Skip to main content Accessibility help
×
Hostname: page-component-7c8c6479df-995ml Total loading time: 0 Render date: 2024-03-28T14:31:52.948Z Has data issue: false hasContentIssue false

Philosophy of Neuroscience

Published online by Cambridge University Press:  08 February 2022

William Bechtel
Affiliation:
University of California, San Diego
Linus Ta-Lun Huang
Affiliation:
The University of Hong Kong

Summary

This Element provides a comprehensive introduction to philosophy of neuroscience. It covers such topics as how neuroscientists procure knowledge, including not just research techniques but the use of various model organisms. It presents examples of knowledge acquired in neuroscience that are then employed to discuss more philosophical topics such as the nature of explanations developed in neuroscience, the different conception of levels employed in discussions of neuroscience, and the invocation of representations in neuroscience explanations. The text emphasizes the importance of brain processes beyond those in the neocortex and then explores what makes processing in neocortex different. It consider the view that the nervous system consists of control mechanisms and considers arguments for hierarchical vs. heterarchical organization of control mechanisms. It concludes by considering implications of findings in neuroscience for how humans conceive of themselves and practices such as embracing norms.
Get access
Type
Element
Information
Online ISBN: 9781108946964
Publisher: Cambridge University Press
Print publication: 03 March 2022

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

Abrahamsen, A., Sheredos, B., & Bechtel, W. (2018). Explaining visually using mechanism diagrams. In Glennan, S. & Illari, P. (Eds.), The Routledge handbook of mechanisms (pp. 238254). London: Routledge.Google Scholar
Akins, K. (1996). On sensory systems and the “aboutness” of mental states. The Journal of Philosophy, 93(7), 337372.Google Scholar
Andersen, R. A., Essick, G. K., & Siegel, R. M. (1985). Encoding of spatial location by posterior parietal neurons. Science, 230, 456458.Google Scholar
Anderson, M. L. (2014). After phrenology: Neural reuse and the interactive brain. Cambridge, MA: MIT Press.CrossRefGoogle Scholar
Ankeny, R. A., & Leonelli, S. (2020). Model organisms. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Ardiel, E. L., & Rankin, C. H. (2010). An elegant mind: Learning and memory in Caenorhabditis elegans. Learning & Memory, 17(4), 191201. DOI:10.1101/lm.960510.CrossRefGoogle ScholarPubMed
Arendt, D., Tosches, M. A., & Marlow, H. (2016). From nerve net to nerve ring, nerve cord and brain–evolution of the nervous system. Nature Reviews Neuroscience, 17(1), 6172. DOI:10.1038/nrn.2015.15.Google Scholar
Arkett, S. A., Mackie, G. O., & Meech, R. W. (1988). Hair cell mechanoreception in the jellyfish Aglantha digitale. Journal of Experimental Biology, 135, 329342.CrossRefGoogle Scholar
Badre, D., & Nee, D. E. (2018). Frontal cortex and the hierarchical control of behavior. Trends in Cognitive Sciences, 22(2), 170188. DOI:10.1016/j.tics.2017.11.005.Google Scholar
Barabási, A.-L., & Bonabeau, E. (2003). Scale-free networks. Scientific American, 288(5), 60–69.Google Scholar
Bargmann, C. I., & Marder, E. (2013). From the connectome to brain function. Nature Methods, 10(6), 483490. DOI: 10.1038/Nmeth.2451.CrossRefGoogle ScholarPubMed
Barsalou, L. W. (2008). Cognitive and neural contributions to understanding the conceptual system. Current Directions in Psychological Science, 17(2), 9195. DOI: 10.1111/j.1467–8721.2008.00555.x.Google Scholar
Bechtel, W. (2008). Mental mechanisms: Philosophical perspectives on cognitive neuroscience. London: Routledge.Google Scholar
