Skip to main content
Log in

The relationship between psychological capacities and neurobiological activities

  • Original Paper in Philosophy of Science
  • Published:
European Journal for Philosophy of Science Aims and scope Submit manuscript

Abstract

This paper addresses the relationship between psychological capacities, as they are understood within cognitive psychology, and neurobiological activities. First, Lycan’s (1987) account of this relationship is examined and certain problems with his account are explained. According to Lycan, psychological capacities occupy a higher level than neurobiological activities in a hierarchy of levels of nature, and psychological entities can be decomposed into neurobiological entities. After discussing some problems with Lycan’s account, a similar, more recent account built around levels of mechanisms is examined (Craver 2007). In the second half of this paper, an alternative is laid out. This new account uses levels of organization and levels of explanation to create a two-dimensional model. Psychological capacities occupy a high level of explanation relative to the cellular and molecular levels of organization. As a result, according to this model, psychological capacities are a particular way of describing the activities that occur at the cellular and molecular levels of organization.

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

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11

Similar content being viewed by others

Notes

  1. This account is similar in some respects to one developed by Dennett (1978), to whom Lycan attributes the basic idea.

  2. The reference to evolutionary processes is not defended by Lycan, but it does position his view within a certain context. He says initially, “I hope, and am inclined to believe, that the teleological characterizations that Homunctionalism [i.e., homuncular functionalism] requires can be independently explained in evolutionary terms” (1987, p. 43). And later, “If teleological characterizations are themselves explicated in evolutionary terms, then our capacities for mental states themselves become more readily explicable by final cause; it is more obvious why we have pains, beliefs, desires, and so on” (1987, p. 45). To illustrate, he offers, “Why does pain hurt? Why could we not have a damage-signaling and repair-instigating system that was not uncomfortable? The answer is simple. Suppose I had just such a system, like the red warning light on my auto engine. Just as I habitually though irrationally ignore the warning light and vaguely hope it will go away, I would ignore a personal warning light if it did not intrinsically provide me with an urgent motive to do something about it” (1987, p. 138, n. 17).

  3. Lycan does not provide a very thorough explanation of the significant features of a level of nature. His definition is “levels [of nature] are nexus of interesting lawlike generalizations, and are individuated according to the types of generalizations involved” (1987, p. 38). Referring to “nexus of interesting lawlike generalizations” does not appear to be (and may not be intended as) a precise criterion. In any case, this definition presumably allows Lycan to draw on anything that falls within the purview of science.

  4. Lycan uses an example to illustrate this notion of degrees of teleologicalness that is discussed in the next section of this paper (see Fig. 2). In that example, as we move down this hierarchy created by homuncular functionalism from a face recognizer to an analyzer to a scanner to a light meter and to photosensitive chemicals, each characterization is, Lycan suggests, less teleological than the previous one.

  5. Lycan locates psychological capacities just below the level of the organism itself (the organism is an “institution” containing the psychological capacities (1987, 40)), but several levels above the neurobiological. To locate the psychological directly at the neuroanatomical level is, he says, “implausible” (1987, p. 59).

  6. Ultimately, the distinction I am after is between purely functional, or nearly purely functional, and physical. This distinction is complicated, however, by the assumption that everything is physical. Thus, to start out, I will try to be more precise by using relational and non-relational (or intrinsic). For example, my belief that the sky is gray has only relational properties—so far as we understand the belief anyway. A neuron in my brain has some relational properties, but it also has intrinsic properties, for instance, being a neuron, as well as having length, width, and mass. And these intrinsic properties are also physical properties. An entity that is only known to have relational properties will not have any (known) physical properties.

  7. A different interpretation of Cummins’ account of functional analysis is offered by Piccinini and Craver (2011). They suggest that a functional decomposition is not explanatorily useful unless it does, at least implicitly, refer to a physical system.

  8. Stepping back for a moment, the reason we are interested in neurobiological activities at all, rather than just neurobiological entities, is because psychological capacities are processes. They are temporally extended, and they typically transform an input into an output. Thus, in order to establish the relationship between psychological capacities and something neurobiological, we have to focus on neurobiological activities, not neurobiological structures.

