Skip to main content
Log in

Fitness: static or dynamic?

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

Abstract

The most consistent definition of fitness makes it a static property of organisms. However, this is not how fitness is used in many evolutionary models. In those models, fitness is permitted to vary with an organism’s circumstances. According to this second conception, fitness is dynamic. There is consequently tension between these two conceptions of fitness. One recently proposed solution suggests resorting to conditional properties. We argue, however, that this solution is unsatisfactory. Using a very simple model, we show that it can lead to incompatible fitness values and indecision about whether selection actually occurs.

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.

Similar content being viewed by others

Notes

  1. The claim that organismal fitness is a causal property can, of course, be denied. While we cannot here address the nuances of this “statisticalist” view, the challenges posed deserve serious scrutiny (cf. Otsuka, 2016; Walsh et al., 2017).

  2. See Beatty and Finsen (1989) for more about the so-called “operationalist fallacy.”

  3. For convergent thoughts, see Bourrat (2015, 2017) who proposes that natural selection (as opposed to other evolutionary processes) results from differences in intrinsic-invariable properties (i.e., static properties) that lead to differences in reproductive output.

  4. For details, see Equation 4 on p.862 of Pence and Ramsey (2013).

  5. Op. 860-861 of Pence and Ramsey (2013). What follows draws very heavily on Pence and Ramsey’s wording.

  6. The adjective “God’s-eye-view” is not intended as pejorative. It is illustrative rather than evaluative. There is nothing philosophically untoward about attempting to establish metaphysical foundations for evolutionary theorizing.

  7. Although studies involving trait types feature prominently, it is worth remembering that this contingent fact is a matter of “occupational necessity” designed to prevent the conflation of individual fitness (as a propensity or propensity-like property) with realized individual fitness. It can be argued that biologists aim to determine the fitness of token organisms, which can only be approximated by way of indirect inferences that rely on selectively relevant trait types. On this point, compare Sober (2013) and Pence and Ramsey (2015).

  8. It is worth noting that models for the evolution of altruism rely on a form of frequency-dependent selection.

  9. Pence and Ramsey are clearly aware of the difficulties raised in this section. They explicitly state, “Our Equation (4) is the density-independent, non-chaotic limit of this more sophisticated work [in adaptive dynamics]” (2013, p.863).

  10. Astute readers will recognize that this is a slightly modified version of Abrams’s (2009a) example involving carotenoids.

  11. See Brandon (pp.60-64, 1990) for details about, what he calls, the “selective environment.”

  12. Epistasis occurs when alleles at two or more genetic loci interact non-additively to determine the phenotype.

  13. Pleiotropy occurs when one genetic locus affects more than one phenotypic trait, which causes a genetic correlation.

  14. In other words, the fitness distributions for the character states are identical.

  15. This local interpretation or “fine-grained partitioning” is equivalent to situation in which there is a metapopulation consisting of two subpopulations of equal size and there is no migration or mutation.

  16. We do not mean to imply that Abrams is oblivious to this problem. See, especially, Abrams (2009b) for a suggested resolution to, what he calls, “the problem of the reference environment.”

  17. Op.20 (Abrams, 2014): “What matters to natural selection is not this or that organism’s particular circumstances and particular fate, but the sorts of conditions that individuals in the population are likely to encounter repeatedly.”

  18. It also threatens to blur the distinction between drift and natural selection, an outcome that runs counter to the intuitions of most biologists and philosophers (Brandon, 2005). If your mathematical expectations differ substantially, then so, too, will your views about which outcomes are due to selection (i.e., are adaptions) and drift (i.e., departure from expectation).

  19. Second scenario: Imagine a single environment in which sand color varies in patches whose average diameter is about a meter. The random assignment of pigeons to a background sand color is then determined by where pigeons happen to land. In either case, we suppose that the pigeons are not selective about sand color” (Abrams, 2013, p.295-296).

  20. “I argue for the stronger claim that it’s plausible that: [1] Different descriptions of a biological population sometimes pick out distinct effects, and [2] There are distinct causes producing these effects. This is a claim about causation rather than explanation. Causal relations exist independently of how the population is characterized, though our way of describing the population can, indeed, focus on one or another of these causal relations” (Abrams, 2013, p.295). Later in the same paper, he claims that: “My view implies that a researcher’s description is able to pick out different causal relationships, but allows that these causal relationships exist, realized by some the of same parts of the world, independent of the choice of causal relationships on which to focus” (p. 300).

