Skip to main content

Representational Mechanisms

  • Chapter
  • First Online:
Neurosemantics

Part of the book series: Studies in Brain and Mind ((SIBM,volume 10))

  • 495 Accesses

Abstract

This chapter attempts to move from the fundamental computational properties of the brain, previously described, into a sketch of how the brain builds a representation of the world. The next part of the book will deal more specifically with the linguistic portrayal we humans have of the world, a topic which has also served as a tentative approach to explaining the neural mechanisms that allow animals to build knowledge. In a sense, the core question of epistemology. Several philosophers, like Jerry Fodor (1983, 1990), have denied that an explanation of representations can be given in terms of neural biophysical properties. Today, such a radical refusal of the neurocomputational approach has become more and more marginal, in any case its thorough defense is beyond the scope of this book, and left to better advocates (Churchland 2002).

On the other hand, any endeavor toward explaining mental representations by neural mechanisms, has first to acknowledge that the notion of “representation” itself is problematic, and at the heart of current philsophical controversies. In addition, there are also positions entirely within a neuroscientific perspective, that deny the concept of representations completely. The title of this chapter leaves no doubt that we, instead, appeal to the notion of representations. It is out of the scope of this book the attempt to settle the philsophical debate on representations, and to lay down any new theory, our approach is to explore the emergence of semantic phenomena through the use of neurocomputational models based on a set of plausible mechanisms, that will be described in this chapter. Before that, we will offer a short overview of the philsophical issues posed by mental representation in general, with more detail on neural representation, and on neural computation over representations.

We will offer a selection of a small number of mechanisms, deemed to be at the core of the bridge between electrochemical activity and world representation. The ability of detecting coincidences, at different scale levels of neural circuits, is regarded as the most general, and probably most effective mechanism. Other more specific strategies will be discussed such as receptive fields, topological organizations, and selective processing pathways.

This is a preview of subscription content, log in via an institution to check access.

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 84.99
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 109.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD 109.99
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

Notes

  1. 1.

    One of the models in this book (Sect. 6.2.3) demonstrates the feasibility of learning from a limited number of experiences, using coincidence detection only.

References

  • Adrian, E. D., & Matthews, R. (1927a). The action of light on the eye: Part I. The discharge of impulses in the optic nerve and its relation to the electric changes in the retina. Journal of Physiology, 63, 378–414.

    Article  Google Scholar 

  • Adrian, E. D., & Matthews, R. (1927b). The action of light on the eye: Part II. The processes involved in retinal excitation. Journal of Physiology, 64, 279–301.

    Article  Google Scholar 

  • Anzai, A., Peng, X., & Van Essen, D. C. (2007). Neurons in monkey visual area V2 encode combinations of orientations. Nature Neuroscience, 10, 1313–1321.

    Article  Google Scholar 

  • Arcizet, F., Jouffrais, C., & Girard, P. (2008). Natural textures classification in area v4 of the macaque monkey. Experimental Brain Research, 189, 109–120.

    Article  Google Scholar 

  • Aristotele (350BCE). On memory and reminiscence (W. D. Ross, Trans., 1930). Oxford: Clarendon Press.

    Google Scholar 

  • Aristotle (335–323BCE). De interpretatione (J. L. Ackrill, Trans., 1975). Oxford: Clarendon Press.

    Google Scholar 

  • Arnauld, A. (1683). Des vraies et des fausses idées. Adolphe Delahays, Paris, published in Oevres Philosophiques, 1843.

    Google Scholar 

  • Arnott, S. R., Binns, M. A., Grady, C. L., & Alain, C. (2004). Assessing the auditory dual-pathway model in humans. NeuroImage, 22, 401–408.

    Article  Google Scholar 

  • Atzori, M., Lei, S., Evans, D. I. P., Kanold, P. O., Phillips-Tansey, E., McIntyre, O., McBain, C. J. (2001). Differential synaptic processing separates stationary from transient inputs to the auditory cortex. Neural Networks, 4, 1230–1237.

    Google Scholar 

  • Avidan, G., Harel, M., Hendler, T., Ben-Bashat, D., Zohary, E., & Malach, R. (2002). Contrast sensitivity in human visual areas and its relationship to object recognition. Journal of Neurophysiology, 87, 3102–3116.

    Google Scholar 

  • Axmacher, N., Mormann, F., Fernández, G., Elger, C. E., & Fell, J. (2006). Memory formation by neuronal synchronization. Brain Research Reviews, 52, 170–182.

    Article  Google Scholar 

  • Ayer, A. (1940). The foundations of empirical knowledge. London: Macmillan.

    Google Scholar 

  • Barnes, J., Howard, R., Senior, C., Brammer, M., Bullmore, E., Simmons, A., & David, A. (1999). The functional anatomy of the mccollough contingent colour after-effect. NeuroReport, 10, 195–199.

    Article  Google Scholar 

  • Bechtel, W. (1998). Representations and cognitive explanations: Assessing the dynamicist’s challenge in cognitive science. Cognition, 22, 295–318.

    Google Scholar 

  • Bechtel, W. (2014). Investigating neural representations: the tale of place cells. Synthese, 1–35. doi:10.1007/s11229-014-0480-8

    Google Scholar 

  • Bechtel, W., & Richardson, R. C. (1993). Discovering complexity: Decomposition and localization as scientific research strategies. Princeton: Princeton University Press.

    Google Scholar 

  • Behrens, T. E., Hunt, L. T., Woolrich, M. W., & Rushworth, M. F. (2008). Associative learning of social value. Nature, 13, 245–249.

    Article  Google Scholar 

  • Belin, P., & Zatorre, R. J. (2000). ‘what’, ‘where’ and ‘how’ in auditory cortex. Nature Neuroscience, 3, 965–966.

    Article  Google Scholar 

  • Belin, P., Fecteau, S., & Bédard, C. (2004). Thinking the voice: Neural correlates of voice perception. Trends in Cognitive Sciences, 8, 129–135.

    Article  Google Scholar 

  • Blasdel, G. G. (1992). Orientation selectivity, preference, and continuity in monkey striate cortex. Journal of Neuroscience, 12, 3139–3161.

    Google Scholar 

  • Blouw, P., Solodkin, E., Thagard, P., & Eliasmith, C. (2015). Concepts as semantic pointers: A framework and computational model. Cognitive Science, 1–35. doi:10.1111/cogs.12265

    Google Scholar 

  • Blumberg, M. S., Freeman, J. H., & Robinson, S. (Eds.). (2010). Oxford handbook of developmental behavioral neuroscience. Oxford: Oxford University Press.

    Google Scholar 

  • Bornkessel-Schlesewsky, I., & Schlesewsky, M. (2013). Reconciling time, space and function: A new dorsalventral stream model of sentence comprehension. Brain and Language, 125, 60–76.

    Article  Google Scholar 

  • Braddick, O., Atkinson, J., & Innocenti, G. M. (Eds.). (2011). The developing brain: From developmental biology to behavioral disorders and their remediation. Cambridge: Cambridge University Press.

    Google Scholar 

  • Bullmore, E., & Sporns, O. (2009). Complex brain networks: Graph theoretical analysis of structural and functional systems. Nature Reviews Neuroscience, 10, 186–198.

    Article  Google Scholar 

  • Burke, C. J., Tobler, P. N., Baddeley, M., & Schultz, W. (2010). Neural mechanisms of observational learning. Proceedings of the Natural Academy of Science USA, 32, 14431–14436.

    Google Scholar 

  • Burkhalter, A., Van Essen, D. C. (1986). Processing of color, form and disparity information in visual areas VP and V2 of ventral extrastriate cortex in the macaque monkey. Journal of Neuroscience, 6, 2327–2351.

    Google Scholar 

  • Caggiano, V., Fogassi, L., Rizzolatti, G., Casile, A., Giese, M. A., & Thier, P. (2012). Mirror neurons encode the subjective value of an observed action. Proceedings of the Natural Academy of Science USA, 29, 11848–11853.

    Google Scholar 

  • Callaway, E. M. (2005). Structure and function of parallel pathways in the primate early visual system. Journal of Physiology, 566, 13–19.

