Skip to main content
Log in

Defending the Semantic View: what it takes

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

Abstract

In this paper, a modest version of the Semantic View is motivated as both tenable and potentially fruitful for philosophy of science. An analysis is proposed in which the Semantic View is characterized by three main claims. For each of these claims, a distinction is made between stronger and more modest interpretations. It is argued that the criticisms recently leveled against the Semantic View hold only under the stronger interpretations of these claims. However, if one only commits to the modest interpretation for all the claims, then the view obtained, the Modest Semantic View, is tenable and fruitful for the philosophy of science.

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. Major developments of the Semantic View are in Suppes (1960, 1962, 1967), Suppe (1974, 1977a, 1979, 1989), Giere (1979, 1988, 1999), and van Fraassen (1972, 1980, 1989, 1991, 2008). In their subsequent works, all these authors typically remained in favor of the Semantic View broadly speaking, even if they each developed particular versions of it, and never formed a unified view. A much more unified view has been initiated by Sneed (1971), and developed in Europe and especially in Germany thanks to the work by Wolfgang Stegmüller (1976, 1979a, b, 1986), C. U. Moulines (1975a, b, 1982, 1991, 2002) and Wolfgang Balzer (1978, 1982, 1985), among others. The view associated with these authors has been called the “non-statement view” but it is now known as the “structuralist program”. One classic work associated with the structuralist program is Balzer et al. (1987). Though this article is devoted to the Anglo-Saxon tradition of the Semantic View, we shall point out some obvious affinities and differences between our view and the structuralist program on various occasions in what follows.

  2. See for example Cartwright (1983, 1989, 1999), Cartwright et al. (1995), Ereshefsky (1991), Downes (1992), Morrison (1999), Suárez (2003), Thomson-Jones (2006), Frigg (2006), Morrison (2007), Suárez and Cartwright (2008), and Krause and Bueno (2008). Even if not explicitly directed against the Semantic View, Redhead (1980) is an important background contribution to the debate.

  3. Important recent work in which the program of the Semantic View is defended and developed has been undertaken by Bueno et al. in various publications, including: Bueno (1997), French and Ladyman (1999), Bueno et al. (2002) and Da Costa and French (2003).

  4. What follows is a conceptual reconstruction of the Semantic View. No pretension is made to give a historically adequate account of how the Semantic View developed. Note also that the differences between the set-theoretical approach of Suppes, the state–space approach of van Fraassen and the relational system approach of Suppe are irrelevant here. We agree with Suppe (1989, p. 4) and da Costa and French (2003, p. 23) that van Fraassen’s and Suppe’s approaches can be conceived within Suppes’ set-theoretical framework.

  5. The “toolbox instrumentalists” conceive of theories as tools for the construction of models through “piecemeal borrowing”. See Cartwright et al. (1995) and Suárez and Cartwright (2008).

  6. To conceive of scientific theories as set of laws or of axioms was part of what has been labeled the “received view” by Putnam in his (1962). It is not clear that the label refers to a single unified view rather than to a variety of positions taken at different times by various authors (for a presentation of the historical development of the received view, see for example Friedman 1999 or Parrini et al. 2003). That said, the view is generally taken to be as presented in Carnap (1966) and Hempel (1963).

  7. Suppe (1977a, pp. 62–66) provides an analysis of what counts as a fruitful axiomatization.

  8. It is well known that there are various kinds of ‘scientific models’. In this paper, no stance is taken as to any particular mode of representation (structural, analogous, iconic etc.) by which scientific models represent the world. For a synthesis of the notions of model in science, see Frigg and Hartman (2006). The relationships between the notions of scientific models (as representing phenomena in the world) and logical models will be discussed in Section 3.

  9. Or, as it is often easier, from the scientific literature in the field. Note that, in doing so, one avoids the problem of having to define what is the theory before being able to consider the models used by the scientists to represent the world. The class of models that the semanticist sets himself to study is simply the class of models that are used to represent a certain class of phenomena.

  10. This distinction between the modest and the stronger interpretation of (Models) is related, but not identical to the distinctions between intrinsic and extrinsic characterization of theories by Suppes (1967, pp. 60–62, and 2002), and between representational and constitutional role of models by da Costa and French (2003, p. 34).

  11. Whether or not this claim holds for structuralist program as well is an interesting question. Clearly, in spirit, and just like the proponents of the Semantic View, the structuralists appear not to adopt the stronger interpretation of (Models), since their “theory-elements”—roughly, elements of what is commonly called a theory—contain, in addition to sets of models, other components, including, among others, the intended domain of application of the theory and the necessary approximations associated with such application. See Schmidt (2008) for a short exposition of the notion of theory-element, and for more details, see Balzer et al. (1987). The question of whether or not the structuralists systematically make the distinction explicit and reject the first interpretation of (Models) requires a detailed investigation that falls beyond the scope of this paper.