Bechtel, W. (2009). Molecules, systems, and behavior: Another view of memory consolidation. In Bickle, J. (Ed.), Oxford handbook of philosophy and neuroscience (pp. 1340). Oxford: Oxford University Press.CrossRefGoogle Scholar
Bechtel, W. (2016). Investigating neural representations: The tale of place cells. Synthese, 193(5), 12871321. DOI:10.1007/s11229-014-0480-8.CrossRefGoogle Scholar
Bechtel, W. (2021). Explaining features of fine-grained phenomena using abstract analyses of phenomena and mechanisms: Two examples from chronobiology. Synthese, 198(1), 123. DOI:10.1007/s11229-017-1469-x.CrossRefGoogle Scholar
Bechtel, W. (2019). Networks and dynamics: 21st century neuroscience. In Robins, S., Symons, J., & Calvo, P. (Eds.), The Routledge companion to philosophy of psychology (pp. 456470). London: Routledge.Google Scholar
Bechtel, W., & Abrahamsen, A. (2002). Connectionism and the mind: Parallel processing, dynamics, and evolution in networks (second ed.). Oxford: Blackwell.Google Scholar
Bechtel, W., & Abrahamsen, A. (2005). Explanation: A mechanist alternative. Studies in History and Philosophy of Biological and Biomedical Sciences, 36(2), 421441.Google Scholar
Bechtel, W., & Abrahamsen, A. (2010). Dynamic mechanistic explanation: Computational modeling of circadian rhythms as an exemplar for cognitive science. Studies in History and Philosophy of Science Part A, 41(3), 321333.Google Scholar
Bechtel, W., & Richardson, R. C. (1993/2010). Discovering complexity: Decomposition and localization as strategies in scientific research. Princeton, NJ: /Princeton, NJ:Princeton University Press (1993)/Cambridge, MA: MIT Press (2010).CrossRefGoogle Scholar
Berger, H. (1930). Über das Elektrenkephalogramm des Menschen. Zweite Mitteilung. Journal für Psychologie und Neurologie, 40, 160179.Google Scholar
Bickle, J. (2006). Reducing mind to molecular pathways: Explicating the reductionism implicit in current cellular and molecular neuroscience. Synthese, 151, 411434.CrossRefGoogle Scholar
Bjursten, L. M., Norrsell, K., & Norrsell, U. (1976). Behavioural repertory of cats without cerebral cortex from infancy. Experimental Brain Research, 25(2), 115130. DOI:10.1007/BF00234897.CrossRefGoogle ScholarPubMed
Bogacz, R., & Gurney, K. (2007). The basal ganglia and cortex implement optimal decision making between alternative actions. Neural Computation, 19(2), 442477. DOI:10.1162/neco.2007.19.2.442.Google Scholar
Briggman, K. L., Abarbanel, H. D. I., & Kristan, W. B. (2005). Optical imaging of neuronal populations during decision-making. Science, 307(5711), 896901. DOI:10.1126/Science.1103736.Google Scholar
Britten, K. H., Shadlen, M. N., Newsome, W. T., & Movshon, J. A. (1992). The analysis of visual motion: A comparison of neuronal and psychophysical performance. The Journal of Neuroscience, 12, 47454765.Google Scholar
Broca, P. (1861). Remarques sur le siége de la faculté du langage articulé, suivies d’une observation d’aphemie (perte de la parole). Bulletin de la Société Anatomique, 6, 343357.Google Scholar
Brodmann, K. (1909/1994). Localization in the cerebral cortex (L. J. Garvey, trans.). New York: Springer.Google Scholar
Brook, A. (2009). Introduction: Philosophy in and philosophy of cognitive science. Topics in Cognitive Science, 1(2), 216230. DOI:10.1111/j.1756-8765.2009.01014.x.CrossRefGoogle ScholarPubMed
Bruner, J. S., & Postman, L. (1949). On the perception of incongruity: A paradigm. The Journal of Personality, 18, 206223.Google Scholar
Buchwald, J. S., & Brown, K. A. (1973). Subcortical mechanisms of behavioral plasticity. In Maser, J. D. (Ed.), Efferent organization and the integration of behavior (pp. 99136). New York: Academic Press.CrossRefGoogle Scholar