    Article  Google Scholar 

  • Calvo-Merino, B., Grezes, J., Glaser, D. E., Passingham, R., & Haggard, P. (2006). Seeing or doing? Influence of visual and motor familiarity in action observation. Current Biology, 16, 1905–1910.

    Article  Google Scholar 

  • Carruthers, P. (2006). The architecture of the mind. Oxford: Oxford University Press.

    Book  Google Scholar 

  • Cerreira-Perpiñán, M., Goodhill, G. J. (2004). Influence of lateral connections on the structure of cortical maps. Journal of Neurophysiology, 92, 2947–2959.

    Article  Google Scholar 

  • Chemero, A. (2009). Radical embodied cognitive science. Cambridge: MIT.

    Google Scholar 

  • Chirimuuta, M., Gold, I. (2009). The embedded neuron, the enactive field? In J. Bickle (Ed.), Handbook of philosophy and neuroscience. Oxford: Oxford University Press.

    Google Scholar 

  • Chklovskii, D. B., Koulakov, A. A. (2004). Maps in the brain: What can we learn from them? Annual Review of Neuroscience, 27, 369–392.

    Article  Google Scholar 

  • Chomsky, N. (1986). Knowledge of language: Its nature, origins and use. New York: Praeger.

    Google Scholar 

  • Church, A. (1941). The calculi of lambda conversion. Princeton: Princeton University Press.

    Google Scholar 

  • Churchland, P. S. (2002). Brain-wise. Studies in neurophilosophy. Cambridge: MIT.

    Google Scholar 

  • Churchland, P. M. (2010). Concept formation via Hebbian learning: The special case of prototypical causal sequences. In P. Machamer & G. Wolters (Eds.), Interpretation – Ways of thinking about the sciences and the arts (pp. 203–219). Pittsburgh: Pittsburgh University Press.

    Google Scholar 

  • Churchland, P. S., Ramachandran, V., & Sejnowski, T. (1994). A critique of pure vision. In C. Koch & J. Davis (Eds.), Large-scale neuronal theories of the brain. Cambridge: MIT.

    Google Scholar 

  • Cloutman, L. L. (2013). Interaction between dorsal and ventral processing streams: Where, when and how? Brain and Language, 127, 251–263.

    Article  Google Scholar 

  • Colombo, M. (2014a). Explaining social norm compliance. A plea for neural representations. Phenomenology and the Cognitive Sciences, 13, 217–238.

    Article  Google Scholar 

  • Colombo, M. (2014b). Neural representationalism, the hard problem of content and vitiated verdicts. A reply to Hutto & Myin (2013). Phenomenology and the Cognitive Sciences, 13, 257–274.

    Article  Google Scholar 

  • Cooper, R. P., Cook, R., Dickinson, A., & Heyes, C. M. (2013). Associative (not Hebbian) learning and the mirror neuron system. Neuroscience Letters, 540, 28–36.

    Article  Google Scholar 

  • Cosmides, L., & Tooby, J. (1997). The modular nature of human intelligence. In A. Scheibel & J. W. Schopf (Eds.), The origin and evolution of intelligence. Oxford: Oxford University Press.

    Google Scholar 

  • Craver, C. F. (2007). Explaining the brain: Mechanisms and the mosaic unity of neuroscience. Oxford: Oxford University Press.

    Book  Google Scholar 

  • Craver, C. (2010). Prosthetic models. Philosophy of Science, 77, 840–851.

    Article  Google Scholar 

  • Cummins, R. (1989). Meaning and mental representation. Cambridge: MIT.

    Google Scholar 

  • Daugman, J. G. (1980). Two-dimensional spectral analysis of cortical receptive field profiles. Vision Research, 20, 847–856.

    Article  Google Scholar 

  • Daugman, J. G. (1985). Uncertainty relation for resolution in space, spatial frequency, and orientation optimized by two-dimensional visual cortical filters. Journal of the Optical Society of America, 2, 1160–1169.

    Article  Google Scholar 

  • David, S. V., Hayden, B. Y., & Gallant, J. L. (2006). Spectral receptive field properties explain shape selectivity in area V4. Journal of Neurophysiology, 96, 3492–3505.

    Article  Google Scholar 

  • Davidson, D. (1987). Knowing one’s own mind. Proceedings and Addresses of the American Philosophical Association, 60, 441–458.

    Google Scholar 

  • de Charms, R. C., & Zador, A. (2000). Neural representation and the cortical code. Annual Review of Neuroscience, 23, 613–647.

    Article  Google Scholar 

  • de Klerk, C. C., Johnson, M. H., Heyes, C. M., & Southgate, V. (2014). Baby steps: Investigating the development of perceptual-motor couplings in infancy. Developmental Science, 8, 1–11.

    Google Scholar 

  • Denys, K., Vanduffel, W., Fize, D., Nelissen, K., Peuskens, H., Van Essen, D., & Orban, G. A. (2004). The processing of visual shape in the cerebral cortex of human and nonhuman primates: A functional magnetic resonance imaging study. Journal of Neuroscience, 24, 2551–2565.

    Article  Google Scholar 

  • Desimone, R., Albright, T. D., Gross, C. D., & Bruce, C. (1984). Stimulus-selective properties of inferior temporal neurons in the macaque. Journal of Neuroscience, 4, 2051–2062.

    Google Scholar 

  • Devitt, M. (1981). Designation. Cambridge: MIT.

    Google Scholar 

  • DeWitt, I., & Rauschecker, J. P. (2012). Phoneme and word recognition in the auditory ventral stream. Proceedings of the Natural Academy of Science USA, 109, E505–E514.

    Google Scholar 

  • Douglas, R. J., Martin, K. A., & Whitteridge, D. (1989). A canonical microcircuit for neocortex. Neural Computation, 1, 480–488.

    Article  Google Scholar 

  • Downing, P. E., Jiang, Y., Shuman, M., & Kanwisher, N. (2001). A cortical area selective for visual processing of the human body. Science, 293, 2470–2473.

    Article  Google Scholar 

  • Dretske, F. I. (1981). Knowledge and the flow of information. Cambridge: MIT.

    Google Scholar 

  • Dretske, F. I. (1986). Misrepresentation. In R. Bogdan (Ed.), Belief: Form, content and function. Oxford: Oxford University Press.

    Google Scholar 

  • Dreyfus, H. (2002). Intelligence without representation – Merleau-ponty’s critique of mental representation. The relevance of phenomenology to scientific explanation. Phenomenology and the Cognitive Sciences, 1, 367–383.

    Article  Google Scholar 

  • Dudai, Y. (2007). Post-activation state: A critical rite of passage of memories. In Bontempi, B., Silva, A., & Christen, Y. (Eds.), Memories: Molecules and circuits (pp. 69–82). Berlin: Springer.

    Chapter  Google Scholar 

  • Durbin, R., & Mitchison, G. (1990). A dimension reduction framework for understanding cortical maps. Nature, 343, 644–647.

    Article  Google Scholar 

  • Dylla, K. V., Galili, D. S., Szyszka, P., & Lüdke, A. (2013). Trace conditioning in insects – Keep the trace! Frontiers in Psychology, 4, 67.

    Google Scholar 

  • Ebbinghaus, H. (1885). Memory: A contribution to experimental psychology. New York: Dover.

    Google Scholar 

  • Edelman, S. (1999). Representation and recognition in vision. Cambridge: MIT.

    Google Scholar 

  • Eliasmith, C. (2013). How to build a brain: A neural architecture for biological cognition. Oxford: Oxford University Press.

    Book  Google Scholar 

  • Eliasmith, C., & Anderson, C. H. (2003). Neural engineering computation, representation, and dynamics in neurobiological systems. Cambridge: MIT.

    Google Scholar 

  • Engel, S. A., Glover, G. H., & Wandell, B. A. (1997). Retinotopic organization in human visual cortex and the spatial precision of functional MRI. Cerebral Cortex, 7, 181–192.

    Article  Google Scholar 

  • Epstein, R., & Kanwisher, N. (1998). A cortical representation of the local visual environment. Nature, 9, 598–601.