  12. Our modest interpretation of (Models) is similar to Downes’ “Deflationary Semantic View” as described in his (1992), which is constituted by the claim that “model construction is an important part of scientific theorizing”, while rejecting the claim that “all scientific theories are simply families of models” (p. 151). That said, the Modest Semantic View defended here differs from Downes’ version of the Semantic View, because it contains, in addition to the modest interpretation of (Models), the modest interpretations of (Scientific=Logical) and of (→ Adequacy), as described in the following two sections.

  13. The common claim that the received view demands that scientific theories be axiomatized in first order logic has been recently challenged in the literature (see Lutz 2010).

  14. On the question of whether the received view and the Semantic View are equivalent due to the Completeness Theorem, see the reviews on van Fraassen (1980) by Friedman (1982) and by Worrall (1984) along with van Fraassen’s answers in van Fraassen (1985, p. 301 sq.) and van Fraassen (1989, p. 211 sq.), which we believe are satisfactory. The discussion is summarized in da Costa and French (2003, pp. 30–31). On the question of wether the Semantic View—using first order model theory—is no better off than the received view—using first order logic plus identity, see French and Ladyman (1999) and references therein.

  15. Rigorously speaking, the interpretation is a domain along with a function assigning extensions to non-logical terms.

  16. It should be clear that “truth” is used here in the weak sense of “satisfaction”, or “truth within an interpretation”.

  17. It is important to note that scientific models do not have to be mathematical structures. It suffices that scientific models possesses a structure (roughly, a domain and relations between the members of this domain). Any structure, whether or not this structure is captured by nice mathematical equations, can be considered from an abstract point of view. This leaves room for an account of structures within the sciences in which nice mathematical equations are not often possible to obtain.

  18. A distinction between at least between three kinds of approximation arises from the literature: (1) Aristotelian abstractions consist in choosing the relevant parameters and variables of the system at hand—cf. Cartwright (1989, chap. 5); (2) Galilean idealizations consist in neglecting some parameters and variables that are clearly relevant the situation studied—cf. McMullin (1985); and (3) mathematical impoverishments consist in radically modifying the original equations such that solutions be tractable—cf. Redhead (1980).

  19. This is true at least within the Anglo-Saxon tradition of the Semantic View. The structuralist program also developed, using the notion of a theory–net, a more subtle view of the complexities of the relationships between theories, models, and the data models. In a theory–net, various more “specialized” theories branch from a core theory-element. For more details on such an account, see Balzer et al. (1987, chap. IV).

  20. It should be noted that Lloyd’s analysis does not include the assumption that population genetics is all there is to evolutionary theory. Rather, she aims at giving an account of the structure of evolutionary theory in a broader sense, including issues about species formation and extinction, group selection, and the tempo and mode of selection (1994, chap. 1, note 4 and chap. 3) [1988].

  21. Morrisson (1999) is often cited for posing the criticism of the Semantic View along these lines. That said, Cartwright (1983, 1989) has been developing similar views over the past twenty years. More recently, Thomson-Jones (2006) gives an analysis of the relation between logical models and scientific models, the former being characterized by their truth making function, and the latter by their representational function. Brading and Landry (2006) and Frigg (2006) gives also a systematic analysis of the problem. It should be noted that van Fraassen presents the same problem in his (2006), and offers an extended discussion of it in his (2008, chap. 11), where he defends his “empiricist structuralism”.

  22. This applies to the structuralist treatment of the “empirical claims” associated with a theory as well. See Balzer at al. (1987, chap. II) for more details.

  23. Note also that model theory is useless when it comes to addressing the question of whether theories represent the features of the world that are not captured by the data models—a question on which the realist and the empiricist depart from one another. More on this point later.

  24. Technically, they distinguish between four types of approximation (Balzer et al. 1987, chap. VII.1). The one we are interested in here is the first one: “model construction approximation” (Balzer et al. 1987, p. 325).

  25. It should be noted that Balzer et al. (1987, VII.2.2) manage to formally characterize a set of necessary (but not sufficient) conditions for an approximation to be “admissible”. For example, they formalize the rather intuitive notion that a given admissible approximation should remain so if one changes only “purely theoretical” aspects of the models associated with it. They also clearly recognize, however, that some strong pragmatic components of model construction are not formally explicable.

  26. The exegesis of van Fraassen’s thought, and in particular, the study of the difference between constructive empiricism and empiricist structuralism falls outside of the scope of this paper.

  27. See for example Da Costa and French (2003) for a systematic treatment of their view.

  28. Note that this is not contradictory to our claim that the tools offered by the Semantic View are useless when it comes to solve the fundamental problem of representation. French and Ladyman are well aware of this and commit only to the claim that the Semantic Approach allows us to give an account of the relationships between the various models in the hierarchy of models as described in Section 2 (see note 12, p. 119, in their 1999). This remains true within Ladyman’s most recent ontic structural realism as exposed in Ladyman and Ross (2007). There, Ladyman and Ross explain that formal methods are good for representing and investigating “the relationships between theoretical models and models of the phenomena” (p. 117), but not for explaining how what they call “structures”, which “are to be understood as mathematical models”, can represent “real patterns” in the world (pp. 118–120). This holds true also for the accounts of evidence by Suppes, of confirmation by Lloyd, and of empirical grounding by van Fraassen in his (2008).

