Skip to main content

Contrasting Cases: The Lotka-Volterra Model Times Three

  • Chapter
  • First Online:
The Philosophy of Historical Case Studies

Part of the book series: Boston Studies in the Philosophy and History of Science ((BSPS,volume 319))

Abstract

How do philosophers of science make use of historical case studies? Are their accounts of historical cases purpose-built and lacking in evidential strength as a result of putting forth and discussing philosophical positions? We will study these questions through the examination of three different philosophical case studies. All of them focus on modeling and on Vito Volterra, contrasting his work to that of other theoreticians. We argue that the worries concerning the evidential role of historical case studies in philosophy are partially unfounded, and the evidential and hermeneutical roles of case studies need not be played against each other. In philosophy of science, case studies are often tied to conceptual and theoretical analysis and development, rendering their evidential and theoretic/hermeneutic roles intertwined. Moreover, the problems of resituating or generalizing local knowledge are not specific to philosophy of science but commonplace in many scientific practices—which show similarities to the actual use of historical case studies by philosophers of science.

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 89.00
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 119.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD 119.99
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.

    Kinzel gives also Chang (2012) as an example of a proponent of a hermeneutic approach.

  2. 2.

    Some vestiges of the confrontation model seem to be at play here, even in the rejection of it.

  3. 3.

    In recent discussion on modeling, the idea of models as fictions has been entertained by several authors, e.g. Suárez (2008).

  4. 4.

    Godfrey-Smith (2006) likewise distinguishes between indirect representation and abstract direct representation—and invokes examples in trying to account for the difference between the two strategies of theorizing. Godfrey-Smith studies two influential books on evolutionary theory: Leo Buss’s The Evolution of Individuality (1987) and Maynard Smith and Szathmáry’s The Major Transitions in Evolution (1995). Buss examines the “actual relations between cellular reproduction and whole-organism reproduction in known organisms” (Godfrey-Smith 2006, p. 731), while Maynard Smith and Szathmáry put forth “idealized, schematic causal mechanisms.”

  5. 5.

    On models as independent or autonomous entities, see also Morgan and Morrison (1999) and Knuuttila (2005).

  6. 6.

    We are here largely following the discussion by Scholl and Räz (2013).

  7. 7.

    Our case studies are based on our earlier work (Knuuttila and Loettgers 2016).

  8. 8.

    Volterra had started his scientific career as a mathematician and had made important contributions to the theory of calculus. This work is summarized in Volterra’s book Theory of Functionals and of Integral and Integro-Differential Equations (Volterra 1930).

  9. 9.

    Volterra (1928) is a partial English translation of the Italian original (Volterra 1926b); in the following, references are made to the 1928 translation.

  10. 10.

    For this so-called Volterra principle, see Weisberg and Reisman (2008).

  11. 11.

    Lankester suggested that to protect edible prey-fish their enemies should be destroyed in the same proportion as the adult prey fish were “removed” (Lancaster 1884, p. 416).

  12. 12.

    On Volterra’s Darwinism, see Scudo (1992).

  13. 13.

    Volterra made use of the method of encounters also in his study of the demographic evolution of a single species: There he applied the method of encounters to mating.

  14. 14.

    Today Volterra is mostly known for the Lotka-Volterra equation. For a discussion on how Volterra’s various models anticipated several theoretical advances in theoretical ecology, see Scudo (1971).

  15. 15.

    A partial English translation of this paper can be found in Scudo and Ziegler (1978).

  16. 16.

    For example in Volterra (1934a, 1936) he discussed the connection between his theories and biological data.

  17. 17.

    The biologists with whom Volterra corresponded included Georgii F. Gause, R.N. Chapman, Jean Régnier, Raymond Pearl, Karl Pearson, D’Arcy W. Thompson, William R. Thompson, Alfred J. Lotka, and Vladimir A. Kostitzin. The correspondence of Vito Volterra on mathematical biology also provides interesting material as regards modeling methods. Among other things, it provides material on the choices between deterministic and probabilistic approaches; between continuous and discrete models; between closed-form solutions and numerical solutions, and between qualitative and quantitative models.

  18. 18.

    Lotka dealt with the rhythmic effects of chemical reactions already in his earlier writings, see e.g. Lotka (1910).

  19. 19.

    Our reading of Scholl and Räz (2013) differs somewhat from their later reading of their own article (this volume). In their original article, they do not clearly offer causal inference as a contrast to modeling (that would provide an alternative for Weisberg’s contrast between modeling and abstract direct representation). According to them “much of our discussion will focus on models of causal structures” (Scholl and Räz 2013, p. 117, emphasis added). Their earlier focus was on causal inference in general and modeling a strategy to deal with insufficient epistemic success.