    Article  Google Scholar 

  • Feldman, D. E. (2012). The spike-timing dependence of plasticity. Neuron, 75, 556–571.

    Article  Google Scholar 

  • Felleman, D. J., & Van Essen, D. C. (1991). Distributed hierarchical processing in the primate cerebral cortex. Cerebral Cortex, 1, 1–47.

    Article  Google Scholar 

  • Felleman, D. J., Carman, G., & Van Essen, D. C. (1984). Distributed hierarchical processing in the primate cerebral cortex. Investigative Ophthalmology and Visual Science, 25, 278.

    Google Scholar 

  • Ferber, S., Humphrey, G. K., & Vilis, T. (2005). Segregation and persistence of form in the lateral occipital complex. Neuropsychologia, 43, 41–45.

    Article  Google Scholar 

  • Fodor, J. (1975). The language of thought. Cambridge: Harvard University Press.

    Google Scholar 

  • Fodor, J. (1983). Modularity of mind: And essay on faculty psychology. Cambridge: MIT.

    Google Scholar 

  • Fodor, J. (1987). Psychosemantics: The problem of meaning in the philosophy of mind. Cambridge: MIT.

    Google Scholar 

  • Fodor, J. (1990). A theory of content and other essays. Cambridge: Cambridge University Press.

    Google Scholar 

  • Fodor, J. (2008). Against darwinism. Mind & Language, 23, 1–24.

    Article  Google Scholar 

  • Fodor, J., & Pylyshyn, Z. W. (2015). Minds without meanings: An essay on the content of concepts. Cambridge: MIT.

    Google Scholar 

  • Fresco, N. (2014). Physical computation and cognitive science. Berlin: Springer.

    Book  Google Scholar 

  • Friederici, A. (2012). The cortical language circuit: From auditory perception to sentence comprehension acquisition, comprehension, and production. Trends in Cognitive Sciences, 16, 262–268.

    Article  Google Scholar 

  • Fries, P. (2005). A mechanism for cognitive dynamics: Neuronal communication through neuronal coherence. Trends in Cognitive Sciences, 9, 475–480.

    Article  Google Scholar 

  • Gabor, D. (1946). Theory of communication. Journal IEE, 93, 429–459.

    Google Scholar 

  • Gallistel, C. R. (2000). The replacement of general-purpose learning models with adaptively specialized learning modules. In M. S. Gazzaniga (Ed.), The new cognitive neurosciences (2nd ed., pp. 1179–1191). Cambridge: MIT.

    Google Scholar 

  • Gallistel, C. R. (2010). Learning organs. In J. Bricmont & J. Franck (Eds.), Chomsky notebook (pp. 573–586). Cambridge: Cambridge University Press.

    Google Scholar 

  • Gallistel, C. R., & Balsam, P. D. (2014). Time to rethink the neural mechanisms of learning and memory. Neurobiology of Learning and Memory, 108, 136–144.

    Article  Google Scholar 

  • Gauthier, I., & Bukach, C. (2007). Should we reject the expertise hypothesis? Cognition, 103, 322–330.

    Article  Google Scholar 

  • Gauthier, I., & Tarr, M. (2002). Unraveling mechanisms for expert object recognition: Bridging brain activity and behavior. Journal of Experimental Psychology Human Perception and Performance, 28, 431–446.

    Article  Google Scholar 

  • Gauthier, I., Skudlarski, P., Gore, J. C., & Anderson, A. W. (2000). Expertise for cars and birds recruits brain areas involved in face recognition. Nature Neuroscience, 3, 191–197.

    Article  Google Scholar 

  • Gegenfurtner, K. R., Kiper, D. C., & Levitt, J. B. (1997). Functional properties of neurons in macaque area V3. Journal of Neurophysiology, 77, 1906–1923.

    Google Scholar 

  • Gelder, T. v. (1995). What might cognition be, if not computation? Journal of Philosophy, 91, 345–381.

    Article  Google Scholar 

  • Gennari, F. (1782). De peculiari structura cerebri, nonnulisque ejus morbis. Parma: Ex regio typographeo.

    Google Scholar 

  • George, D., & Hawkins, J. (2009). Towards a mathematical theory of cortical micro-circuits. PLoS Computational Biology, 5, e1000532.

    Article  Google Scholar 

  • Gilbert, C. D., & Wiesel, T. N. (1983). Clustered intrinsic connections in cat visual cortex. Journal of Neuroscience, 3, 1116–1133.

    Google Scholar 

  • Glenberg, A. M., & Gallese, V. (2012). Action-based language: A theory of language acquisition, comprehension, and production. Cognition, 48, 905–922.

    Google Scholar 

  • Goodale, M. A., & Milner, A. D. (1992). Separate visual pathways for perception and action. Trends in Neuroscience, 15, 20–25.

    Article  Google Scholar 

  • Goodman, N. (1976). Languages of art: An approach to a theory of symbols. Indianapolis: Hackett.

    Google Scholar 

  • Gould, S. J., & Lewontin, R. C. (1979). The spandrels of San Marco and the Panglossian paradigm: A critique of the adaptationist programme. Philosophical Transactions of the Royal Society of London Series B, Biological Sciences, 205, 581–598.

    Google Scholar 

  • Griffiths, T. D., & Warren, J. D. (2004). What is an auditory object? Neuron, 43, 237–249.

    Article  Google Scholar 

  • Griffiths, T. L., Chater, N., Kemp, C., Perfors, A., & Tenenbaum, J. B. (2010). Probabilistic models of cognition: Exploring representations and inductive biases. Trends in Cognitive Sciences 14, 357–364.

    Article  Google Scholar 

  • Grill-Spector, K. (2003). The functional organization of the ventral visual pathway and its relationship to object recognition. In N. Kanwisher & J. Duncan (Eds.), Attention and performance XX. Functional brain imaging of visual cognition. Oxford: Oxford University Press.

    Google Scholar 

  • Grill-Spector, K., Kushnir, T., Edelman, S., Itzchak, Y., & Malach, R. (1998). Cue-invariant activation in object-related areas in the human occipital lobe. Neuron, 21, 191–202.

    Article  Google Scholar 

  • Grill-Spector, K., Kushnir, T., Edelman, S., Avidan-Carmel, G., Itzchak, Y., & Malach, R. (1999). Differential processing of objects under various viewing conditions in the human lateral occipital complex. Neuron, 24, 187–203.

    Article  Google Scholar 

  • Grush, R. (1997). The architecture of representation. Philosophical Psychology, 10, 5–23.

    Article  Google Scholar 

  • Grush, R., & Mandik, P. (2002). Representational parts. Phenomenology and the Cognitive Sciences, 1, 389–394.

    Article  Google Scholar 

  • Hagmann, P., Cammoun, L., Gigandet, X., Gerhard, S., Grant, P. E., Wedeen, V., Meuli, R., Thiran, J. P., Honey, C. J., & Sporns, O. (2010). MR connectomics: Principles and challenges. Journal of Neuroscience Methods, 194, 34–45.

    Article  Google Scholar 

  • Hall, G. (1991). Perceptual and associative learning. Oxford: Oxford University Press.

    Book  Google Scholar 

  • Hartline, H. K. (1938). The response of single optic nerve fibers of the vertebrate eye to illumination of the retina. American Journal of Physiology, 121, 400–415.

    Google Scholar 

  • Hartline, H. K. (1967). Visual receptors and retinal interaction. Science, 164, 270–278.

    Article  Google Scholar 

  • Hartline, H. K., Ratliff, F., & Miller, W. H. (1961). Inhibitory interaction in the retina and its significance in vision. In E. Florey (Ed.), Nervous inhibition. New York: Pergamon Press.

    Google Scholar 

  • Haselager, P., de Groot, A., & van Rappard, H. (2003). Representationalism vs. anti-representationalism: A debate for the sake of appearance. Philosophical Psychology, 16, 5–23.

    Article  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.

    Article  Google Scholar 

  • Haxby, J. V., Gobbini, M. I., Furey, M. L., Ishai, A., Schouten, J. L., & Pietrini, P. (2001). Distributed and overlapping representations of objects and faces in ventral temporal cortex. Science, 293, 2425–2430.

