Skip to main content
Log in

Theoretical considerations on cognitive niche construction

  • Published:
Synthese Aims and scope Submit manuscript

Abstract

Cognitive niche theories consist in a theoretical framework that is proving extremely profitable in bridging evolutionary biology, philosophy, cognitive science, and anthropology by offering an inter-disciplinary ground, laden with novel approaches and debates. At the same time, cognitive niche theories are multiple, and differently related to niche theories in theoretical and evolutionary biology. The aim of this paper is to clarify the theoretical and epistemological relationships between cognitive and ecological niche theories. Also, by adopting a constructionist approach (namely by referring principally to ecological and cognitive niche construction theories) we will try to explain the shift from ecological to cognitive niches and their actual and theoretical overlaps. In order to do so, we will take two concepts expressing loose forms of causation in the interaction between organisms and their environment: the biological notion of “enablement” and the psycho-cognitive one of “affordance”.

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

Similar content being viewed by others

Notes

  1. Cf. from Sect. 4.2 onwards.

  2. When reading this quote, one should still keep in mind that Gibson was a psychologist and not an evolutionary biologist.

  3. “The bark of trees is part of the woodpecker’s environment, but the stones at the base of the tree, even though physically present, are not. [...] Not only do organisms determine their own food, but they make their own climate” (Levins and Lewontin 1985, p. 98).

  4. “Organisms are both the consumers and the producers of the resources necessary to their own continued existence. Plant roots alter the physical structure and chemical composition of the soil in which they grow, withdrawing nutrients but also conditioning the soil so that nutrients are more easily mobilized. Grazing animals actually increase the rate of production of forage, both by fertilizing the ground with their droppings and by stimulating plant growth by cropping. Organisms also influence the species composition of the plant community on which they depend” (Levins and Lewontin 1985, p. 99).

  5. Similar arguments seem to elicit the idea that the construction of an ecological niche is not only external but also internal in the sense of psycho-cognitive construction. This view is nevertheless opposed by Odling-Smee and colleagues who, in their definition of ecological niche construction, clearly stressed that their “use of the term construction refers to a physical modification of the selective environment or actual movement in physical space, and not to the perceptual processes responsible for construction a mental representation of the world from sensory inputs (as used in psychology), not to the creation of scientific theories or facts (as used by postmodernist philosophers)” (Odling-Smee et al. 2003, p. 41, footnote 4).

  6. “Both the amplitude and the frequency of external fluctuations are transformed by biological processes in the organism. Fluctuations are damped by various storage devices that average over space and time” (Levins and Lewontin 1985, p. 100).

  7. Magnani does not deny the fact that cognitive niche construction signifies a jump to the next curb in cognition, but instead of conceiving the cognitive niche as something to be entered he concentrates on the cognitive evolution of human beings into masters of ecological-cognitive engineering: “Accordingly, we may argue that the creation of cognitive niches is the way cognition evolves, and humans can be considered as ecological cognitive engineers” (Magnani 2009, p. 332).

  8. Cf. p. 3 and footnote 5 earlier in this article.

  9. This overlap is evident and should not be taken as a sign of theoretical inconsistency: it will be dealt with toward the conclusion of this article, from page 20.

  10. In agreement with Odling-Smee et al. (2003, pp. 256–257), the concept of fitness here has to be intended as loosely Darwinian because of the following reasons: extragenetically informed behavior patterns are broadly adaptive and maladaptive; variants occurring during genetic evolution are random, whereas those of extra-genetic information are not. It is also consistent with Pinker’s take on the cognitive niche, for which the gathering and exchange of information is crucial for survival (Pinker 2003).

  11. As far as the biological and pre-linguistic levels are concerned, it can be argued that the detected cause–effect relationships do not matter for their truth-reliability but rather for their fitness-reliability (Sage 2004)—understanding “fitness” as comprehending the less rigorous notion of “welfare”. While our human language-dominated world informs the fact that we are used to consider the notion of truth, naïvely, as correspondence, from a biological perspective (which is often engaged by human beings as well) the favored detected relationship is the most successful, the one leading to survival.