Buckner, C., & Garson, J. (2019). Connectionism. In E. N. Zalta (Ed.), The Stanford encyclopedia of philosophy. Retrieved from https://plato.stanford.edu/archives/fall2019/entries/connectionism.Google Scholar
Burk, J. A., & Fadel, J. R. (Eds.). (2019). The orexin/hypocretin system: Functional roles and therapeutic potential. New York: Academic Press.Google Scholar
Burnston, D. C. (2016). A contextualist approach to functional localization in the brain. Biology & Philosophy, 31(4), 527550. DOI:10.1007/s10539-016-9526-2.Google Scholar
Buzsáki, G. (2006). Rhythms of the brain. Oxford: Oxford University Press.Google Scholar
Buzsáki, G. (2010). Neural syntax: cell assemblies, synapsembles, and readers. Neuron, 68(3), 362–385. doi:10.1016/j.neuron.2010.09.023Google Scholar
Carlson, R. W., & Crockett, M. J. (2018). The lateral prefrontal cortex and moral goal pursuit. Current Opinion in Psychology, 24, 7782. DOI:10.1016/j.copsyc.2018.09.007.Google Scholar
Chakravarthy, V. S., & Balasubramani, P. P. (2018). The basal ganglia system as an engine for exploration. In Chakravarthy, V. S. & Moustafa, A. A. (Eds.), Computational neuroscience models of the basal ganglia (pp. 59–96). New York: Springer. DOI:10.1007/978-981-10-8494-2_5.CrossRefGoogle Scholar
Chalfie, M., Sulston, J. E., White, J. G., Southgate, E., Thomson, J. N., & Brenner, S. (1985). The neural circuit for touch sensitivity in Caenorhabditis elegans. Journal of Neuroscience, 5(4), 956964.Google Scholar
Chemero, A. (2000). Anti-representationalism and the dynamical stance. Philosophy of Science, 67(4), 625647. DOI:10.1086/392858.Google Scholar
Chemero, A., & Silberstein, M. (2008). After the philosophy of mind: Replacing scholasticism with science. Philosophy of Science, 75(1), 127. DOI:10.1086/587820.CrossRefGoogle Scholar
Churchland, P. M. (1981). Eliminative materialism and propositional attitudes. The Journal of Philosophy, 78, 6790.Google Scholar
Churchland, P. S. (1986). Neurophilosophy: Toward a unified science of the mind-brain. Cambridge, MA: MIT Press/Bradford Books.Google Scholar
Cisek, P. (2019). Resynthesizing behavior through phylogenetic refinement. Attention, Perception, & Psychophysics, doi:10.3758/s13414-019-01760-1.Google Scholar
Cisek, P., & Kalaska, J. F. (2010). Neural mechanisms for interacting with a world full of action choices. Annual Review of Neuroscience, 33, 269298. DOI:10.1146/annurev.neuro.051508.135409.CrossRefGoogle ScholarPubMed
Clark, A. (2013). Whatever next? Predictive brains, situated agents, and the future of cognitive science. The Behavioral and Brain Sciences, 36(3), 181204. DOI:10.1017/S0140525X12000477.CrossRefGoogle ScholarPubMed
Clark, A. (2014). Mindware: An introduction to the philosophy of cognitive science (second ed.). New York: Oxford University Press.Google Scholar
Colaço, D. (2018). Rip it up and start again: The rejection of a characterization of a phenomenon. Studies in History and Philosophy of Science Part A, 72, 3240. DOI:10.1016/j.shpsa.2018.04.003.CrossRefGoogle ScholarPubMed
Colgin, L. L., Moser, E. I., & Moser, M.-B. (2008). Understanding memory through hippocampal remapping. Trends in Neurosciences, 31(9), 469477. DOI:10.1016/j.tins.2008.06.008.CrossRefGoogle ScholarPubMed
Corkin, S. (2013). Permanent present tense: The unforgettable life of the amnesic patient, H.M. New York: Basic Books.Google Scholar
Craver, C. F. (2003). The making of a memory mechanism. Journal of the History of Biology, 36, 153195.Google Scholar
Craver, C. F. (2006). When mechanistic models explain. Synthese, 153(3), 355376. DOI:10.1007/s11229-006-9097-x.Google Scholar
Craver, C. F. (2007). Explaining the brain: Mechanisms and the mosaic unity of neuroscience. New York: Oxford University Press.CrossRefGoogle Scholar
Craver, C. F. (2008). Physical law and mechanistic explanation in the Hodgkin and Huxley model of the action potential. Philosophy of Science, 75(5), 10221033. DOI:10.1086/594543.Google Scholar