  20. 20.

    Scholl and Räz adopt the distinction between “how possibly” and “how actually” from the discussion on mechanistic explanation (Machamer et al. 2000), a discussion that has been up until recently relatively disinterested in modeling and that considered models as explanation sketches only (see Knuuttila and Loettgers 2013).

  21. 21.

    The “how-possible” may be a bit misleading expression in this context, since what Volterra accomplished was an alternative explanation for the prevailing explanations that attributed the fluctuations to external causes.

  22. 22.

    A model template is an abstract conceptual idea concerning usually a certain kind of interaction and associated with particular mathematical forms and computational methods, see for further discussion Knuuttila and Loettgers (2014).

  23. 23.

    Even though the differences between the three case studies with respect to Volterra’s work were more substantial, such underdetermination of theories by data would be a common feature of other scientific practices, too.

References

  • Ankeny, R.A. 2012. Detecting themes and variations: The use of cases in developmental biology. Philosophy of Science 79: 644–654.

    Article  Google Scholar 

  • Bailer-Jones, D. 2009. Scientific models in philosophy of science. Pittsburgh: University Press.

    Google Scholar 

  • Burian, R. 2001. The dilemma of case studies resolved: The virtues of using case studies in the history and philosophy of science. Perspectives on Science 9: 383–404.

    Article  Google Scholar 

  • Buss, L. 1987. The evolution of individuality. Princeton University Press.

    Google Scholar 

  • Chang, H. 2012. Is water \(H_{2}O\) ? evidence, pluralism and realism. Boston studies in the philosophy and history of science. Dordrecht: Springer.

    Google Scholar 

  • Contessa, G. 2007. Scientific representation, interpretation, and surrogative reasoning. Philosophy of Science 74: 48–68.

    Article  Google Scholar 

  • da Costa, N.C.A., and S. French. 2000. Models, theories and structures: Thirty years on. Philosophy of Science 67: 116–127.

    Article  Google Scholar 

  • Darwin, C. 1882. The origin of the species by means of natural selection. Murray.

    Google Scholar 

  • French, S., and J. Ladyman. 1999. Reinflating the semantic approach. International Studies in the Philosophy of Science 13: 103–21.

    Article  Google Scholar 

  • Frigg, R. 2010. Models and fiction. Synthese 172: 251–268.

    Article  Google Scholar 

  • Gause, G. 1935. Studies in the ecology of the orthoptera. Ecology 11: 307–325.

    Article  Google Scholar 

  • Giere, R.N. 2004. How models are used to represent reality. Philosophy of Science (Symposia) 71: 742–752.

    Article  Google Scholar 

  • Godfrey-Smith, P. 2006. The strategy of model-based science. Biology and Philosophy 21: 725–740.

    Article  Google Scholar 

  • Haydon, D., and A. Lloyd. 1999. On the origins of the Lotka-Volterra equations. Bulletin of the Ecological Society of America 80: 205–206.

    Article  Google Scholar 

  • Helm, G. 1898. Die Energetik. Verlag von Veit.

    Google Scholar 

  • Howard, L.O. 1897. A study in insect parasitism: A consideration of the parasites of the white-marked tussok moth with an account of their habits and interrelations, and with descriptions of new species. U.S. Department of Agriculture Technical Bulletin 5: 1–57.

    Google Scholar 

  • Humphreys, P. 2002. Computational models. Proceedings of the Philosophy of Science Association 3: S1–S11.

    Article  Google Scholar 

  • Humphreys, P. 2004. Extending ourselves: Computational science, empiricism, and scientific method. Oxford University Press.

    Google Scholar 

  • Israel, G. 1991. Volterra’s analytical mechanics of biological associations (second part). U.S. Department of Agriculture Technical Bulletin 5: 1–57.

    Google Scholar 

  • Israel, G. 1993. The emergence of biomathematics and the case of population dynamics: A revival of mechanical reductionism and darwinism. Science in Context 6: 469–509.

    Article  Google Scholar 

  • Israel, G., and Gasca, A.M. 2002. The biology of numbers: The correspondence of Vito Volterra on mathematical biology. Birkhäuser.

    Google Scholar 

  • Kinzel, K. 2015. Narrative and evidence: How can case-studies from the history of science support claims in the philosophy of science? Studies in History and Philosophy of Science Part A 49: 48–57.