    Article  Google Scholar 

  • Hebb, D. O. (1949). The organization of behavior. New York: Wiley.

    Google Scholar 

  • Heider, B., Meskenaite, V., & Peterhans, E. (2000). Anatomy and physiology of a neural mechanism defining depth order and contrast polarity at illusory contours. European Journal of Neuroscience, 12, 4117–4130.

    Article  Google Scholar 

  • Hempel, C. G. (1965). Aspects of scientific explanation and other essays in the philosophy of science. New York: Free Press.

    Google Scholar 

  • Heyes, C. (2010). Where do mirror neurons come from? Neuroscience & Biobehavioral Reviews, 34, 575–583.

    Article  Google Scholar 

  • Heyes, C. (2012). Simple minds: A qualified defence of associative learning. Philosophical Transactions of the Royal Society B, 367, 2695–2703.

    Article  Google Scholar 

  • Hickok, G. (2009). Eight problems for the mirror neuron theory of action understanding in monkeys and humans. Journal of Cognitive Neuroscience, 21, 1229–1243.

    Article  Google Scholar 

  • Hickok, G. (2010). The role of mirror neurons in speech perception and action word semantics. Language and Cognitive Processes, 25, 749–776.

    Article  Google Scholar 

  • Hickok, G., & Poeppel, D. (2007). The cortical organization of speech processing. Nature Reviews Neuroscience, 8, 393–402.

    Article  Google Scholar 

  • Horton, J. C., & Adams, D. L. (2005). The cortical column: A structure without a function. Philosophical Transactions of the Royal Society B, 360, 837–862.

    Article  Google Scholar 

  • Hubel, D. H., & Livingstone, M. S. (1987). Segregation of form, color, and stereopsis in primate area 18. Journal of Neuroscience, 7, 3378–3415.

    Google Scholar 

  • Hubel, D., & Wiesel, T. (1959). Receptive fields of single neurones in the cat’s striate cortex. Journal of Physiology, 148, 574–591.

    Article  Google Scholar 

  • Hubel, D., & Wiesel, T. (1962). Receptive fields, binocular interaction, and functional architecture in the cat’s visual cortex. Journal of Physiology, 160, 106–154.

    Article  Google Scholar 

  • Hubel, D., & Wiesel, T. (1968). Receptive fields and functional architecture of monkey striate cortex. Journal of Physiology, 195, 215–243.

    Article  Google Scholar 

  • Hubel, D., & Wiesel, T. (1974a). Ordered arrangement of orientation columns in monkeys lacking visual experience. Journal of Comparative Neurology, 158, 307–318.

    Article  Google Scholar 

  • Hubel, D., & Wiesel, T. (1974b). Uniformity of monkey striate cortex: A parallel relationship between field size, scatter, and magnification factor. Journal of Comparative Neurology, 158, 295–305.

    Article  Google Scholar 

  • Hubel, D. H., & Wiesel, T. N. (2004). Brain and visual perception: The story of a 25-year collaboration. Oxford: Oxford University Press.

    Book  Google Scholar 

  • Hume, D. (1739). A treatise of human nature (Vols. 1, 2). London: John Noon.

    Google Scholar 

  • Hume, D. (1748). An enquiry concerning human understanding. London: A. Millar.

    Book  Google Scholar 

  • Hunt, J. J., Bosking, W. H., & Goodhill, G. J. (2011). Statistical structure of lateral connections in the primary visual cortex. Neural Systems & Circuits, 1, 1–12.

    Article  Google Scholar 

  • Hutto, D. D., & Myin, E. (2013). Radicalizing enactivism: Basic minds without content. Cambridge: MIT.

    Google Scholar 

  • Hutto, D. D., & Myin, E. (2014). Neural representations not needed – No more pleas, please. Phenomenology and the Cognitive Sciences, 13, 241–256.

    Article  Google Scholar 

  • Iacoboni, M., Woods, R. P., Brass, M., Bekkering, H., Mazziotta, J. C., & Rizzolatti, G. (1999). Cortical mechanisms of human imitation. Science, 286, 2526–2528.

    Article  Google Scholar 

  • Ito, M. (1984). The cerebellum and neural control. New York: Raven Press.

    Google Scholar 

  • Ito, M., & Komatsu, H. (2004). Representation of angles embedded within contour stimuli in area V2 of macaque monkeys. Journal of Neuroscience, 24, 3313–3324.

    Article  Google Scholar 

  • Johnson-Laird, P. (1983). Mental models: Towards a cognitive science of language, inference and consciousness. Cambridge: Cambridge University Press.

    Google Scholar 

  • Kaas, J. H. (1997). Topographic maps are fundamental to sensory processing. Brain Research Bulletin, 44, 107–112.

    Article  Google Scholar 

  • Kaas, J. H., & Lyon, D. C. (2001). Visual cortex organization in primates: Theories of V3 and adjoining visual areas. Progress in Brain Research, 134, 285–295.

    Article  Google Scholar 

  • Kaas, J. H., Gharbawie, O. A., & Stepniewska, I. (2011). The organization and evolution of dorsal stream multisensory motor pathways in primates. Frontiers in Nauroanatomy, 5, 34.

    Google Scholar 

  • Kanwisher, N. (2003). The ventral visual object pathway in humans: Evidence from fMRI. In L. Chalupa & J. Werner (Eds.), The visual neurosciences. Cambridge: MIT.

    Google Scholar 

  • Kanwisher, N., McDermott, J., & Chun, M. (1997). The fusiform face area: A module in human extrastriate cortex specialized for face perception. Journal of Neuroscience, 17, 4302–4311.

    Google Scholar 

  • Kaplan, D. M. (2011). Explanation and description in computational neuroscience. Synthese, 183, 339–373.

    Article  Google Scholar 

  • Kaplan, D. M., & Craver, C. F. (2011). Towards a mechanistic philosophy of neuroscience. In S. French & J. Saatsi (Eds.), Continuum companion to the philosophy of science (pp. 268–292). London: Continuum Press.

    Google Scholar 

  • Keysers, C., & Perrett, D. I. (2004). Demystifying social cognition: A Hebbian perspective. Trends in Cognitive Sciences, 8, 501–507.

    Article  Google Scholar 

  • Kim, J. G., Biederman, I., Lescroart, M. D., & Hayworth, K. J. (2009). Adaptation to objects in the lateral occipital complex (loc): Shape or semantics? Vision Research, 49, 2297–2305.

    Article  Google Scholar 

  • Kleene, S. C. (1936). General recursive functions of natural numbers. Mathematische Annalen, 112, 727–742.

    Article  Google Scholar 

  • Kobatake, E., & Tanaka, K. (1994). Neuronal selectivities to complex object features in the ventral visual pathway of the macaque cerebral cortex. Journal of Neurophysiology, 71, 856–867.

    Google Scholar 

  • Koch, G., Ponzo, V., Lorenzo, F. D., Caltagirone, C., & Veniero, D. (2013). Hebbian and anti-Hebbian spike-timing-dependent plasticity of human cortico-cortical connections. Journal of Neuroscience, 33, 9725–9733.

    Article  Google Scholar 

  • Kohonen, T., & Hari, R. (2000). Where the abstract feature maps of the brain might come from. Trends in Neurosciences, 22, 135–139.

    Article  Google Scholar 

  • Komosinski, M. (2000). The world of framsticks: Simulation, evolution, interaction. In Proceedings of the 2nd international conference on virtual worlds (pp. 214–224). Paris: Springer.

    Google Scholar 

  • Kourtzi, Z., & Kanwisher, N. (2000). Cortical regions involved in perceiving object shape. Journal of Neuroscience, 20, 3310–3318.

    Google Scholar 

  • Kourtzi, Z., Erb, M., Grodd, W., & Bülthoff, H. H. (2003). Representation of the perceived 3-d object shape in the human lateral occipital complex. Cerebral Cortex, 13, 911–920.

    Article  Google Scholar 

  • Koutstaal, W., Wagner, A. D., Rotte, M., Maril, A., Buckner, R., & Schacter, D. L. (2001). Perceptual specificity in visual object priming: Functional magnetic resonance imaging evidence for a laterality difference in fusiform cortex. Neuropsychologia, 2, 184–199.