Craver, C. F., & Darden, L. (2013). In search of mechanisms: Discoveries across the life sciences. Chicago: University of Chicago Press.Google Scholar
Craver, C. F., & Tabery, J. (2019). Mechanisms in science. In E. N. Zalta (Ed.), The Stanford excyclopedia of philosophy. Retrieved from https://plato.stanford.edu/archives/sum2019/entries/science-mechanisms/.Google Scholar
Crisp, K. M., & Mesce, K. A. (2006). Beyond the central pattern generator: Amine modulation of decision-making neural pathways descending from the brain of the medicinal leech. Journal of Experimental Biology, 209(9), 17461756. DOI:10.1242/jeb.02204.CrossRefGoogle ScholarPubMed
Cushing, H. (1909). A note upon the faradic stimulation of the post-central gyrus in conscious patients. Brain, 32, 4453.Google Scholar
Damasio, A. R. (1995). Descartes’ error. New York: G. P. Putnam.Google Scholar
Davis, Z. W., Muller, L., Martinez-Trujillo, J., Sejnowski, T., & Reynolds, J. H. (2020). Spontaneous travelling cortical waves gate perception in behaving primates. Nature, 587(7834), 432436. DOI:10.1038/s41586-020-2802-y.CrossRefGoogle ScholarPubMed
Dennett, D. C. (1991). Consciousness explained. New York: Little, Brown.Google Scholar
Diba, K., & Buzsáki, G. (2007). Forward and reverse hippocampal place-cell sequences during ripples. Nature Neuroscience, 10(10), 12411242. DOI:10.1038/nn1961.CrossRefGoogle ScholarPubMed
Egan, F. (2019). The nature and function of content in computational models. In Sprevak, M. & Colombo, M. (Eds.), The Routledge handbook of the computational mind (pp. 247–258). London: Routledge.Google Scholar
Eichenbaum, H. (2002). The cognitive neuroscience of memory: An introduction. Oxford: Oxford University Press.Google Scholar
Fodor, J. A., & Pylyshyn, Z. W. (1988). Connectionism and cognitive architecture: A critical analysis. Cognition, 28, 371.Google Scholar
Fowler, O. S. (1890). The illustrated self-instructor in phrenology and physiology. New York: Fowler and Wells.Google Scholar
Frankish, K. (2004). Mind and supermind. Cambridge: Cambridge University Press.Google Scholar
Friston, K. (2010). The free-energy principle: A unified brain theory? Nature Reviews Neuroscience, 11(2), 127138. DOI:10.1038/nrn2787.Google Scholar
Fuster, J. M., Bodner, M., & Kroger, J. K. (2000). Cross-modal and cross-temporal association in neurons of frontal cortex. Nature, 405(6784), 347351. DOI:10.1038/35012613.Google Scholar
Gall, F. J. (1812). Anatomie et physiologie du systême nerveaux et général, et du cerveau en particulier, avec des observations sur la possibilité de reconnoitre plusieurs dispositions intellectuelles et morales de l’homme et des animaux, par la configuration de leur têtes. Paris: F. Schoell.Google Scholar
Galvani, L. (1791). De viribus electricitatis in motu musculari commentarius. Bologna: Ex typographia Instituti Scientiarum.CrossRefGoogle Scholar
Gaudry, Q., Ruiz, N., Huang, T., Kristan, W. B. III, & Kristan, W. B. Jr. (2010). Behavioral choice across leech species: chacun à son goût. The Journal of Experimental Biology, 213(8), 13561365. DOI:10.1242/jeb.039495.Google Scholar
Gibson, J. J. (1979). The ecological approach to visual perception. Boston: Houghton Mifflin.Google Scholar
Goldbeter, A. (1995). A model for circadian oscillations in the Drosophila period protein (PER). Philosophical Transactions of the Royal Society B: Biological Sciences, 261(1362), 319324. DOI:10.1098/rspb.1995.0153.Google Scholar
Goldman-Rakic, P. S. (1995). Cellular basis of working memory. Neuron, 14(3), 477485. DOI:10.1016/0896-6273(95)90304-6.Google Scholar
Grillner, S., & El Manira, A. (2019). Current principles of motor control, with special reference to vertebrate locomotion. Physiological Reviews, 100(1), 271320. DOI:10.1152/physrev.00015.2019.Google Scholar