  21. In an earlier paper, Abrams says the following: “Rather than an organism’s fitness literally changing when an effect C occurs, there is instead a relationship between conditional probabilities: a type’s fitness conditional on the occurrence of C is different from its fitness conditional on C’s non-occurrence” (2009a, p.493).

  22. On this point, Abrams (2013, p.296, emphasis added) says the following: “It should be clear, though, that if what we are concerned with is the evolution of the pigeon population over several generations […] then any relevant sense of fitness and selection must take into account probabilities of pigeons being found on light or dark sand when hawks are overhead.”

References

  • Abrams, M. (2007). Fitness and Propensity’s annulment? Biology and Philosophy, 22, 115–130.

    Article  Google Scholar 

  • Abrams, M. (2009a). Fitness ‘Kinematics’: Biological Function, Altruism, and Organism–Environment Development. Biology and Philosophy, 24, 487–504.

    Article  Google Scholar 

  • Abrams, M. (2009b). What determines biological fitness? The problem of the reference environment. Synthese, 166, 21–40.

    Article  Google Scholar 

  • Abrams, M. (2012). Measured, modeled, and causal conceptions of fitness. Frontiers in Genetics, 3, 196.

    Article  Google Scholar 

  • Abrams, M. (2013). Populations and pigeons: Prosaic pluralism about evolutionary causes. Studies in History and Philosophy of Biological and Biomedical Sciences, 44, 294–301.

    Article  Google Scholar 

  • Abrams, M. (2014). Environmental grain, organism fitness, and type fitness. In G. Barker, E. Dejardins, & T. Pearce (Eds.), Entangled life: Organism and environment in the biological and social sciences. Springer.

    Google Scholar 

  • Ariew, A., & Ernst, Z. (2009). What fitness Can’t be. Erkenntnis, 71(3), 289–301.

    Article  Google Scholar 

  • Ariew, A., & Lewontin, R. C. (2004). The confusions of fitness. The British Journal for the Philosophy of Science, 55(2), 347–363.

    Article  Google Scholar 

  • Beatty, J.H. and Finsen S.K. (1989). Rethinking the Propensity Interpretation: A Peek inside Pandora’s Box, in What the Philosophy of Biology Is: Essays Dedicated to David Hull, ed. Michael Ruse (Dordrecht: Kluwer Publishers), pp. 17–30.

  • Bouchard, F., & Rosenberg, A. (2004). Fitness, probability and the principles of natural selection. British Journal for the Philosophy of Science, 55, 693–712.

    Article  Google Scholar 

  • Bourrat, P. (2015). Distinguishing natural selection from other evolutionary processes in the evolution of altruism. Biological Theory, 10, 311–321.

    Article  Google Scholar 

  • Bourrat, P. (2017). Explaining drift from a deterministic setting. Biological Theory, 12, 27–38.

    Article  Google Scholar 

  • Bourrat, P 2019. Natural selection and the reference grain problem”. Studies in History and Philosophy of Science Part A. https://doi.org/10.1016/j.shpsa.2019.03.003

  • Bourrat, P 2021. Facts, conventions, and the levels of selection. Elements in the Philosophy of Biology. Cambridge: Cambridge University Press, 2021. https://doi.org/10.1017/9781108885812

  • Brandon, R. N. (1978). Adaptation and evolutionary theory. Studies in History and Philosophy of Science Part A, 9, 181–206.

    Article  Google Scholar 

  • Brandon, R. N. (1990). Adaptation and environment. Princeton University Press.

    Google Scholar 

  • Brandon, R. N. (2005). The difference between selection and drift: A reply to Millstein. Biology and Philosophy, 20, 153–170.

    Article  Google Scholar 

  • Brandon, R. N., & Beatty, J. H. (1984). The propensity interpretation of ‘fitness’—No interpretation is no substitute. Philosophy of Science, 51, 342–347.