    Article  Google Scholar 

  • Krubitzer, L. (1995). The organization of neocortex in mammals: Are species differences really so different? Trends in Neuroscience, 8, 408–417.

    Article  Google Scholar 

  • Kubovy, M., & Valkenburg, D. V. (2000). Auditory and visual objects. Nature Neuroscience, 3, 191–197.

    Article  Google Scholar 

  • Kuffler, S. W. (1953). Discharge patterns and functional organization of mammalian retina. Journal of Neurophysiology, 16, 37–68.

    Google Scholar 

  • Landisman, C. E., & Ts’o, D. Y. (2002). Color processing in macaque striate cortex: Relationships to ocular dominance, cytochrome oxidase, and orientation. Journal of Neurophysiology, 87, 3126–3137.

    Google Scholar 

  • Large, M. E., Aldcroft, A., & Vilis, T. (2007). Task-related laterality effects in the lateral occipital complex. Brain, 1128, 130–138.

    Article  Google Scholar 

  • LeDoux, J. E. (2000). Emotion circuits in the brain. Annual Review of Neuroscience, 23, 155–184.

    Article  Google Scholar 

  • Liebenthal, E., Binder, J. R., Spitzer, S. M., Possing, E. T., & Medler, D. A. (2005). Neural substrates of phonemic perception. Cerebral Cortex, 15, 1621–1631.

    Article  Google Scholar 

  • Lingnau, A., Gesierich, B., & Caramazza, A. (2009). Asymmetric fMRI adaptation reveals no evidence for mirror neurons in humans. Proceedings of the Natural Academy of Science USA, 106, 9925–9930.

    Google Scholar 

  • Linsker, R. (1989). How to generate ordered maps by maximizing the mutual information between input and output signals. Neural Computation, 1, 402–411.

    Article  Google Scholar 

  • Liu, H., & Wen, X. (2013). On formalizing causation based on constant conjunction theory. The Review of Symbolic Logic, 6, 160–181.

    Article  Google Scholar 

  • Livingstone, M. S., & Hubel, D. H. (1987). Psychophysical evidence for separate channels for the perception of form, color, movement, and depth. Journal of Neuroscience, 7, 3416–3468.

    Google Scholar 

  • Locke, J. (1690). An essay concerning human understanding. Cleveland: Meridian Books.

    Book  Google Scholar 

  • Lorente de Nó, R. (1938). Architectonics and structure of the cerebral cortex. In J. Fulton (Ed.), Physiology of the nervous system (pp. 291–330). Oxford: Oxford University Press.

    Google Scholar 

  • Maçarico da Costa, N., Martin, K. A. C. (2010). Whose cortical column would that be? Frontiers in Nauroanatomy, 4, 16.

    Google Scholar 

  • Machamer, P., Darden, L., & Craver, C. F. (2000). Thinking about mechanisms. Philosophy of Science, 67, 1–84.

    Article  Google Scholar 

  • Macko, K., Jarvis, C., Kennedy, C., Miyaoka, M., Shinohara, M., Sololoff, L., & Mishkin, M. (1982). Mapping the primate visual system with [2–14c]deoxyglucose. Science, 218, 394–397.

    Article  Google Scholar 

  • Malach, R., Reppas, J. B., Benson, R. R., Kwong, K. K., Jiang, H., Kennedy, W. A., Ledden, P. J., Brady, T. J., Rosen, B. R., & Tootell, R. B. (1995). Object-related activity revealed by functional magnetic resonance imaging in human occipital cortex. Proceedings of the Natural Academy of Science USA, 92, 8135–8139.

    Google Scholar 

  • Malach, R., Levy, I., & Hasson, U. (2002). The topography of high-order human object areas. Trends in Cognitive Sciences, 6, 176–184.

    Article  Google Scholar 

  • Malebranche, N. (1675). De la recherche de la vérité. Amsterdam: Henry Desbordes.

    Google Scholar 

  • Mandik, P. (2003). Varieties of representation in evolved and embodied neural networks. Biology and Philosophy, 18, 95–130.

    Article  Google Scholar 

  • Marr, D. (1970). A theory for cerebral neocortex. Proceedings of the Royal Society of London B, 176, 161–234.

    Google Scholar 

  • Maunsell, J., Van Essen, D. C. (1983). The connections of the middle temporal visual area (mt) and their relation ship to a cortical hierarchy in the macaque monkey. Journal of Neuroscience, 3, 2563–2586.

    Google Scholar 

  • McCollough, C. (1965). Color adaptation of edge-detectors in the human visual system. Science, 149, 1115–1116.

    Article  Google Scholar 

  • Miłkowski, M. (2015a). The hard problem of content: Solved (long ago). Studies in Logic, Grammar and Rhetoric, 41, 73–88.

    Article  Google Scholar 

  • Miłkowski, M. (2015b). Satisfaction conditions in anticipatory mechanisms. Biology and Philosophy, 30, 709–728.

    Article  Google Scholar 

  • Miller, L. M., Escabí, M. A., Read, H. L., & Schreiner, C. E. (2002b). Spectrotemporal receptive fields in the lemniscal auditory thalamus and cortex. Journal of Neurophysiology, 87, 516–527.

    Google Scholar 

  • Millikan, R. G. (1984). Language, thought, and other biological categories: New foundations for realism. Cambridge: MIT.

    Google Scholar 

  • Millikan, R. G. (2000). On Clar and confused ideas: An essay about substance concepts. Cambridge: Cambridge University Press.

    Book  Google Scholar 

  • Morita, T., Kochiyama, T., Okada, T., Yonekura, Y., Matsumura, M., & Sadato, N. (2004). The neural substrates of conscious color perception demonstrated using fMRI. NeuroImage, 21, 1665–1673.

    Article  Google Scholar 

  • Mountcastle, V. (1957). Modality and topographic properties of single neurons in cats somatic sensory cortex. Journal of Neurophysiology, 20, 408–434.

    Google Scholar 

  • Mountcastle, V. (2003). Introduction. Cerebral Cortex, 13, 2–4.

    Article  Google Scholar 

  • Murphy, E. H., & Berman, N. (1979). The rabbit and the cat: A comparison of some features of response properties of single cells in the primary visual cortex. Journal of Comparative Neurology, 188, 401–427.

    Article  Google Scholar 

  • Murray, S. O., & He, S. (2001). Contrast invariance in the human lateral occipital complex depends on attention. Cerebral Cortex, 16, 606–611.

    Google Scholar 

  • Nair-Collins, M. (2013). Representation in biological systems: Teleofunction, etiology, and structural preservation. In L. Swan (Ed.), Origins of mind (pp. 161–185). New York: Academic.

    Chapter  Google Scholar 

  • Nealey, T. A., & Maunsell, J. H. R. (1994). Magnocellular and parvocellular contributions to the responses of neurons in macaque striate cortex. Journal of Neuroscience, 14, 2069–2079.

    Google Scholar 

  • Nessler, B., Pfeiffer, M., Buesing, L., & Maass, W. (2013). Bayesian computation emerges in generic cortical microcircuits through spike-timing-dependent plasticity. PLoS Computational Biology, 9, e1003037.

    Article  Google Scholar 

  • Nestor, A., Plaut, D. C., & Behrmann, M. (2011). Unraveling the distributed neural code of facial identity through spatiotemporal pattern analysis. Proceedings of the Natural Academy of Science USA, 108, 9998–10003.

    Google Scholar 

  • Newell, A., & Simon, H. A. (1976). Computer science as empirical enquiry: Symbols and search. Communications of the Association for Computing Machinery, 19, 113–126.

    Article  Google Scholar 

  • Niell, C., & Stryker, M. P. (2010). Modulation of visual responses by behavioral state in mouse visual cortex. Neuron, 65, 472–479.

    Article  Google Scholar 

  • Nieuwenhuys, R., Voogd, J., & van Huijzen, C. (2008). The human central nervous system. Berlin: Springer.