  9. Craver’s levels of mechanisms and Wimsatt’s levels of organization have some similarities, most notably, an emphasis on causal interactions. I won’t review the differences between the two positions here, but see Craver (2007, chapter 5) for his criticisms of Wimsatt.

  10. For more information on methodological functionalism, see Polger (2004, 2009).

  11. Marr does not make a distinction between the terms explanation and description. I will refer to descriptions that occupy levels of explanation. Also, although it will not have any bearing on what we do here, it is useful to note that there are other types of descriptions besides the three that Marr uses. (Statistical descriptions are one example.) Consequently, there can be other types of levels of explanation in addition to Marr’s. And levels of explanation do not only apply to mental processes, but are used any time a higher level description is provided for a process.

    It is also useful to note that while Marr speaks of descriptions of information processing tasks, this type of task is essentially the same as a capacity. A capacity is just the ability to perform a certain task.

  12. It has been pointed out that computational theory is a confusing label for this level because it is not where a computational operation is described (Bechtel 1994; Bechtel et al. 1998). As Bechtel et al. say, “[Marr] called his highest level computational theory (a label that many have found misleading; it is somewhat akin to Chomsky’s notion of competence and might best be called task analysis)” (1998, p. 65).

  13. The descriptions that are in the quotations from Scherer and Robinson are brief; a complete description at this level can be quite substantial if the capacity is characterized in detail. See Scherer (2001) and Robinson (2004, 2005) for complete descriptions of their theories of emotion. Using Scherer’s and Robinson’s descriptions is complicated by the fact that, while Scherer is a cognitive psychologist, Robinson is a philosopher. Nonetheless, a little blurring between fields is acceptable. And these two quotations are a good and concise illustration of Marr’s highest level.

  14. This is a substantial topic. Nevertheless, there are plenty of examples of neuroscientific investigations into how psychological processes are carried out at the cellular level. See, for example, Graziano’s work on microstimulation of the neurons in the motor cortex (Graziano et al. 2002ab); Rolls’ investigation of the responses of neurons in the orbitofrontal cortex to certain kinds of visual information (Rolls et al. 2005, 2006); or the progress that is being made on the control of prosthetic devices by cellular activity (Carmena et al. 2003; Tillery and Taylor 2004; Velliste et al. 2008).

    It is also interesting to note that the descriptions that Marr offered for the lowest level of explanation all involved activities at the cellular level of organization. The visual processes that he investigated are carried out, he suggested, by the activities of the neurons in the retina, lateral geniculate nucleus, and primary visual cortex (Marr 1982; Marr and Ullman 1981).

  15. There are other processes that produce plasticity and probably occur in the neurons in the hippocampus. Long-term depression, spike-timing-dependent plasticity, and depotentiation are a few (Neves et al. 2008, p. 66). These other means of plasticity won’t be discussed here, but they could be included in the descriptions that reside at the molecular level of organization (i.e., the lowest level of explanation that occurs at the molecular level of organization). But notice that the descriptions of molecular activities need only be those that are relevant to explaining spatial memory. Of course, “explanatorily relevant” can be widely interpreted, but, on the face of it, there does not seem to be a need to provide a molecular level explanation of every aspect of the activity that occurs at the cellular level in order accurately describe spatial memory.

  16. Some changes have been made to Craver’s highest level (his “level of spatial memory”), and that description has become the highest level of explanation. Also, one of Craver’s mechanistic levels has been eliminated. Besides a level for the cellular activities, he includes a level for LTP and then a separate level for the molecular process that carries out LTP. Dropping one of the latter two descriptions seems warranted.

References

  • Bechtel, W. (1994). Levels of description and explanation in cognitive science. Minds and Machines, 4, 1–25.

    Article  Google Scholar 

  • Bechtel, W., Abrahamsen, A., & Graham, G. (1998). The life of cognitive science. In W. Bechtel, & G. Graham (Eds.), A companion to cognitive science (pp. 2–104). Malden, MA: Blackwell.

    Google Scholar 

  • Bickle, J. (2003). Philosophy and neuroscience: A ruthlessly reductive account. Boston: Kluwer.