    Article  Google Scholar 

  • Charbonneau, M., & Bourrat, P. (2021). Fidelity and the grain problem in cultural evolution. Synthese. https://doi.org/10.1007/s11229-021-03047-1

  • Doulcier, G., Takacs, P., & Bourrat, P. (2021). Taming fitness: Organism-environment interdependencies preclude long-term fitness forecasting. BioEssays, 43, 2000157.

    Article  Google Scholar 

  • Gillespie, J. H. (1977). Natural selection for variances in offspring numbers: A new evolutionary principle. American Naturalist, 111, 1010–1014.

    Article  Google Scholar 

  • Glymour, B. (1999). Population level causation and a unified theory of natural selection. Biology and Philosophy, 14, 521–536.

    Article  Google Scholar 

  • Hájek, A. (2003). Conditional probability is the very guide of life. In Probability is the very guide of life: The philosophical uses of chance, edited by Kyburg Jr, E. Henry, and Mariam Thalos, 183–203. Open Court,

  • Hájek, A. (2007). The reference class problem is your problem too. Synthese, 156, 563–585.

    Article  Google Scholar 

  • Kitcher, P., Sterelny, K., & Waters, C. K. (1990). The illusory riches of Sober’s monism. The Journal of Philosophy, 87, 158.

    Article  Google Scholar 

  • Kokko, H., Griffith, S. C., & Pryke, S. R. (2014). The hawk–dove game in a sexually reproducing species explains a colourful polymorphism of an endangered bird. Proceedings of the Royal Society B: Biological Sciences, 281, 20141794.

    Article  Google Scholar 

  • Levins, R. (1968). Evolution in changing environments: Some theoretical explorations. Princeton University Press.

    Book  Google Scholar 

  • Maynard Smith, J. (1982). Evolution and the theory of games. Cambridge University Press.

    Book  Google Scholar 

  • Metz, J. A. J., Nisbet, R. M., & Geritz, S. A. H. (1992). How should we define ‘fitness’ for general ecological scenarios? Trends in Ecology and Evolution, 7, 198–202.

    Article  Google Scholar 

  • Mills, S. K., & Beatty, J. H. (1979). The propensity interpretation of fitness. Philosophy of Science, 46, 263–286.

    Article  Google Scholar 

  • Millstein, R. (2016). Probability in biology: The case of fitness. In A. Hájek & C. R. Hitchcock (Eds.), The Oxford handbook of probability and philosophy (pp. 601–622). Oxford University Press.

    Google Scholar 

  • Otsuka, J. (2016). A critical review of the Statisticalist debate. Biology and Philosophy, 31, 459–482.

    Article  Google Scholar 

  • Otsuka, J., Turner, T., Allen, C., & Lloyd, E. A. (2011). Why the causal view of fitness survives. Philosophy of Science, 78, 209–224.

    Article  Google Scholar 

  • Pence, C. H., & Ramsey, G. (2013). A new Foundation for the Propensity Interpretation of fitness. British Journal for the Philosophy of Science, 64, 851–881.

    Article  Google Scholar 

  • Pence, C. H., & Ramsey, G. (2015). Is organismic fitness at the basis of evolutionary theory? Philosophy of Science, 82, 1081–1091.

    Article  Google Scholar 

  • Ramsey, G. (2006). Block fitness. Studies in History and Philosophy of Science Part C: Studies in History and Philosophy of Biological and Biomedical Sciences, 37, 484–498.

    Article  Google Scholar 

  • Ridley, M. (1996). Evolution (2nd edition). Blackwell.

    Google Scholar 

  • Rosenberg, A. (1982). On the propensity definition of fitness. Philosophy of Science, 49, 268–273.

    Article  Google Scholar 

  • Rosenberg, A. (1983). Fitness. Journal of Philosophy, 80, 457–473.

    Article  Google Scholar 

  • Rosenberg, Alexander and Frédéric Bouchard. 2015. “Fitness,” Stanford Encyclopedia of Philosophy (Spring 2020 edition), ed. Edward N. Zalta, URL = <https://plato.stanford.edu/archives/spr2020/entries/fitness/>

  • Rosenberg, A., & Williams, M. B. (1986). Fitness as primitive and propensity. Philosophy of Science, 53, 412–418.