    Book  Google Scholar 

  • O’Brien, G., & Opie, J. (2004). Notes toward a structuralist theory of mental representation. In H. Clapin, P. Staines, & P. Slezak (Eds.), Representation in mind – New approaches to mental representation. Amsterdam: Elsevier.

    Google Scholar 

  • O’Keefe, J., & Nadel, L. (1978). The hippocampus as a cognitive map. Oxford: Oxford University Press.

    Google Scholar 

  • O’Keefe, J., & Recce, M. (1993). Phase relationship between hippocampal place units and the EEG theta rhytm. Hippocampus, 3, 317–330.

    Article  Google Scholar 

  • Orban, G. A., Zhu, Q., & Vanduffel, W. (2014). The transition in the ventral stream from feature to real-world entity representations. Frontiers in Psychology, 5, 695.

    Google Scholar 

  • Osterhout, L., Kim, A., & Kuperberg, G. R. (2007). The neurobiology of sentence comprehension. In M. Spivey, K. McRae, & M. Joanisse (Eds.), The Cambridge handbook of psycholinguistics (pp. 365–389). Cambridge: Cambridge University Press.

    Google Scholar 

  • O’Toole, A. J., Jiang, F., Abdi, H., & Haxby, J. V. (2005). Partially distributed representations of objects and faces in ventral temporal cortex. Journal of Cognitive Neuroscience, 17, 580–590.

    Article  Google Scholar 

  • Palmer, S. (1978). Fundamental aspects of cognitive representation. In E. Rosch & B. Lloyd (Eds.), Cognition and categorization. Mahwah: Lawrence Erlbaum Associates.

    Google Scholar 

  • Papeo, L., Negri, G. A. L., Zadini, A., & Rumiati, R. I. (2010). Action performance and action-word understanding: Evidence of double dissociations in left-damaged patients. Cognitive Neuropsychology, 27, 428–461.

    Article  Google Scholar 

  • Papineau, D. (1987). Reality and representation. Oxford: Basil Blackwell.

    Google Scholar 

  • Papineau, D. (1993). Philsophical naturalism. Oxford: Basil Blackwell.

    Google Scholar 

  • Paradiso, M., MacEvoy, S. P., Huang, X., & Blau, S. (2005). The importance of modulatory input for V1 activity and perception. Progress in Brain Research, 149, 257–267.

    Article  Google Scholar 

  • Pavlov, I. (1927). Conditioned reflexes. Oxford: Oxford University Press.

    Google Scholar 

  • Peelen, M. V., & Downing, P. E. (2007). The neural basis of visual body perception. Nature Reviews Neuroscience, 8, 636–648.

    Article  Google Scholar 

  • Piccinini, G. (2006). Computational explanation in neuroscience. Synthese, 153, 343–353.

    Article  Google Scholar 

  • Piccinini, G. (2007). Computational modeling vs. computational explanation: Is everything a Turing Machine, and does it matter to the philosophy of mind? Australasian Journal of Philosoph, 85, 93–115.

    Article  Google Scholar 

  • Piccinini, G., & Bahar, S. (2013). Neural computation and the computational theory of cognition. Cognitive Science, 34, 453–488.

    Article  Google Scholar 

  • Piccinini, G., & Scarantino, A. (2010). Computation vs. information processing: Why their difference matters to cognitive science. Studies in History and Philosophy of Science, 41, 237–246.

    Article  Google Scholar 

  • Plate, T. (2003). Holographic reduced representations. Stanford: CSLI Publication.

    Google Scholar 

  • Plebe, A. (2008). The ventral visual path: Moving beyond V1 with computational models. In T. A. Portocello & R. B. Velloti (Eds.), Visual cortex: New research (pp. 97–160). New York: Nova Science Publishers.

    Google Scholar 

  • Pour-El, M. B., & Richards, I. (1981). Wave equation with computable initial data such that its unique solution is not computable. Advances in Mathematics, 39, 215–239.

    Article  Google Scholar 

  • Press, W. A., Brewer, A. A., Dougherty, R. F., Wade, A. R., & Wandell, B. A. (2001). Visual areas and spatial summation in human visual cortex. Vision Research, 41, 1321–1332.

    Article  Google Scholar 

  • Prinz, J. (2002). Furnishing the mind – Concepts and their perceptual basis. Cambridge: MIT.

    Google Scholar 

  • Prinz, J. (2006b). Is the mind really modular? In R. Morris & L. Tarassenko (Eds.), Cognitive systems: Information processing meets brain science (pp. 22–36). Amsterdam: Elsevier.

    Google Scholar 

  • Pulvermüller, F. (2010). Brain embodiment of syntax and grammar: Discrete combinatorial mechanisms spelt out in neuronal circuits. Brain and Language, 112, 167–179.

    Article  Google Scholar 

  • Pulvermüller, F., & Fadiga, L. (2010). Active perception: Sensorimotor circuits as a cortical basis for language. Nature Reviews Neuroscience, 11, 351–360.

    Article  Google Scholar 

  • Putnam, H. (1988). Representation and reality. Cambridge: MIT.

    Google Scholar 

  • Pylyshyn, Z. (1981). Computation and cognition: Issues in the foundations of cognitive science. Behavioral and Brain Science, 3, 111–150.

    Article  Google Scholar 

  • Rajimehr, R., Devaney, K. J., Bilenko, N. Y., Young, J. C., & Tootell, R. B. (2011). The “parahippocampal place area” responds preferentially to high spatial frequencies in humans and monkeys. PLoS Computational Biology, 9, e1000608.

    Article  Google Scholar 

  • Rakic, P. (2008). Confusing cortical columns. Proceedings of the Natural Academy of Science USA, 34, 12,099–12,100.

    Google Scholar 

  • Ramsey, W. M. (2007). Representation reconsidered. Cambridge: Cambridge University Press.

    Book  Google Scholar 

  • Rauschecker, J. P., & Scott, S. K. (2009). Maps and streams in the auditory cortex: Nonhuman primates illuminate human speech processing. Nature Neuroscience, 12, 718–724.

    Article  Google Scholar 

  • Ray, E., & Heyes, C. (2011). Imitation in infancy: The wealth of the stimulus. Developmental Science, 14, 92–105.

    Article  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 II: Current theory and research (pp. 64–99). New York: Appleton Century Crofts.

    Google Scholar 

  • Rice, H. (1954). Classes of recursively enumerable sets and their decision problems. Transaction of American Mathematical Society, 74, 358–366.

    Article  Google Scholar 

  • Rizzolatti, G., & Arbib, M. A. (1998). Language within our grasp. Trends in Neuroscience, 21, 188–194.

    Article  Google Scholar 

  • Rizzolatti, G., & Matelli, M. (2003). Two different streams form the dorsal visual system: Anatomy and functions. Experimental Brain Research, 153, 146–157.

    Article  Google Scholar 

  • Rizzolatti, G., & Sinigaglia, C. (2010). The functional role of the parieto-frontal mirror circuit: Interpretations and misinterpretations. Nature Reviews Neuroscience, 11, 264–274.

    Article  Google Scholar 

  • Rizzolatti, G., Camarda, R., Fogassi, L., Gentilucci, M., Luppino, G., & Matelli, M. (1988). Functional organization of inferior area 6 in the macaque monkey. Experimental Brain Research, 71, 491–507.

    Article  Google Scholar 

  • Rizzolatti, G., Fogassi, L., & Gallese, V. (2001). Neurophysiological mechanisms underlying the understanding and imitation of action. Nature Reviews Neuroscience, 2, 661–670.

    Article  Google Scholar 

  • Robbins, R. A., & McKone, E. (2007). No face-like processing for objects-of-expertise in three behavioural tasks. Cognition, 103, 34–79.

    Article  Google Scholar 

  • Rockel, A., Hiorns, R., & Powell, T. (1980). The basic uniformity in structure of the neocortex. Brain, 103, 221–244.

    Article  Google Scholar 

  • Rodieck, R. W. (1965). Quantitative analysis of cat retinal ganglion cell response to visual stimuli. Vision Research, 5, 583–601.

