., Moul, D. E., Barbato, G., Giesen, H. A., Seidel, J. A., Barker, C., et al. (1993). Conservation of photoperiod-responsive mechanisms in humans. American Journal of Physiology, 265, 846–857.Google ScholarPubMed
Weidman, N. M. (1999). Constructing scientific psychology. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Weisberg, J., Turennout, M., & Martin, A. (2006). A neural system for learning about object function. Cerebral Cortex, 17, 513–521.CrossRefGoogle ScholarPubMed
Wells, G. (1999). Dialogic inquiry. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Wellman, H. (1990). The child's theory of mind. Cambridge, MA: MIT Press.Google Scholar
Weissman, D. (2000). A social ontology. New Haven, CT: Yale University Press.Google Scholar
Weissman, D. (2008). Styles of thought. Albany: State University of New York Press.Google Scholar
Wells, G. (1999). Dialogic inquiry: Toward a sociocultural practice and theory of education. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Wheatley, T., Milleville, S. C., & Martin, A. (2007). Understanding animate agents. Psychological Science, 18, 469–474.CrossRefGoogle ScholarPubMed
Wheeler, M. (2005). Reconstructing the cognitive world. Cambridge, MA: MIT Press.Google Scholar
Whitehead, A. N. (1927/1953). Symbolism. New York: Macmillan.Google Scholar
Whitehead, A. N. (1929/1957). The aims of education. New York: Free Press.Google Scholar
Whitehead, A. N. (1929/1958). The function of reason. Boston: Beacon Press.Google Scholar
Whitehead, A. N. (1933/1961). Adventures of ideas. New York: Free Press.Google Scholar
Whitehead, A. N. (1938/1967). Modes of thought. New York: Free Press.Google Scholar
Whiten, A., & Byrne, R. W. (Eds.). (1997). Machiavellian intelligence 11: Extensions and evaluations. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Williams, W. C. (1949/1969). Selected poems. New York: New Directions Press.Google Scholar
Wilson, E. O. (1994). Naturalist. Washington, DC: Island Press.Google Scholar
Wilson, E. O. (1995). Consilience: The unity of knowledge. New York: Alfred A. Knopf.Google Scholar
Wilson, M. (2002). Six views of embodied cognition. Psychonomic Bulletin and Review, 9, 625–636.CrossRefGoogle ScholarPubMed
Wilson, M., & Knoblich, G. (2005). The case for motor involvement in perceiving conspecifics. Psychological Bulletin, 131, 460–473.CrossRefGoogle ScholarPubMed
Wilson, R. A. (Ed.). (1999). Species: New interdisciplinary essays. Cambridge, MA: MIT Press.
Wilson, R. A. (2004). Boundaries of the mind. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Wilson, R. A. (2005). Genes and the agents of life. Cambridge: Cambridge University Press.Google Scholar
Wingfield, J. C. (2004). Allostatic load and life cycles: Implication for neuroendocrine control mechanisms. In Schulkin, J. (Ed.), Allostasis, homeostasis and the costs of physiological adaptation. Cambridge: Cambridge University Press, pp. 302–342.CrossRefGoogle Scholar
Wittgenstein, L. (1953/1968). Philosophical investigations. New York: Macmillan.Google Scholar
Wolff, P., & Medin, D. L. (2001). Measuring the evolution and devolution of folk-biological knowledge. In Maffi, L. (Ed.), On biocultural diversity: Linking language, knowledge, and the environment. Washington, DC: Smithsonian Institution, pp. 212–227.Google Scholar
Woodward, C. V. (1955/1966). The strange career of Jim Crow. New York: Oxford University Press.Google Scholar
Woodward, C. V. (1986). Thinking back. Baton Rouge: Louisiana State University.Google Scholar
Woodward, C. V. (1989). Future of the past. Oxford: Oxford University Press.Google Scholar
Worster, D. (1977/1991). Nature's economy. Cambridge: Cambridge University Press.Google Scholar
Wu, S., Jia, M., Ruan, Y., Liu, J., Guo, Y., Shuang, M., et al. (2005). Positive association of the oxytocin receptor gene (OXTR) with autism in the Chinese Han population. Biological Psychiatry, 58, 74–77.CrossRefGoogle ScholarPubMed
Wuerfel, J., Krischamoorthy, E. S., Brown, R. J., Lemieus, L., Koepp, M., Tebartz van Elst, L., et al. (2004). Religiosity is associated with hippocampal, but not amygdala volume in patients with refractory epilepsy. Journal of Neurology, Neurosurgery and Psychiatry, 75, 640–642.CrossRef
Young, A. W. (1998). Face and mind. Oxford: Oxford University Press.CrossRefGoogle Scholar
Zeki, S. (1999). Art and the brain. Journal of Consciousness Studies, 6, 76–96.Google Scholar
Zucker, I. (1988). Neuroendocrine substrates of circannual rhythms. In D. J. Kufpler, T. R. Monk, & J. D. Barchas (Eds.), Biological rhythms and mental disorders. New York: Guilford Press, pp. 219–251.Google Scholar

Save book to Kindle

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

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

Find out more about the Kindle Personal Document Service.

  • References
  • Jay Schulkin, Georgetown University, Washington DC
  • Book: Cognitive Adaptation
  • Online publication: 08 August 2009
  • Chapter DOI: https://doi.org/10.1017/CBO9780511499982.009
Available formats
×

Save book to Dropbox

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

  • References
  • Jay Schulkin, Georgetown University, Washington DC
  • Book: Cognitive Adaptation
  • Online publication: 08 August 2009
  • Chapter DOI: https://doi.org/10.1017/CBO9780511499982.009
Available formats
×

Save book to Google Drive

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

  • References
  • Jay Schulkin, Georgetown University, Washington DC
  • Book: Cognitive Adaptation
  • Online publication: 08 August 2009
  • Chapter DOI: https://doi.org/10.1017/CBO9780511499982.009
Available formats
×