Gross, C. G., Rocha-Miranda, C. E., & Bender, D. B. (1972). Visual properties of neurons in inferotemporal cortex of the macaque. Journal of Neurophysiology, 35, 96111.CrossRefGoogle ScholarPubMed
Haken, H., Kelso, J. A. S., & Bunz, H. (1985). A theoretical model of phase transitions in human hand movements. Biological Cybernetics, 51(5), 347356.Google Scholar
Hardin, P. E., Hall, J. C., & Rosbash, M. (1990). Feedback of the Drosophila period gene product on circadian cycling of its messenger RNA levels. Nature, 343(6258), 536540.Google Scholar
Haselager, W. F. G., de Groot, A., & van Rappard, H. (2003). Representationalism vs. anti-representationalism: A debate for the sake of appearance. Philosophical Psychology, 16, 523.Google Scholar
Hazy, T. E., Frank, M. J., & O’Reilly, R. C. (2007). Towards an executive without a homunculus: Computational models of the prefrontal cortex/basal ganglia system. Philosophical Transactions of the Royal Society B: Biological Sciences, 362(1485), 16011613. DOI:10.1098/rstb.2007.2055.CrossRefGoogle ScholarPubMed
Hempel, C. G. (1965). Aspects of scientific explanation and other essays in the philosophy of science. New York: The Free Press.Google Scholar
Hendricks, J. C., Finn, S. M., Panckeri, K. A., et al. (2000). Rest in Drosophila is a sleep-like state. Neuron, 25(1), 129138.CrossRefGoogle ScholarPubMed
Hills, T. T. (2006). Animal foraging and the evolution of goal-directed cognition. Cognitive Science, 30(1), 341. DOI:10.1207/s15516709cog0000_50.Google Scholar
Hodgkin, A. L., & Huxley, A. F. (1952). A quantitative description of membrane current and its application to the conduction and excitation of nerve. Journal of Physiology, 117, 500544.CrossRefGoogle Scholar
Hubel, D. H., & Wiesel, T. N. (1959). Receptive fields of single neurones in the cat’s striate cortex. Journal of Physiology, 148, 574591.Google Scholar
Huebner, B. (2014). Macrocognition: A theory of distributed minds and collective intentionality. Oxford: Oxford University Press.Google Scholar
Huneman, P. (2010). Topological explanations and robustness in biological sciences. Synthese, 177(2), 213245. DOI:10.1007/s11229-010-9842-z.CrossRefGoogle Scholar
Huneman, P. (2018). Diversifying the picture of explanations in biological sciences: Ways of combining topology with mechanisms. Synthese, 195(1), 115146. DOI:10.1007/s11229-015-0808-z.CrossRefGoogle Scholar
Joiner, W. J. (2016). Unraveling the evolutionary determinants of skoizleep. Current Biology, 26(20), R1073R1087. DOI:10.1016/j.cub.2016.08.068.Google Scholar
Kanwisher, N., McDermott, J., & Chun, M. M. (1997). The fusiform face area: A module in human extrastriate cortex specialized for face perception. Journal of Neuroscience, 17(11), 43024311.Google Scholar
Keene, A. C., & Duboue, E. R. (2018). The origins and evolution of sleep. The Journal of Experimental Biology, 221(11), jeb159533. DOI:10.1242/jeb.159533.Google Scholar
Keijzer, F., van Duijn, M., & Lyon, P. (2013). What nervous systems do: Early evolution, input–output, and the skin brain thesis. Adaptive Behavior, 21(2), 6785. DOI:10.1177/1059712312465330.Google Scholar
Koizumi, O. (2016). Origin and evolution of the nervous system considered from the diffuse nervous system of Cnidarians. In Goffredo, S. & Dubinsky, Z. (Eds.), The Cnidaria, past, present and future: The world of Medusa and her sisters (pp. 7391). Cham: Springer International Publishing.Google Scholar
Konopka, R. J., & Benzer, S. (1971). Clock mutants of Drosophila melanogaster. Proceedings of the National Academy of Sciences of the United States of America, 89, 21122116.Google Scholar