    Article  Google Scholar 

  • Rogalsky, C., Love, T., Driscoll, D., Andersond, S. W., & Hickok, G. (2011). Are mirror neurons the basis of speech perception? Evidence from five cases with damage to the purported human mirror system. Neurocase, 17, 178–187.

    Article  Google Scholar 

  • Romanski, L. M., Tian, B., Fritz, J., Mishkin, M., Goldman-Rakici, P. S., & Rauschecker, J. (1999). Dual streams of auditory afferents target multiple domains in the primate prefrontal cortex. Nature Neuroscience, 2, 1131–1136.

    Article  Google Scholar 

  • Rose, D. (1979). Mechanisms underlying the receptive field properties of neurons in cat visual cortex. Vision Research, 19, 533–544.

    Article  Google Scholar 

  • Rubenstein, J. L. R., Rakic, P. (Eds.). (2013a). Comprehensive developmental neuroscience: Neural circuit development and function in the healthy and diseased brain. New York: Academic.

    Google Scholar 

  • Rubenstein, J. L. R., Rakic, P. (Eds.). (2013b). Comprehensive developmental neuroscience: Patterning and cell type specification in the developing CNS and PNS. New York: Academic.

    Google Scholar 

  • Rupert, R. D. (1999). The best test theory of extension: First principle(s). Minds and Language, 14, 321–355.

    Article  Google Scholar 

  • Russell, B. (1927). The analysis of matter. London: Harcourt.

    Google Scholar 

  • Ryder, D. (2004). SINBAD neurosemantics: A theory of mental representation. Minds and Machines, 19, 211–240.

    Google Scholar 

  • Ryder, D. (2009a). Problems of representation I: Nature and role. In J. Symons & P. Calvo (Eds.), The Routledge companion to philosophy of psychology (pp. 233–250). London: Routledge.

    Google Scholar 

  • Ryder, D. (2009b). Problems of representation II: Naturalizing content. In J. Symons & P. Calvo (Eds.), The Routledge companion to philosophy of psychology (pp. 251–279). London: Routledge.

    Google Scholar 

  • Saur, D., Kreher, B. W., Schnell, S., Kümmerer, D., Kellmeyer, P., Vry, M. S., Umarova, R., Musso, M., Glauche, V., Abel, S., Huber, W., Rijntjes, M., Hennig, J., & Weiller, C. (2008). Ventral and dorsal pathways for language. Proceedings of the Natural Academy of Science USA, 105, 18035–18040.

    Google Scholar 

  • Schneider, G. E. (1967). Visual receptors and retinal interaction. Science, 164, 270–278.

    Google Scholar 

  • Schouenborg, J., Garwicz, M., & Danielsen, N. (Eds.). (2011). Brain machine interfaces – Implications for science, clinical practice and society. Amsterdam: Elsevier.

    Google Scholar 

  • Searle, J. R. (1990). Is the brain a digital computer? Proceedings and Addresses of the American Philosophical Association, 64, 21–37.

    Google Scholar 

  • Shagrir, O. (2012). Structural representations and the brain. British Journal for the Philosophy of Science, 63, 519–545.

    Article  Google Scholar 

  • Shanks, D. R. (1995). The psychology of associative learning. Cambridge: Cambridge University Press.

    Book  Google Scholar 

  • Shepard, R. N., & Chipman, S. (1970). Second-order isomorphism of internal representations: Shapes of states. Cognitive Psychology, 1, 1–17.

    Article  Google Scholar 

  • Shepherd, G. M. (1988). A basic circuit for cortical organization. In M. S. Gazzaniga (Ed.), Perspectives on memory research (pp. 93–134). Cambridge: MIT.

    Google Scholar 

  • Simons, J. S., Koutstaal, W., Prince, S., Wagner, A. D., & Schacter, D. L. (2003). Neural mechanisms of visual object priming: Evidence for perceptual and semantic distinctions in fusiform cortex. NeuroImage, 19, 613–626.

    Article  Google Scholar 

  • Singer, W. (1995). Synchronization of neuronal responses as a putative binding mechanism. In M. Arbib (Ed.), The handbook of brain theoriey and neural networks (pp. 960–964). Cambridge: MIT.

    Google Scholar 

  • Sirosh, J., Miikkulainen, R., & Choe, Y. (Eds.). (1996). Lateral interactions in the cortex: Structure and function. Austin: The UTCS Neural Networks Research Group.

    Google Scholar 

  • Slezak, P. (2002). The tripartite model of representation. Philosophical Psychology, 15, 239–270.

    Article  Google Scholar 

  • Smith, B. C. (2002). The foundations of computing. In M. Scheutz (Ed.), Computationalism – New directions (pp. 23–58). Cambridge: MIT.

    Google Scholar 

  • Sprevak, M. (2011). William M. Ramsey, representation reconsidered. British Journal for the Philosophy of Science, 62, 669–675.

    Article  Google Scholar 

  • Stettler, D. D., Das, A., Bennett, J., & Gilbert, C. D. (2002). Lateral connectivity and contextual interactions in macaque primary visual cortex. Neuron, 36, 739–750.

    Article  Google Scholar 

  • Stowe, L. A., Haverkort, M., & Zwarts, F. (2004). Rethinking the neurological basis of language. Lingua, 115, 997–1042.

    Article  Google Scholar 

  • Swindale, N. V. (2001). Keeping the wires short: A singularly difficult problem. Neuron, 29, 316–317.

    Article  Google Scholar 

  • Swoyer, C. (1991). Structural representation and surrogative reasoning. Synthese, 87, 449–508.

    Article  Google Scholar 

  • Tanaka, K. (1996). Inferotemporal cortex and object vision. Annual Review of Neuroscience, 19, 109–139.

    Article  Google Scholar 

  • Tanaka, K., Saito, H., Fukada, Y., & Moriya, M. (1991). Coding visual images of objects in the inferotemporal cortex of the macaque monkey. Journal of Neurophysiology, 66, 170–189.

    Google Scholar 

  • Tarr, M. J., & Bülthoff, H. H. (1998). Image-based object recognition in man, monkey, and machine. In M. J. Tarr & H. H. Bülthoff (Eds.), Object recognition in man, monkey, and machine. Cambridge: MIT.

    Google Scholar 

  • Tarr, M., & Gauthier, I. (2000). FFA: A flexible fusiform area for subordinate-level visual processing automatized by expertise. Nature Neuroscience, 3, 764–769.

    Article  Google Scholar 

  • Tenenbaum, J. B., Kemp, C., Griffiths, T. L., & Goodman, N. D. (2011). How to grow a mind: Statistics, structure, and abstraction. Science, 331, 1279–1285.

    Article  Google Scholar 

  • Tettamanti, M., & Moro, A. (2012). Can syntax appear in a mirror (system)? Cognition, 48, 923–935.

    Google Scholar 

  • Thivierge, J. P., & Marcus, G. F. (2007). The topographic brain: From neural connectivity to cognition. Trends in Neuroscience, 30, 251–259.

    Article  Google Scholar 

  • Thorndike, E. (1892). Animal intelligence: An experimental study of the associative processes in animals. Psychological Monographs, 2, 192–205.

    Google Scholar 

  • Tiao, Y., & Blakemore, C. (1976). Functional organization in the visual cortex of the golden hamster. Journal of Comparative Neurology, 168, 459–481.

    Article  Google Scholar 

  • Toni, I., de Lange, F. P., Noordzij, M. L., & Hagoort, P. (2008). Language beyond action. Journal of Physiology – Paris, 102, 71–79.

    Article  Google Scholar 

  • Tootell, R. B., Silverman, M. S., Hamilton, S. L., Switkes, E., & De Valois, R. (1988a). Functional anatomy of the macaque striate cortex. V. spatial frequency. Journal of Neuroscience, 8, 1610–1624.

    Google Scholar 

  • Tootell, R. B., Switkes, E., Silverman, M. S., & Hamilton, S. L. (1988b). Functional anatomy of the macaque striate cortex. I. Ocular dominance, binocular interactions, and baseline conditions. Journal of Neuroscience, 8, 1531–1568.

