Skip to main content
Log in

Using conceptual spaces to exhibit conceptual continuity through scientific theory change

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

Abstract

There is a great deal of justified concern about continuity through scientific theory change. Our thesis is that, particularly in physics, such continuity can be appropriately captured at the level of conceptual frameworks (the level above the theories themselves) using conceptual space models. Indeed, we contend that the conceptual spaces of three of our most important physical theories—Classical Mechanics (CM), Special Relativity Theory (SRT), and Quantum Mechanics (QM)—have already been so modelled as phase-spaces. Working with their phase-space formulations, one can trace the conceptual changes and continuities in transitioning from CM to QM, and from CM to SRT. By offering a revised severity-ordering of changes that conceptual frameworks can undergo, we provide reasons to doubt the commonly held view that CM is conceptually closer to SRT than QM.

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

Similar content being viewed by others

Notes

  1. Connectionism (Bechtel and Abrahamsen 2002) is a special case of associationism that models associations using artificial neuron networks.

  2. Given a temporal dimension, the qualitative evolution of an object can be represented by a trajectory through the conceptual space.

  3. Quality dimensions can be continuous or discrete, and discrete spaces can also have metrics allowing measures of similarity. For instance, if three objects, A, B, and C, differ along two discrete dimensions such that A differs from B only along one dimension, but A differs from C along both dimensions, then A is more similar to B than A is to C, which is to say that the distance from A to B in the (Euclidean) plane of these dimensions (taking the value 1) is less than the distance from A to C (taking the value \( \sqrt{2} \)) in that plane.

  4. Holonomic system constraints can be expressed by a zero-valued function like \( f\left({x}_1,\;{x}_2,\;{x}_3, \dots, {x}_n,t\right)=0 \). It is characteristic of holonomic constraints, moreover, that they do not have velocities as arguments. For instance, the constraint \( {x}^2+{y}^2-{L}^2=0 \) on the position of a 2D pendulum of fixed length L is holonomic. A constraint that cannot be expressed in the above form is a non-holonomic constraint.

  5. Those inclined to grapple with scientific realism, of course, take the question ‘Which dimensions are basic, which derivative?’ to be one of principle. When Isaac Newton introduced the force domain into mechanics, for instance, many natural philosophers advocated—on grounds of ontological suspicion—a reformulation that made force a defined domain. Similarly, those following Niels Bohr in claiming that all measurements are ultimately spatio-temporal determinations, from which the values along other dimensions are inferred, view space-time as the basic domain. Whether conventional or principled, the division itself is (with the possible exception of such “Babylonian” physicists as Richard Feynman) present, and important to observe.

  6. We have previously argued that Kuhn’s account unduly assumes the primacy of the symbolic level of representation (Zenker and Gärdenfors 2015).

  7. This version of the criterion is a strengthening of the criterion proposed in Gärdenfors and Zenker (2011, 2013) where only changes of scales; e.g., the change from Celsius to Kelvin, were considered. This strengthening corresponds to a more severe change of a theory, so we now put it after changes in the importance of dimensions.

  8. Phase-spaces have been predominantly used to represent the degrees of freedom of complex systems and to model chaotic behavior. This use is related to their use in formulating physical theories in the tradition of Boltzmann (that we are interested in here) but often obscures that original purpose. For a historical introduction see Nolte (2010). For a brief introduction to classical and quantum phase-space see Tao (2007).

  9. The phase space formulation of quantum mechanics is particularly difficult to work in relative to the other formulations of that theory.

  10. There is a lesson for psychology, and other disciplines using conceptual spaces, here. In the literature on conceptual spaces one does not see the multiplying of dimensions with the multiplying of objects being located in the conceptual space, but if our thesis—that phase-spaces are just an example of a conceptual space—is correct, then this aspect of conceptual spaces should be recognized more widely.

  11. Besides our own, for instance, we know of no other work on non-Euclidean conceptual spaces, which is obviously needed if one admits that the conceptual space of relativity theory is non-Euclidean.

  12. One intriguing possibility that we cannot explore further here is to motivate a ranking—hopefully, but not necessarily, our ranking—of the severity of conceptual space changes by entailment relations, in a manner similar to how topological morphisms are ranked.

  13. Perhaps the metric test can be reformulated to account for such examples in the following manner: Dimensions are treated as separable if the data supports the city-block metric as the best measure of “distance” in the space formed by those dimensions and are otherwise treated as integral.

  14. The general schema is: \( T=\underset{\varepsilon \to 0}{ \lim }{T}^{*} \)

References

  • Batterman, R. W. (2001). The devil in the details: Asymptotic reasoning in explanation, reduction and emergence. Oxford: Oxford University Press.

    Book  Google Scholar 

  • Batterman, R. W. (2006). Reduction. In D. Borchert (Ed.), Encyclopedia of philosophy (2nd ed.). Detroit: Macmillan Reference.

    Google Scholar 

  • Bechtel, W., & Abrahamsen, A. (2002). Connectionism and the mind: Parallel processing, dynamics, and evolution in networks (2nd ed.). Oxford: Wiley.

    Google Scholar 

  • Bohr, N. (1963). Essays 1958–1962 on atomic physics and human knowledge. Woodbridge: Ox Bow Press.

    Google Scholar 

  • Douven, I., Decock, L., Dietz, R., & Égré, P. (2011). Vagueness: a conceptual spaces approach. Journal of Philosophical Logic, 42(1), 137–160.

    Article  Google Scholar 

  • Feynman, R. (1965). The character of physical law. Cambridge: MIT press.

  • Gärdenfors, P. (2000). Conceptual spaces: The geometry of thought. Cambridge: The MIT Press.