  12. Suggesting that organisms may affect evolutionary dynamics by modifying their environment and thus enacting another form of inheritance may actually bridge non-Lamarckian and neo-Lamarckian insights. This discourse connects with Jablonka and Lamb’s arguments for epi-genetic inheritance, namely the possibility that genetic inheritance is affected by the way “variants are inherited and what final form they assume depends on various ‘filtering’ and ‘editing’ processes that occur before and during the transmission” (Jablonka and Lamb 2005, p. 309), that is by how cells regulate the activation and deactivation of specific genetic material.

  13. This view is also embedded in Wallach’s evolutionary argument against niche construction, where he denies the theory’s explanatory power inasmuch as it cannot give a causal account of how similar situations developed very different outcomes (Wallach 2015).

  14. Civilization is a fortunate series of role-based strategy video-games, in which players are prompted to take one action (where in their possessions or against their adversaries) during each turn, then awaiting the reaction or another move by the other players or the computer.

  15. Almost ironically, such a view that breaks the dichotomy between what causes and what is caused turns groundbreaking insights such as the existence of an ecological inheritance system into a matter of fact. Indeed, given the enabled nature of all biological phenomena, and considering the constant reference to the environment as a constraint life develops on, around, within, and affecting it at the same time (and thus affecting its enabling capacities), the genetic and the ecological inheritance systems blend into one both at the ontogenetic and phylogenetic level. Furthermore, understanding a niche as the result of a set of interrelated and modifiable enablements makes better sense of how the direction of niche construction activities is not limited to \(organism \longrightarrow inactive \ environment\) (as \( beaver \longrightarrow river \& wood\) or \(termites \longrightarrow earth mound\)) but allows many instances of \(organism \longrightarrow organism\), since organisms can be each other’s environment inasmuch as they consist in a precise pool of mutual enablements.

  16. Affordance theory is extensively employed in design, technology, and communication studies (Norman 1988b; Withagen et al. 2012; Nagy and Neff 2015). It is are also used by law and government scholars (Calo 2016). Hildebrandt (2015) bases part of her analysis of the law on the “affordances of written text.”

  17. “The different substances of the environment have different affordances for nutrition and for manufacture. The different objects of the environment have different affordances for manipulation. The other animal afford, above all, a rich and complex set of interactions, sexual, predatory, nurturing, fighting, playing, cooperating, and communicating. What other persons afford, comprises the whole realm of social significance for human beings” (Gibson 1979, p. 128).

  18. The use of “man” as a general word indicating female and male human beings is now outdated. Gibson, writing at the end of the Seventies, was of course not arguing that only male human beings were engaged in environmental modification activities.

  19. Moreover, according to Reed, the opinion that mediated perception or cognition, hence learning, is inconsistent with Gibson’s view of ecological psychology is “simply mistaken” (Reed 1988, p. 305), as clearly illustrated by Gibson: “At least three separate levels [of theorising] will be required: first, a theory of how we perceive the surfaces of objects [...]; second, a theory of how we perceive representations, pictures, displays, and diagrams; and third, a theory of how we apprehend symbols. There is no reason to suppose that the physiological concomitants of all these experiences will be the same; in fact, since pictures and symbols presuppose objects, their physiological explanations will probably have to be found at increasing levels of complexity” (Gibson 1951, p. 413).

  20. Cf. p. 10 earlier in this paper.

  21. Cf. p. 10 earlier in this article.

References

  • Bertolotti, T., & Cinerari, C. (2013). The gospel according to Google: The future of religious niches and technological spirituality. European Journal of Science and Theology, 9(1), 41–53.

    Google Scholar 

  • Bertolotti, T., & Magnani, L. (2015). Contemporary finance as a critical cognitive niche. Mind & Society, 14(2), 273–293.

    Article  Google Scholar 

  • Calo, R. (2016). Can Americans resist surveillance? The University of Chicago Law Review, 83, 23–46.

    Google Scholar 

  • Clark, A. (1997). Being there: Putting brain, body, and world together again. Cambridge, MA: The MIT Press.