    Google Scholar 

  • Tootell, R. B., Switkes, E., Silverman, M. S., & Hamilton, S. L. (1988c). Functional anatomy of the macaque striate cortex. II. Retinotopic organization. Journal of Neuroscience, 8, 1531–1568.

    Google Scholar 

  • Tootell, R. B., Switkes, E., Silverman, M. S., & Hamilton, S. L. (1988d). Functional anatomy of the macaque striate cortex. III. Color. Journal of Neuroscience, 8, 1531–1568.

    Google Scholar 

  • Trevarthen, C., & Sperry, R. W. (1968). Two mechanisms of vision in primates. Psychological Research, 31, 299–337.

    Article  Google Scholar 

  • Tsien, J. Z. (2007). The organizing principles of real-time memory encoding: Neural clique assemblies and universal neural codes. In B. Bontempi, A. Silva, & Y. Christen (Eds.), Memories: Molecules and circuits (pp. 100–182). Berlin: Springer.

    Google Scholar 

  • Turing, A. (1936) On computable numbers, with an application to the Entscheidungsproblem. Proceedings of the London Mathematical Society, 42, 230–265.

    Google Scholar 

  • Umarova, R. M., Saur, D., Schnell, S., Kaller, C. P., Vry, M. S., Glauche, V., Rijntjes, M., Hennig, J., Kiselev, V., & Weiller, C. (2010). Structural connectivity for visuospatial attention: Significance of ventral pathways. Cerebral Cortex, 20, 121–129.

    Article  Google Scholar 

  • Umiltà, M., Kohler, E., Gallese, V., Fogassi, L., Fadiga, L., Keysers, C., & Rizzolatti, G. (2001). I know what you are doing – A neurophysiological study. Neuron, 31, 155–165.

    Article  Google Scholar 

  • Ungerleider, L., & Mishkin, M. (1982). Two cortical visual systems. In D. J. Ingle, M. A. Goodale, & R. J. W. Mansfield (Eds.), Analysis of visual behavior (pp. 549–586). Cambridge: MIT.

    Google Scholar 

  • Usher, M. (2004). Comment on Ryder’s SINBAD neurosemantics: Is teleofunction isomorphism the way to understand representations? Minds and Language, 19, 241–248.

    Article  Google Scholar 

  • Vanduffel, W., Tootell, R. B., Schoups, A. A., & Orban, G. A. (2002). The organization of orientation selectivity throughout the macaque visual cortex. Cerebral Cortex, 12, 647–662.

    Article  Google Scholar 

  • Van Essen, D. C. (2005). A population-average, landmark- and surface-based (PALS) atlas of human cerebral cortex. NeuroImage, 28, 635–662.

    Article  Google Scholar 

  • Van Essen, D. C., & DeYoe, E. A. (1994). Concurrent processing in the primate visual cortex. In M. S. Gazzaniga (Ed.), The cognitive neurosciences. Cambridge: MIT.

    Google Scholar 

  • Van Essen, D. C., Newsome, W., & Maunsell, J. (1984). The visual field representation in striate cortex of the macaque monkey: Asymmetries, anisotropies, and individual variability. Vision Research, 24, 429–448.

    Article  Google Scholar 

  • Van Essen, D. C., Lewis, J. W., Drury, H. A., Hadjikhani, N., Tootell, R. B., Bakircioglu, M., & Miller, M. I. (2001). Mapping visual cortex in monkeys and humans using surface-based atlases. Vision Research, 41, 1359–1378.

    Article  Google Scholar 

  • van Fraassen, B. C. (2008). Scientific representation. Oxford: Oxford University Press.

    Book  Google Scholar 

  • Van Hooser, S. D., Heimel, J. A. F., Chung, S., Nelson, S. B., & Toth, L. J. (2005). Orientation selectivity without orientation maps in visual cortex of a highly visual mammal. Journal of Neuroscience, 25, 19–28.

    Article  Google Scholar 

  • Verkindt, C., Bertrand, O., Echallier, F., & Pernier, J. (1995). Tonotopic organization of the human auditory cortex: N100 topography and multiple dipole model analysis. Electroencephalography and Clinical Neurophisiology, 96, 143–156.

    Article  Google Scholar 

  • von der Malsburg, C. (1973). Self-organization of orientation sensitive cells in the striate cortex. Kybernetic, 14, 85–100.

    Article  Google Scholar 

  • von der Malsburg, C. (1995a). Binding in models of perception and brain function. Current Opinion in Neurobiology, 5, 520–526.

    Article  Google Scholar 

  • von Economo, C., & Koskinas, G. N. (1925). Die Cytoarchitektonik der Hirnrinde des erwachsenen Menschen. Berlin: Springer.

    Google Scholar 

  • von Neumann, J. (1961). The general and logical theory of automata. In A. Taub (Ed.), Collected works (Vol. V, pp. 288–328). New York: Pergamon Press.

    Google Scholar 

  • Vuilleumier, P., Henson, R. N., Driver, J., & Dolan, R. J. (2002). Multiple levels of visual object constancy revealed by event-related fMRI of repetition priming. Nature Neuroscience, 5, 491–499.

    Article  Google Scholar 

  • Wandell, B. A. (1999). Computational neuroimaging of human visual cortex. Annual Review of Neuroscience, 10, 145–173.

    Article  Google Scholar 

  • Wandell, B. A., Brewer, A. A., & Dougher, R. F. (2005). Visual field map clusters in human cortex. Philosophical Transactions of the Royal Society of London, 360, 693–707.

    Article  Google Scholar 

  • Weigelt, S., Kourtzi, Z., Kohler, A., Singer, W., & Muckli, L. (2007). The cortical representation of objects rotating in depth. Journal of Neuroscience, 27, 3864–3874.

    Article  Google Scholar 

  • Wiesel, T., & Hubel, D. (1965). Binocular interaction in striate cortex of kittens reared with artificial squint. Journal of Neurophysiology, 28, 1041–1059.

    Google Scholar 

  • Williams, D., Sagness, K., & McPhee, J. (1994). Configural and elemental strategies in predictive learning. Journal of Experimental Psychology, Learning, Memory, and Cognition, 20, 694–709.

    Article  Google Scholar 

  • Winer, J. A., Miller, L. M., Lee, C. C., & Schreiner, C. E. (2005). Auditory thalamocortical transformation: Structure and function. Neuron, 28, 255–263.

    Google Scholar 

  • Wittgenstein, L. (1953). Philosophische Untersuchung. Oxford: Basil Blackwell.

    Google Scholar 

  • Wolfram, S. (2002). A new kind of science. Champaign: Wolfram Media.

    Google Scholar 

  • Wunderlich, K., Symmonds, M., Bossaerts, P., & Dolan, R. J. (2011). Hedging your bets by learning reward correlations in the human brain. Neuron, 71, 1141–1152.

    Article  Google Scholar 

  • Xu, Y. (2005). Revisiting the role of the fusiform face area in visual expertise. Cerebral Cortex, 15, 1234–1242.

    Article  Google Scholar 

  • Zeki, S. (1971). Functional organization of a visual area in the posterior bank of the superior temporal sulcus of the rhesus monkey. Journal of Physiology, 236, 549–573.

    Article  Google Scholar 

  • Zeki, S. (1983a). Colour coding in the cerebral cortex: The reaction of cells in monkey visual cortex to wavelenghts and colours. Neuroscience, 9, 741–765.

    Article  Google Scholar 

  • Zeki, S. (1983b). Colour coding in the cerebral cortex: The responses of wavelength-selective and colour-coded cells in monkey visual cortex to changes in wavelength composition. Neuroscience, 9, 767–781.

    Article  Google Scholar 

  • Zeki, S. (2003). Improbable areas in the visual brain. Trends in Neuroscience, 26, 23–26.

    Article  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Rights and permissions

Reprints and permissions

Copyright information

© 2016 Springer International Publishing Switzerland

About this chapter

Cite this chapter

Plebe, A., De La Cruz, V.M. (2016). Representational Mechanisms. In: Neurosemantics. Studies in Brain and Mind, vol 10. Springer, Cham. https://doi.org/10.1007/978-3-319-28552-8_3

Download citation

Publish with us

Policies and ethics