Kosslyn, S. M. (1994). Image and brain: The resolution of the imagery debate. Cambridge, MA: MIT Press.Google Scholar
Leng, G. (2018). The heart of the brain: The hypothalamus and its hormones. Cambridge, MA: MIT Press.CrossRefGoogle Scholar
Levy, A. (2013). What was Hodgkin and Huxley’s achievement? The British Journal for the Philosophy of Science, 65(3), 469492. DOI:10.1093/bjps/axs043.Google Scholar
Levy, A., & Bechtel, W. (2013). Abstraction and the organization of mechanisms. Philosophy of Science, 80(2), 241261. DOI:10.1086/670300.Google Scholar
List, C. (2013). Social choice theory. In E. N. Zalta (Ed.), The Stanford encyclopedia of philosophy. Retrieved from https://plato.stanford.edu/archives/win2013/entries/social-choice.Google Scholar
Loftus, E. F. (1975). Leading questions and the eyewitness report. Cognitive Psychology, 7, 550572.CrossRefGoogle Scholar
Machamer, P., Darden, L., & Craver, C. F. (2000). Thinking about mechanisms. Philosophy of Science, 67, 125.Google Scholar
Mackie, G. O. (2004). Central neural circuitry in the jellyfish Aglantha – A model “simple nervous system.” Neurosignals, 13(1–2), 519. DOI:10.1159/000076155.CrossRefGoogle Scholar
Mackie, G. O., Meech, R. W., & Spencer, A. N. (2012). A new inhibitory pathway in the jellyfish Polyorchis penicillatus. Canadian Journal of Zoology, 90(2), 172181. DOI:10.1139/Z11-124.CrossRefGoogle Scholar
Marr, D. C. (1982). Vision: A computation investigation into the human representational system and processing of visual information. San Francisco: Freeman.Google Scholar
Maturana, H. R., & Varela, F. J. (1980). Autopoiesis: The organization of the living. In Maturana, H. R. & Varela, F. J. (Eds.), Autopoiesis and cognition: The realization of the living (pp. 73138). Dordrecht: Reidel.Google Scholar
Maxwell, J. C. (1868). On governors. Proceedings of the Royal Society of London, 16, 270283.Google Scholar
McCulloch, W. S. (1945). A heterarchy of values determined by the topology of nervous nets. The Bulletin of Mathematical Biophysics, 7(2), 8993. DOI:10.1007/BF02478457.Google Scholar
McGeer, V. (2015). Mind-making practices: The social infrastructure of self-knowing agency and responsibility. Philosophical Explorations, 18(2), 259281. DOI:10.1080/13869795.2015.1032331.Google Scholar
Miller, E. K., & Cohen, J. D. (2001). An integrative theory of prefrontal cortex function. Annual Review of Neuroscience, 24(1), 167202. DOI:10.1146/annurev.neuro.24.1.167.Google Scholar
Milner, A. D., & Goodale, M. G. (1995). The visual brain in action. Oxford: Oxford University Press.Google Scholar
Minsky, M. (1986). The society of mind. New York: Simon and Schuster.Google Scholar
Mishkin, M., Ungerleider, L. G., & Macko, K. A. (1983). Object vision and spatial vision: Two cortical pathways. Trends in Neurosciences, 6, 414417.Google Scholar
Moreno, A., & Mossio, M. (2015). Biological autonomy: A philosophical and theoretical inquiry. Dordrecht: Springer.Google Scholar
Mundale, J. (2001). Neuroanatomical foundations of cognition: Connecting the neuronal level with the study of higher brain areas. In Bechtel, W., Mandik, P., Mundale, J., & Stufflebeam, R. S. (Eds.), Philosophy and the neurosciences (pp. 3754). Oxford: Blackwell.Google Scholar
Nielsen, K. (2010). Representation and dynamics. Philosophical Psychology, 23(6), 759773. DOI:10.1080/09515089.2010.529045.Google Scholar
O’Keefe, J. A., & Conway, D. H. (1978). Hippocampal place units in the freely moving rat: Why they fire where they fire. Experimental Brain Research, 31(4), 573590. DOI:10.1007/bf00239813.Google ScholarPubMed
O’Keefe, J. A., & Nadel, L. (1978). The hippocampus as a cognitive map. Oxford: Oxford University Press.Google Scholar
O’Keefe, J. A., & Recce, M. L. (1993). Phase relationship between hippocampal place units and the EEG theta rhythm. Hippocampus, 3, 317330.Google Scholar