    Article  Google Scholar 

  • Knuuttila, T. 2005. Models, representation, and mediator. Philosophy of Science 72: 1260–1271.

    Article  Google Scholar 

  • Knuuttila, T. 2006. From representation to production: Parsers and parsing in language technology. Sociology of Science 25: 41–55.

    Google Scholar 

  • Knuuttila, T. 2013. Science in a new mode: Good old (theoretical) science vs. brave new (commodified) knowledge production. Science and Education 22: 2443–2461.

    Article  Google Scholar 

  • Knuuttila, T., and Loettgers, A. 2013. The productive tension: Mechanisms vs. templates in modeling the phenomena. In Representations, models, and simulations, ed. P. Humphreys, and C. Imbert, 3–24. Routledge.

    Google Scholar 

  • Knuuttila, T., and A. Loettgers. 2014. Varieties of noise: Analogical reasoning in synthetic biology. Studies in History and Philosophy of Science Part A 48: 76–88.

    Article  Google Scholar 

  • Knuuttila, T., and Loettgers, A. 2016. Modelling as indirect representation? the Lotka-Volterra model revisited. The British Journal for the Philosophy of Science.

    Google Scholar 

  • Lancaster, E.R. 1884. The scientific results of the exhibition. Fisheries Exhibition Literature 4: 405–442.

    Google Scholar 

  • Loettgers, A. 2007. Model organisms, mathematical models, and synthetic models in exploring gene regulatory mechanisms. Biological Theory 2: 134–142.

    Article  Google Scholar 

  • Lotka, A. 1910. Contribution to the theory of periodic reactions. The Journal of Physical Chemistry 14: 271–274.

    Article  Google Scholar 

  • Lotka, A. 1920a. Analytical note on certain rhythmic relations in organic systems. Proceedings of the National Academy of Art and Science 42: 410–415.

    Article  Google Scholar 

  • Lotka, A. 1920b. Undamped oscillations derived from the law of mass action. Journal of the American Chemical Society 42: 1595–1598.

    Article  Google Scholar 

  • Lotka, A. (1925). Elements of physical biolgy. Williams and Wilkins.

    Google Scholar 

  • Machamer, P.K., L. Darden, and C. Craver. 2000. Thinking about mechanisms. Philosophie of. Science 67: 1–25.

    Google Scholar 

  • Maeki, U. 2009. Missing the world: Models as isolations and credible surrogate systems. Erkenntnis 70: 29–43.

    Article  Google Scholar 

  • May, R. 1974. Biological populations with nonoverlapping genertations: Stable cycles, and chaos. Science 186: 645–647.

    Article  Google Scholar 

  • Maynard-Smith, J., and E. Smathmáry. 1995. The major transitions in evolution. Oxford: Oxford University Press.

    Google Scholar 

  • Morgan, M. 2012. Case studies: One observation or many? justification or discovery? Philosophy of Science 79: 667–677.

    Article  Google Scholar 

  • Morgan, M. 2014. Resulting knowledge: Generic strategies and case studies. Philosophy of Science 81: 1012–1024.

    Article  Google Scholar 

  • Morgan, M., and Morrison, M. 1999. Models as mediating instruments. In Models as mediators. Perspectives on natural and social science, ed. M. Morgan, and M. Morrison, 97–146. Cambridge University Press.

    Google Scholar 

  • Ostwald, W. 1893. Lehrbuch der allgemeinen Chemie: Chemische Energie. Wilhelm Engelman.

    Google Scholar 

  • Pitt, J.C. 2001. The dilemma of case studies: Toward a heraclitian philosophy of science. Perspectives on Science 9: 373–382.

    Article  Google Scholar 

  • Schickore, J. 2011. More thoughts on HPS: Another 20 years later. Perspectives on Science 19: 453–481.

    Article  Google Scholar 

  • Scholl, R., and T. Räz. 2013. Modeling causal structures: Volterra’s struggle and Darwin’s success. European Journal for Philosophy of Science 3: 115–132.

    Article  Google Scholar 

  • Scudo, F. 1971. Vito Volterra and theoretical ecology. Theoretical Population Biology 2: 1–23.

    Article  Google Scholar 

  • Scudo, F. 1992. Vito Volterra, ecology and the quantification of Darwinism. Accademia Nazionale dei Lincei: Atti del Convegno internazionale in memorie di Vito Volterra. Rome.

    Google Scholar 

  • Scudo, F.M., and J.R. Ziegler. 1978. The golden age of theoretical ecology: 1923–1940. Berlin: Springer.