    Google Scholar 

  • Gärdenfors, P., & Zenker, F. (2011). Using conceptual spaces to model the dynamics of empirical theories. In E. J. Olsson & S. Enqvist (Eds.), Philosophy of science meets belief revision theory (pp. 137–153). Berlin: Springer.

    Google Scholar 

  • Gärdenfors, P., & Zenker, F. (2013). Theory change as dimensional change: conceptual spaces applied to the dynamics of empirical theories. Synthese, 190(6), 1039–1058.

    Article  Google Scholar 

  • Giovanni, A.-C., Laurent, F., Jerzy, K.-G., & Lee, S. (2011). The principle of relative locality. General Relativity and Gravitation, 43(10), 2547–2553. preprint, URL = http://arxiv.org/pdf/1101.0931v2.pdf . Accessed 1 Apr 2016.

  • Gleick, J. (1987). Chaos: Making a new science. New York: Viking.

    Google Scholar 

  • Hooker, C. A. (1992). Idealization, naturalism and rationality: some lessons from minimal rationality. Synthese, 99, 181–231.

    Google Scholar 

  • Kuhn, T. (1970). The structure of scientific revolutions (2nd ed.). Chicago: Chicago University Press.

    Google Scholar 

  • Maddox, W. T. (1992). Perceptual and decisional separability. In G. F. Ashby (Ed.), Multidimensional models of perception and cognition (pp. 147–180). Hillsdale, NJ: Erlbaum.

  • Maxwell, G. (1970). Structural realism and the meaning of theoretical terms. In S. Winokur & M. Radner (Eds.), Minnesota studies in the philosophy of science: Volume 4 (pp. 181–192). Minneapolis: University of Minnesota Press.

    Google Scholar 

  • Murdoch, D. (1987). Niels Bohr’s philosophy of physics. Cambridge: Cambridge University Press.

    Book  Google Scholar 

  • Nickles, T. (1973). Two concepts of intertheoretic reduction. The Journal of Philosophy, 70(7), 181–201.

    Article  Google Scholar 

  • Nolte, D. D. (2010). The tangled tale of phase space. Physics Today, 63(4), 33–38.

    Article  Google Scholar 

  • Petersen, G., & Zenker, F. (2014). From Euler to Navier–Stokes: a spatial analysis of conceptual changes in 19th-century fluid dynamics. International Studies in the Philosophy of Science, 28(3), 1–19.

    Article  Google Scholar 

  • Pylyshyn, Z. W. (1986). Computation and cognition: Towards a foundation for cognitive science. Massachusetts: MIT Press.

    Google Scholar 

  • Raubal, M. (2004). Formalizing conceptual spaces. In Varzi, A. & Vieu, L. (Eds.), Formal Ontology in Information Systems, Proceedings of the Third International Conference, (FOIS 2004) (Vol. 114, pp. 153–164). Amsterdam: IOS Press.

  • Rothman, M. A. (1972). Discovering the natural laws: The experimental basis of physics. Garden City: Anchor Books.

    Google Scholar 

  • Sankey, H. (2009). Scientific realism and the semantic incommensurability thesis. Studies in History and Philosophy of Science, 40(2), 196–202.

    Article  Google Scholar 

  • Schurz, G. (2009). When empirical success implies theoretical reference: a structural correspondence theorem. British Journal for the Philosophy of Science, 60, 101–133.

  • Suppes, P. (2002). Representation and invariance of scientific structures. Stanford: CSLI Publications.

    Google Scholar 

  • Tao, T. (2007). Phase space. http://www.math.ucla.edu/~tao/preprints/phase_space.pdf. Accessed 1 Apr 2016.

  • Thébault, K. P. Y. (2014). Quantization as a guide to Ontic structure. British Journal for the Philosophy of Science, 67(1), 89–114.

    Article  Google Scholar 

  • Zenker, F. (2014). From features via frames to spaces. Modeling scientific conceptual change without incommensurability or aprioricity. In T. Gamerschlag, D. Gerland, R. Osswald, & W. Petersen (Eds.), Concept types and frames. Applications in language and philosophy (pp. 69–89). Dordrecht: Springer.

    Chapter  Google Scholar 

  • Zenker, F., & Gärdenfors, P. (2014). Modeling diachronic changes in structuralism and in conceptual spaces. Erkenntnis, 79(8), 1547–1561.

    Article  Google Scholar 

  • Zenker, F., & Gärdenfors, P. (2015). Communication, rationality, and conceptual changes in scientific theories. In F. Zenker & P. Gärdenfors (Eds.), Applications of conceptual spaces: The case for geometric knowledge representation (pp. 259–277). Dordrecht: Springer.

    Google Scholar 

Download references

Acknowledgments

Parts of this paper were presented at GAP.9, held September 14–17 September 2015 in Osnabrück, Germany, and at EPSA15, held 23–26 September 2015 in Düsseldorf, Germany. We would like to thank audiences at these events, two anonymous reviewers for this journal, Keizo Matsubara and Lars-Göran Johansson for comments and criticisms that helped to improve earlier versions of this manuscript. The authors acknowledge funding from the Swedish Research Council (G.M., F.Z., P.G.), the European Union’s FP7 program as well as the Volkswagen Foundation (F.Z.)

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to George Masterton.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Masterton, G., Zenker, F. & Gärdenfors, P. Using conceptual spaces to exhibit conceptual continuity through scientific theory change. Euro Jnl Phil Sci 7, 127–150 (2017). https://doi.org/10.1007/s13194-016-0149-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s13194-016-0149-x

Keywords

Navigation