O’Reilly, R. C., Petrov, A. A., Cohen, J. D., Lebiere, C. J., Herd, S. A., & Kriete, T. (2014). How limited systematicity emerges: A computational cognitive neuroscience approach. In Calvo, P. & Symons, J. (Eds.), The architecture of cognition: Rethinking Fodor and Pylyshyn’s systematicity challenge (pp. 191225). Cambridge, MA: MIT Press.Google Scholar
Pattee, H. H. (1991). Measurement-control heterarchical networks in living systems. International Journal of General Systems, 18(3), 213221.Google Scholar
Penfield, W., & Rasmussen, T. (1950). The cerebral cortex in man: A clinical study of localization of function. New York: Macmillan.Google Scholar
Pitt, D. (2020). Mental representation. In Zalta, E. N. (Ed.), The Stanford encyclopedia of philosophy. Retrieved at https://plato.stanford.edu/archives/win2013/entries/mental-representation/.Google Scholar
Puhl, J. G., & Mesce, K. A. (2008). Dopamine activates the motor pattern for crawling in the medicinal leech. Journal of Neuroscience, 28(16), 41924200. DOI:10.1523/JNEUROSCI.0136-08.2008.Google Scholar
Quiroga, R. Q., Reddy, L., Kreiman, G., Koch, C., & Fried, I. (2005). Invariant visual representation by single neurons in the human brain. Nature, 435(7045), 11021107.CrossRefGoogle ScholarPubMed
Raichle, M. E., MacLeod, A. M., Snyder, A. Z., Powers, W. J., Gusnard, D. A., & Shulman, G. L. (2001). A default mode of brain function. Proceedings of the National Academy of Sciences of the United States of America, 98(2), 676682.Google Scholar
Roseberry, T. K., Lee, A. M., Lalive, A. L., Wilbrecht, L., Bonci, A., & Kreitzer, A. C. (2016). Cell-type-specific control of brainstem locomotor circuits by basal ganglia. Cell, 164(3), 526537. DOI:10.1016/j.cell.2015.12.037.Google Scholar
Rosen, R. (1972). Some relational cell models: The metabolism-repair systems. In Rosen, R. (Ed.), Foundations of mathematical biology (vol. 2, pp. 217–253). New York: Academic Press.Google Scholar
Rupert, R. D. (2011). Embodiment, consciousness, and the massively representational mind. Philosophical Topics, 39(1), 99120.Google Scholar
Satterlie, R. A. (2018). Jellyfish locomotion. In Oxford research encyclopedia, neuroscience. New York: Oxford University Press. DOI:10.1093/acrefore/9780190264086.013.147.Google Scholar
Schwitzgebel, E. (2019). Introspection. In Zalta, E. N. (Ed.), The Stanford encyclopedia of philosophy. Retrieved from https://plato.stanford.edu/archives/win2019/entries/introspection/.Google Scholar
Searle, J. R. (1980). Minds, brains, and programs. Behavioral and Brain Sciences, 3, 417424.Google Scholar
Sejnowski, T. J. (2018). The deep learning revolution. Cambridge, MA: MIT Press.Google Scholar
Sellars, W. (1956). Empiricism and the philosophy of mind. In Feigl, H. & Scriven, M. (Eds.), The foundations of science and the concepts of psychology and psychoanalysis (Minnesota studies in the philosophy of science, vol. 1, pp. 253329). Minneapolis: University of Minnesota Press.Google Scholar
Shagrir, O. (2010). Marr on computational-level theories. Philosophy of Science, 77(4), 477500.Google Scholar
Shaw, P. J., Cirelli, C., Greenspan, R. J., & Tononi, G. (2000). Correlates of sleep and waking in Drosophila melanogaster. Science, 287(5459), 1834–1837. DOI:10.1126/science.287.5459.1834.Google Scholar
Shepherd, G. M. (2016). Foundations of the neuron doctrine (twenty-fifth anniversary ed.). Oxford: Oxford University Press.Google Scholar
Sohn, J. W. (2015). Network of hypothalamic neurons that control appetite. BMB Reports, 48(4), 229233. DOI:10.5483/BMBRep.2015.48.4.272.Google Scholar
Sporns, O. (2010). Networks of the brain. Cambridge, MA: MIT Press.Google Scholar
Sporns, O. (2012). Discovering the human connectome. Cambridge, MA: MIT Press.Google Scholar