    Google Scholar 

  • Servos, J.W. 1990. Physical chemistry from ostwald to pauling: the making of a science in America. Princeton University Press.

    Google Scholar 

  • Shrader-Frechette, K.S., and E.D. McCoy. 1994. How the tail wags the dog: How value judgments determine ecological science. Environmental Values 3: 107–120.

    Article  Google Scholar 

  • Simon, H. 1981. The sciences of the artifical. MIT Press.

    Google Scholar 

  • Suárez, M. (2008). Scientific fictions as rules of inference. In Fictions in science: philosophical essays on modeling and idealisation, ed. M. Suárez, 258–270. Routledge.

    Google Scholar 

  • Thomson, W.R. 1922. ThĂ©orie de l’action des parasites entomophages. les formules mathĂ©matiques du parasitisme cyclique. Comptes Rendus Academie des Sciences Paris 174: 1202–1204.

    Google Scholar 

  • Volterra, V. 1901. On the attempts to apply mathematics to the biological and social sciences. In The Volterra chronicles: Life and times of an extraordinary mathematician 1860–1940, ed. J.D. Goodstein, 247–260. AMS.

    Google Scholar 

  • Volterra, V. 1926a. Fluctuations in the abundance of a species considered mathematically. Nature, CXVIII:558–560.

    Google Scholar 

  • Volterra, V. 1926b. Variazioni e fluttuazioni del numero d’indivui in specie animali conviventi. Memorie della R. Accademia Lincei 2: 31–113.

    Google Scholar 

  • Volterra, V. 1927. Letter to nature. Nature 119: 12.

    Google Scholar 

  • Volterra, V. 1928. Variations and fluctuations of the number of individuals in animal species living together. Journal du Conseil International Pour l’Exploration de la Mer 3: 3–51.

    Article  Google Scholar 

  • Volterra, V. 1930. Theory of functionals and of integral and integro-differential equations. Blackie and Son.

    Google Scholar 

  • Volterra, V. (1931). Leçons sur la ThĂ©orie MathĂ©matique de la Lutte Pour la Vie. Gauthier-Villars.

    Google Scholar 

  • Volterra, V. 1936. La thĂ©orie mathĂ©matique de la lutte pour la vie et l’expĂ©rience (a propos de deux ouvrages de C.F. Gause). Scientia LX: 169–174.

    Google Scholar 

  • Volterra, V. 1937a. Applications des mathĂ©matiques Ă  la biologie. L’Enseignement MathĂ©matique XXXVI: 297–330.

    Google Scholar 

  • Volterra, V. 1937b. Principes de biologie mathĂ©matique. Acta Biotheoretica III: 1–36.

    Article  Google Scholar 

  • Volterra, V., and D’Ancona, U. 1935. Les associations biologiques au point de vue mathĂ©matique. Hermann.

    Google Scholar 

  • von Bertalanffy, L. 1968. General system theory: Foundations, development, applications. George Braziller.

    Google Scholar 

  • Weisberg, M. 2007. Who is a modeler? British Journal for the Philosophy of Science 58(2): 207–233.

    Article  Google Scholar 

  • Weisberg, M. 2013. Simulation and similarity: Using models to understand the World. Oxford University Press.

    Google Scholar 

  • Weisberg, M., and K. Reisman. 2008. The robust Volterra principle. Philosophy of Science 75: 106–131.

    Article  Google Scholar 

  • Whyte, F. 1943. The street corner society. The social structure of an Italian slum: The University of Chicago Press.

    Google Scholar 

  • Wiener, N. 1948. Cybernetics or control and communication in the animal and the machine. MIT Press.

    Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Tarja Knuuttila .

Editor information

Editors and Affiliations

Rights and permissions

Reprints and permissions

Copyright information

© 2016 Springer International Publishing Switzerland

About this chapter

Cite this chapter

Knuuttila, T., Loettgers, A. (2016). Contrasting Cases: The Lotka-Volterra Model Times Three. In: Sauer, T., Scholl, R. (eds) The Philosophy of Historical Case Studies. Boston Studies in the Philosophy and History of Science, vol 319. Springer, Cham. https://doi.org/10.1007/978-3-319-30229-4_8

Download citation

  • DOI: https://doi.org/10.1007/978-3-319-30229-4_8

  • Published:

  • Publisher Name: Springer, Cham

  • Print ISBN: 978-3-319-30227-0

  • Online ISBN: 978-3-319-30229-4

  • eBook Packages: HistoryHistory (R0)

Publish with us

Policies and ethics