    Article  Google Scholar 

  • Scriven, M. (1959). Explanation and prediction in evolutionary theory: Satisfactory explanation of the past is possible even when prediction of the future is impossible. Science, 130, 477–482.

    Article  Google Scholar 

  • Sober, E. (1984). The nature of selection: Evolutionary theory in philosophical focus. University of Chicago Press.

    Google Scholar 

  • Sober, E. (2001). The two faces of fitness. In R. S. Singh, C. B. Krimbas, D. B. Paul, & J. H. Beatty (Eds.), Thinking about evolution: Historical, philosophical, and political perspectives (pp. 309–321). Cambridge University Press.

    Google Scholar 

  • Sober, E. (2011). Realism, conventionalism, and causal decomposition in units of selection: Reflections on Samir Okasha’s evolution and the levels of selection. Philosophy and Phenomenological Research, 82(1), 221–231.

    Article  Google Scholar 

  • Sober, E. (2013). Trait fitness is not a propensity, but fitness variation is. Studies in History and Philosophy of Biological and Biomedical Sciences, 44(3), 336–341.

    Article  Google Scholar 

  • Stevens, Lori. 2011. “Selection: Frequency-dependent,” in eLS (Wiley Online Library). https://doi.org/10.1002/9780470015902.a0001763.pub2

  • Tuljapurkar, Shripad. 2013. Population dynamics in variable environments (2nd edition). Springer: Berlin Heidelberg.

  • Wagner, G. P. (2010). The measurement theory of fitness. Evolution, 64–65, 1358–1376. https://doi.org/10.1111/j.1558-5646.2009.00909.x

  • Walsh, D. M. (2007). The pomp of superfluous causes: The interpretation of evolutionary theory. Philosophy of Science, 74(3), 281–303.

    Article  Google Scholar 

  • Walsh, D. M. (2010). Not a sure thing: Fitness, probability, and causation. Philosophy of Science, 77, 147–171.

    Article  Google Scholar 

  • Waters, K. C. (2011). Okasha’s unintended argument for toolbox theorizing. Philosophy and Phenomenological Research, 82, 232–240.

    Article  Google Scholar 

  • Walsh, D.M., Ariew, A., and Matthen, M. (2017). Four Pillars of Statisticalism, Philosophy, Theory, and Practice in Biology, vol. 9, no. 1, https://doi.org/10.3998/ptb.6959004.0009.001

  • Williams, M. B. (1970). Deducing the consequences of evolution: A mathematical model. Journal of Theoretical Biology, 29, 343–385.

    Article  Google Scholar 

Download references

Funding

Peter Takacs’s research was supported under Australian Research Council's Discovery Projects funding scheme (project number FL170100160).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Peter Takacs.

Ethics declarations

Conflict of interest

The Authors declare that there are no conflicts of interest.

Additional information

Publisher’s note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendix

Appendix

Statistical measures (e.g., relative fitness inequalities) in Table 1 are based on the following data

Individuals by Trait Type

Number of Offspring

Statistics

Local Partition E1

Global Partition (Union of E1 and ~ E1)

A

11

Arithmetic Mean for A in E1

10

  

A

9

  

A

10

  

A

9

Arithmetic Mean for B in E1

6

  

A

11

Variance for A in E1

0.8

  

B

7

Variance for B in E1

0.8

  

B

5

Geometric Mean for A in E1

9.96

  

B

6

Geometric Mean for B in E1

5.93

Arithmetic Mean for A

8

B

5

Arithmetic Mean for B

8

B

7

Variance for A

4.8

Local Partition ~ E1

Variance for B

4.8

A

7

Arithmetic Mean for A in ~E1

6

Geometric Mean for A

7.69

A

5

Geometric Mean for B

7.69

A

6

A

5

Arithmetic Mean for B in ~E1

10

  

A

7

Variance for A in ~E1

0.8

  

B

11

Variance for B in ~E1

0.8

  

B

9

Geometric Mean for A in ~E1

5.93

  

B

10

Geometric Mean for B in ~E1

9.96

  

B

9

    

B

11

    

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Takacs, P., Bourrat, P. Fitness: static or dynamic?. Euro Jnl Phil Sci 11, 112 (2021). https://doi.org/10.1007/s13194-021-00430-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s13194-021-00430-0

Keywords

Navigation