    Book  Google Scholar 

  • Bourne, J. N., & Harris, K. M. (2011). Coordination of size and number of excitatory and inhibitory synapses results in a balanced structural plasticity along mature hippocampal CA1 dendrites during LTP. Hippocampus, 21, 354–373.

    Article  Google Scholar 

  • Carmena, J. M., Lebedev, M. A., Crist, R. E., O’Doherty, J. E., Santucci, D. M., Dimitrov, D. F., et al. (2003). Learning to control a brain–machine interface for reaching and grasping by primates. PLoS Biology, 1, 193–208.

    Article  Google Scholar 

  • Churchland, P. S., & Sejnowski, T. J. (1988). Perspectives on cognitive neuroscience. Science, 242, 741–745.

    Article  Google Scholar 

  • Churchland, P. S., & Sejnowski, T. J. (1992). The computational brain. Cambridge, MA: MIT Press.

    Google Scholar 

  • Craver, C. F. (2002). Interlevel experiments and multilevel mechanisms in the neuroscience of memory. Philosophy of Science, 69(Proceedings), S83–S97.

    Article  Google Scholar 

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

    Google Scholar 

  • Cummins, R. (1975). Functional analysis. The Journal of Philosophy, 72, 741–765.

    Article  Google Scholar 

  • Cummins, R. (1983). The nature of psychological explanation. Cambridge, MA: MIT Press.

    Google Scholar 

  • Cummins, R. (2000). “How does it work?” versus “what are the laws?”: Two conceptions of psychological explanation. In F. C. Keil, & R. A. Wilson (Eds.), Explanation and cognition (pp. 117–144). Cambridge, MA: MIT Press.

    Google Scholar 

  • Dennett, D. C. (1978). Brainstorms: Philosophical essays on mind and psychology. Montgomery, VT: Bradford.

    Google Scholar 

  • Graziano, M. S. A., Taylor, C. S. R., & Moore, T. (2002a). Complex movements evoked by microstimulation of precentral cortex. Neuron, 34, 841–851.

    Article  Google Scholar 

  • Graziano, M. S. A., Taylor, C. S. R., Moore, T., & Cooke, D. F. (2002b). The cortical control of movement revisited. Neuron, 36, 349–362.

    Article  Google Scholar 

  • Johnson, G. (2009). Mechanisms and functional brain areas. Minds and Machines, 19, 255–271.

    Article  Google Scholar 

  • Kandel, E. R. (2001). The molecular biology of memory storage: A dialogue between genes and synapses. Science, 294, 1030–1038.

    Article  Google Scholar 

  • Lycan, W. G. (1981). Form, function, and feel. The Journal of Philosophy, 78, 24–50.

    Article  Google Scholar 

  • Lycan, W. G. (1987). Consciousness. Cambridge, MA: MIT Press.

    Google Scholar 

  • Lycan, W. G. (1991). Homuncular functionalism meets PDP. In W. Ramsey, S. P. Stich, & D. E. Rumelhart (Eds.), Philosophy and connectionist theory (pp. 259–286). Hillsdale, NJ: Lawrence Erlbaum.

    Google Scholar 

  • Marr, D. (1977). Artificial intelligence–A personal view. Artificial Intelligence, 9, 37–48.

    Article  Google Scholar 

  • Marr, D. (1982). Vision: A computational investigation into the human representation and processing of visual information. San Francisco: W.H. Freeman.

    Google Scholar 

  • Marr, D., & Ullman, S. (1981). Directional selectivity and its use in early visual processing. Proceedings of the Royal Society of London B, 211, 151–180.

    Article  Google Scholar 

  • Mou, W., & McNamara, T. P. (2002). Intrinsic frames of reference in spatial memory. Journal of Experimental Psychology: Learning, Memory, and Cognition, 28, 162–170.

    Article  Google Scholar 

  • Neves, G., Cooke, S. F., & Bliss, T. V. P. (2008). Synaptic plasticity, memory and the hippocampus: A neural network approach to causality. Nature Reviews Neuroscience, 9, 65–75.

    Article  Google Scholar 

  • Piccinini, G., & Craver, C. (2011). Integrating psychology and neuroscience: Functional analyses as mechanism sketches. Synthese, 183, 283–311.