    Google Scholar 

  • Clark, A. (2005). Word, niche and super-niche: How language makes minds matter more. Theoria, 54, 255–268.

    Google Scholar 

  • Clark, A. (2006). Language, embodiment, and the cognitive niche. Trends in Cognitive Science, 10(8), 370–374.

    Article  Google Scholar 

  • Clark, A., & Chalmers, D. J. (1998). The extended mind. Analysis, 58, 10–23.

    Article  Google Scholar 

  • Day, R. L., Laland, K., & Odling-Smee, F. J. (2003). Rethinking adaptation. The niche-construction perspective. Perspectives in Biology and Medicine, 46(1), 80–95.

    Article  Google Scholar 

  • Elton, C. S. (1927). Animal ecology. New York: Macmillan.

    Google Scholar 

  • Gibson, J. J. (1951). What is a form? Psychological Review, 58, 403–413.

    Article  Google Scholar 

  • Gibson, J. J. (1979). The ecological approach to visual perception. Boston: Houghton Mifflin.

    Google Scholar 

  • Godfrey-Smith, P. (1998). Complexity and the function of mind in nature. Cambridge: Cambridge University Press.

    Google Scholar 

  • Griesemer, J. (2000). Development, culture, and the units of inheritance. Philosophy of Science, 67, S348–S368.

    Article  Google Scholar 

  • Grinnell, J., & Swarth, H. S. (1913). An account of the birds and mammals of the San Jacinto area of southern California with remarks upon the behavior of geographic races on the margins of their habitats. Berkeley: University of California Press.

    Book  Google Scholar 

  • Hildebrandt, M. (2015). Smart technologies and the end(s) of law: Novel entanglements of law and technology. Cheltenham, UK: Edward Elgar.

    Book  Google Scholar 

  • Hutchinson, G. E. (1957). Concluding remarks. Cold Spring Harbor Symposia on Quantitative Biology, 22, 415–427.

    Article  Google Scholar 

  • Iriki, A., & Taoka, M. (2012). Triadic (ecological, neural, cognitive) niche construction: A scenario of human brain evolution extrapolating tool use and language from the control of reaching actions. Philosophical Transactions of the Royal Society B, 367, 10–23.

    Article  Google Scholar 

  • Jablonka, E., & Lamb, M. J. (2005). Evolution in four dimensions. Genetic, epigenetic, behavioral, and symbolic variation in the history of life. Cambridge, MA: MIT Press.

  • Laland, K. N., Odling-Smee, F. J., & Feldman, M. W. (2000). Niche construction, biological evolution and cultural change. Behavioral and Brain Sciences, 23(1), 131–175.

    Article  Google Scholar 

  • Levins, R., & Lewontin, R. C. (1985). The dialectical biologist. Harvard: Harvard University Press.

    Google Scholar 

  • Lewontin, R. C. (1983). Gene, organism and environment. In D. S. Bendall (Ed.), Evolution from molecules to men. Cambridge: Cambridge University Press.

    Google Scholar 

  • Longo, G., & Montévil, M. (2014). Perspectives on organisms. Berlin: Springer.

    Book  Google Scholar 

  • Longo, G., Montévil, M., & Kauffman, S. (2012). No entailing laws, but enablement in the evolution of the biosphere. GECCO companion ’12 proceedings of the fourteenth international conference on genetic and evolutionary computation conference companion (pp. 1379–1392). New York: ACM.

    Chapter  Google Scholar 

  • MacArthur, R., & Levins, R. (1967). The limiting similarity, convergence, and divergence of coexisting species. American Naturalist, 101, 377–385.

    Article  Google Scholar 

  • Magnani, L. (2007). Creating chances through niche construction. The role of affordances. In B. Apolloni (Ed.), Knowledge-based intelligent information and engineering systems: 11th International conference, KES 2007, proceedings, Part II, Vietri sul Mare, Italy, September 12–14, 2007. Lecture notes in computer science. Berlin: Springer.

  • Magnani, L. (2009). Abductive cognition: The epistemological and eco-cognitive dimensions of hypothetical reasoning. Berlin: Springer.