Sterling, P., & Laughlin, S. (2015). Principles of neural design. Cambridge, MA: MIT Press.Google Scholar
Takahashi, J. S. (2017). Transcriptional architecture of the mammalian circadian clock. Nature Reviews Genetics, 18(3), 164179. DOI:10.1038/nrg.2016.150.Google Scholar
Tolman, E. C. (1948). Cognitive maps in rats and men. Psychological Review, 55, 189208.Google Scholar
Tosches, M. A., & Arendt, D. (2013). The bilaterian forebrain: An evolutionary chimaera. Current Opinion in Neurobiology, 23(6), 10801089. DOI:10.1016/j.conb.2013.09.005.Google Scholar
Tulving, E. (1983). Elements of episodic memory. New York: Oxford University Press.Google Scholar
Valenstein, E. S. (2005). The war of the soups and the sparks: The discovery of neurotransmitters and the dispute over how nerves communicate. New York: Columbia University Press.CrossRefGoogle Scholar
van den Heuvel, M. P., & Sporns, O. (2011). Rich-club organization of the human connectome. The Journal of Neuroscience, 31(44), 1577515786. DOI:10.1523/jneurosci.3539-11.2011.Google Scholar
van Essen, D. C., & Gallant, J. L. (1994). Neural mechanisms of form and motion processing in the primate visual system. Neuron, 13(1), 110.Google Scholar
van Gelder, T. (1998). The dynamical hypothesis in cognitive science. Behavioral and Brain Sciences, 21, 615628.Google Scholar
Voigt, J. P., & Fink, H. (2015). Serotonin controlling feeding and satiety. Behavioural Brain Research, 277, 1431. DOI:10.1016/j.bbr.2014.08.065.Google Scholar
Watts, D., & Strogratz, S. (1998). Collective dynamics of small worlds. Nature, 393, 440442.CrossRefGoogle Scholar
Weber, M. (2005). Philosophy of experimental biology. Cambridge: Cambridge University Press.Google Scholar
Weber, M. (2008). Causes without mechanisms: Experimental regularities, physical laws, and neuroscientific explanation. Philosophy of Science, 75(5), 9951007. DOI:10.1086/594541.Google Scholar
Welsh, D. K., Takahashi, J. S., & Kay, S. A. (2010). Suprachiasmatic nucleus: Cell autonomy and network properties. Annual Review of Physiology, 72(1), 551577. DOI:10.1146/annurev-physiol-021909-135919.Google Scholar
White, J. G., Southgate, E., Thomson, J. N., & Brenner, S. (1986). The structure of the nervous system of the nematode Caenorhabditis elegans. Philosophical Transactions of the Royal Society B: Biological Sciences, 314(1165), 1340. DOI:10.1098/rstb.1986.0056.Google Scholar
Winfree, A. T. (1987). The timing of biological clocks. New York: W. H. Freeman.Google Scholar
Winning, J. (2020). Internal perspectivalism: The solution to generality problems about proper function and natural norms. Biology & Philosophy, 35(3), 33. DOI:10.1007/s10539-020-09749-z.CrossRefGoogle Scholar
Winning, J., & Bechtel, W. (2018). Rethinking causality in neural mechanisms: Constraints and control. Minds and Machines, 28(2), 287310.CrossRefGoogle Scholar
Woodward, J. (2019). Scientific explanation. In E. N. Zalta (Ed.), The Stanford encyclopedia of philosophy. Retrieved from https://plato.stanford.edu/archives/win2019/entries/scientific-explanation/.Google Scholar
Yamins, D. L. K., & DiCarlo, J. J. (2016). Using goal-driven deep learning models to understand sensory cortex. Nature Neuroscience, 19(3), 356365. DOI:10.1038/nn.4244.Google Scholar
Zeki, S. M. (1971). Cortical projections from two prestriate areas in the monkey. Brain Research, 34, 1935.Google Scholar

Save element to Kindle

To save this element 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.

Philosophy of Neuroscience
Available formats
×

Save element 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.

Philosophy of Neuroscience
Available formats
×

Save element 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.

Philosophy of Neuroscience
Available formats
×