    Article  Google Scholar 

  • Polger, T. W. (2004). Natural minds. Cambridge, MA: MIT Press.

    Google Scholar 

  • Polger, T. W. (2009). Computational functionalism. In J. Symons, & P. Calvo (Eds.), Routledge companion to the philosophy of psychology (pp. 148–163). New York: Routledge.

    Google Scholar 

  • Robinson, J. (2004). Emotion: Biological fact or social construction? In Robert C. Solomon (Ed.), Thinking about feeling: Contemporary philosophers on emotion (pp. 28–43). New York: Oxford University Press.

    Google Scholar 

  • Robinson, J. (2005). Deeper than reason: Emotion and its role in literature, music, and art. Oxford: Oxford University Press.

    Google Scholar 

  • Rock, I. (1973). Orientation and form. New York: Academic Press.

    Google Scholar 

  • Rolls, E. T., Browning, A. S., Inoue, K., & Hernadi, I. (2005). Novel visual stimuli activate a population of neurons in the primate orbitofrontal cortex. Neurobiology of Learning and Memory, 84, 111–123.

    Article  Google Scholar 

  • Rolls, E. T., Critchley, H. D., Browning, A. S., & Inoue, K. (2006). Face-selective and auditory neurons in the primate orbitofrontal cortex. Experimental Brain Research, 170, 74–87.

    Article  Google Scholar 

  • Scherer, K. R. (1997). Profiles of emotion-antecedent appraisal: Testing theoretical predictions across cultures. Cognition & Emotion, 11, 113–150.

    Article  Google Scholar 

  • Scherer, K. R. (2001). Appraisal considered as a process of multilevel sequential checking. In K. R. Scherer, A. Schorr, & T. Johnstone (Eds.), Appraisal processes in emotion: Theory, methods, research (pp. 92–120). New York: Oxford University Press.

    Google Scholar 

  • Shelton, A. L., & McNamara, T. P. (2001). Systems of spatial reference in human memory. Cognitive Psychology, 43, 274–310.

    Article  Google Scholar 

  • Smith, C. A., & Kirby, L. D. (2000). Consequences require antecedents: Toward a process model of emotion elicitation. In J. P. Forgas (Ed.), Feeling and thinking: The role of affect in social cogition (pp. 83–106). Cambridge, U.K.: Cambridge University Press.

    Google Scholar 

  • Squire, L. R., & Kandel, E. R. (1999). Memory: From mind to molecules. New York: Scientific American Library.

    Google Scholar 

  • Tillery, S. I. H., & Taylor, D. M. (2004). Signal acquisition and analysis for cortical control of neuroprosthetics. Current Opinion in Neurobiology, 14, 758–762.

    Article  Google Scholar 

  • Velliste, M., Perel, S., Spalding, M. C., Whitford, A. S., & Schwartz, A. B. (2008). Cortical control of a prosthetic arm for self-feeding. Nature, 453, 1098–1101.

    Article  Google Scholar 

  • Wimsatt, W. C. (1976). Reductionism, levels of organization, and the mind-body problem. In G. G. Globus, M. Grover, & I. Savodnik (Eds.), Consciousness and the brain: A scientific and philosophical inquiry (pp. 205–267). New York: Plenum Press.

    Google Scholar 

  • Wimsatt, W. C. (1994). The ontology of complex systems: Levels of organization, perspectives, and causal thickets. In M. Matthen & R. Ware (Eds.), Biology and society: Reflections on methodology (Supplementary vol. 20 of the Canadian Journal of Philosophy, pp. 207–274). Calgary: University of Calgary Press.

  • Wimsatt, W. C. (2007). Re-engineering philosophy for limited beings: Piecewise approximations to reality. Cambridge, MA: Harvard University Press.

    Google Scholar 

Download references

Acknowledgements

I am grateful for the help that I received from Thomas Polger and John Bickle on the many drafts of this paper. I am also indebted to two anonymous reviewers for this journal who provided very helpful comments on the paper’s final versions.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Gregory Johnson.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Johnson, G. The relationship between psychological capacities and neurobiological activities. Euro Jnl Phil Sci 2, 453–480 (2012). https://doi.org/10.1007/s13194-012-0053-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s13194-012-0053-y

Keywords

Navigation