    Book  Google Scholar 

  • Magnani, L., & Bardone, E. (2008). Sharing representations and creating chances through cognitive niche construction. The role of affordances and abduction. In S. Iwata, Y. Oshawa, S. Tsumoto, N. Zhong, Y. Shi, & L. Magnani (Eds.), Communications and discoveries from multidisciplinary data (pp. 3–40). Berlin: Springer.

  • Menary, R. (Ed.). (2010). The extended mind. Cambridge, MA: The MIT Press.

    Google Scholar 

  • Nagy, P., & Neff, G. (2015). Imagined affordance: Reconstructing a keyword for communication theory. Social Media & Society, 1(2), 1–9.

  • Norman, D. A. (1988a). The design of everyday things. New York: Addison Wesley.

    Google Scholar 

  • Norman, D. A. (1988b). The psychology of everyday things. New York: Basic Books.

    Google Scholar 

  • Odling-Smee, F. J., Laland, K. N., & Feldman, M. W. (2003). Niche construction. The neglected process in evolution. Princeton, NJ: Princeton University Press.

    Google Scholar 

  • Pinker, S. (2003). Language as an adaptation to the cognitive niche. In M. H. Christiansen & S. Kirby (Eds.), Language evolution (pp. 16–37). Oxford: Oxford University Press.

    Chapter  Google Scholar 

  • Pocheville, A. (2015). The ecological niche: History and recent controversies. In T. Heams, P. Huneman, G. Lecointre, & M. Silberstein (Eds.), Handbook of evolutionary thinking in the sciences (pp. 547–586). New York: Springer.

    Google Scholar 

  • Reed, E. S. (1988). James J. Gibson and the psychology of perception. New Haven, CT: Yale University Press.

    Google Scholar 

  • Sage, J. (2004). Truth-reliability and the evolution of human cognitive faculties. Philosophical Studies, 117, 95–106.

    Article  Google Scholar 

  • Shavit, A., & Griesemer, J. (2011). Mind the gaps: Why are niche construction models so rarely used? In S. B. Gissis & E. Jablonka (Eds.), Transformations of Lamarckism (pp. 307–317). Cambridge, MA: The MIT Press.

    Chapter  Google Scholar 

  • Sinha, C. (2015). Language and other artifacts: Socio-cultural dynamics of niche construction. Frontiers in Psychology, 6(1601). doi:10.3389/fpsyg.2015.01601.

  • Sinha, C. (2015). Ontogenesis, semiosis and the epigenetic dynamics of biocultural niche construction. Cognitive Development Online First. doi:10.1016/j.cogdev.2015.09.006.

  • Tooby, J., & DeVore, I. (1987). The reconstruction of hominid behavioral evolution through strategic modeling. In W. G. Kinzey (Ed.), Primate models of hominid behavior (pp. 183–237). Albany: Suny Press.

    Google Scholar 

  • Wallach, E. (2015). Niche construction theory as an explanatory framework for human phenomena. Synthese Online First. doi:10.1007/s11229-015-0868-0.

  • Weir, A. A. S., & Kacelnik, A. (2006). A new Caledonian crow (Corvus moneduloides) creatively re-designs tools by bending or unbending aluminium strips. Animal Cognition, 9, 317–334.

    Article  Google Scholar 

  • Withagen, R., de Poel, H. J., Araújo, D., & Pepping, G. J. (2012). Affordances can invite behavior: Reconsidering the relationship between affordances and agency. New Ideas in Psychology, 30, 250–258.

    Article  Google Scholar 

Download references

Acknowledgments

Research for this article was supported by the 2012 National Research Grant (PRIN) “Models and Inferences in Science: Logical, Epistemological, and Cognitive Aspects” from the Italian Ministry of Education, University and Research (MIUR). The authors wish to express their gratitude to the two anonymous referees, whose insightful comments helped shaping the final version of this article.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Tommaso Bertolotti.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Bertolotti, T., Magnani, L. Theoretical considerations on cognitive niche construction. Synthese 194, 4757–4779 (2017). https://doi.org/10.1007/s11229-016-1165-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11229-016-1165-2

Keywords

Navigation