References

  • Balzer, W. (1978). Empirische geometrie und Raum–Zeit–Theorie in mengentheoretischer Darstellung. Kronberg: Scriptor.

    Google Scholar 

  • Balzer, W. (1982). Empirische theorien: Modelle, strukturen, beispiele. Braunschweig: Vieweg.

    Google Scholar 

  • Balzer, W. (1985). Theorie und messung. Berlin: Springer.

    Book  Google Scholar 

  • Balzer, W., Ulises Moulines, C., & Sneed, J. D. (1987). An architectonic for science: The structuralist program. Dordrecht: Reidel.

    Book  Google Scholar 

  • Balzer, W., Ulises Moulines, C., & Sneed, J. (Eds.) (2000). Structuralist knowledge representation: Paradigmatic examples. Amsterdam: Rodopi.

    Google Scholar 

  • Bickle, J. (2002). Concepts structured through reduction: A structuralist resource illuminates the consolidation-long-term potentiation (LTP). Synthese, 130(1), 123–133.

    Article  Google Scholar 

  • Bourbaki, N. (pseud.) (1954). Eléments de mathematiques: Theorie des ensembles. Paris: Herman.

  • Brading, K., & Landry, E. M. (2006). Scientific structuralism: Presentation and representation. Philosophy of Science, 73, 571–581.

    Article  Google Scholar 

  • Bueno, O. (1997). Empirical adequacy: A partial structures account. Studies in History and Philosophy of Science, 28, 585–610.

    Article  Google Scholar 

  • Bueno, O., French, S., & Ladyman, J. (2002). On representing the relationships between the mathematical and the empirical. Philosophy of Science, 69, 497–518.

    Article  Google Scholar 

  • Caamaño, M. (2009). A structural analysis of the Phlogiston case. Erkenntnis, 70, 331–364.

    Article  Google Scholar 

  • Carnap, R. (1966). Philosophical foundations of physics. New York: Macmillan.

    Google Scholar 

  • Cartwright, N. D. (1983). How the laws of physics lie. Oxford: Oxford University Press.

    Book  Google Scholar 

  • Cartwright, N. D. (1989). Nature’s capacities and their measurement. Oxford: Oxford University Press.

    Google Scholar 

  • Cartwright, N. D. (1999). Models and the limits of theory: Quantum Hamiltonians and the BCS model of superconductivity. In M. S. Morgan, & M. Morrison (Eds.), Models as mediators (pp. 241–281). Cambridge: Cambridge University Press.

    Chapter  Google Scholar 

  • Cartwright, N. D., Shomar, T., & Suárez, M. (1995). The tool box of science: Tools for the building of models with a superconductivity example. In W. E. Herfel, W. Krajewski, I. Niiniluoto, & R. Wojcicki (Eds.), Theories and models in science: Poznan studies in the philosophy of the sciences and the humanities (vol. 44, pp. 137–149). Amsterdam: Rodopi.

    Google Scholar 

  • da Costa, N. C. A., & French, S. (2003). Science and partial truth: A unitary approach to models and scientific reasoning. Oxford: Oxford University Press.

    Google Scholar 

  • Downes, S. M. (1992). The importance of models in theorizing: A deflationary Semantic View. PSA 1992, 1, 142–153.

    Google Scholar 

  • Ereshefsky, M. (1991). The semantic approach to evolutionary theory. Biology and Philosophy, 5, 7–28.

    Google Scholar 

  • French, S., & Ladyman, J. (1999). Reinflating the semantic approach. International Studies in the Philosophy of Science, 13(2): 103–121.

    Article  Google Scholar 

  • Friedman, M. (1982). Review of the scientific image. Journal of Philosophy, 79(5), 274–283.

    Article  Google Scholar 

  • Friedman, M. (1999). Reconsidering logical positivism. Cambridge: Cambridge University Press.

    Book  Google Scholar 

  • Frigg, R. (2006). Scientific representation and the Semantic View of theories. Theoria, 55, 49–65.

    Google Scholar 

  • Frigg, R., & Hartman, S. (2006). Models in science. In E. N. Zalta (Ed.), Stanford encyclopedia of philosophy. http://plato.stanford.edu/archives/spr2006/entries/models-science/.

  • Gähde, U. (2002). Holism, underdetermination, and the dynamics of empirical theories. Synthese, 130(1), 69–90.

    Article  Google Scholar 

  • Gamut, L. T. F. (1990). Logic, language and meaning: Introduction to logic. Chicago: University of Chicago Press.

    Google Scholar 

  • Giere, R. (1979). Understanding scientific reasoning. New York: Holt, Rinehart and Winston.

    Google Scholar 

  • Giere, R. (1988). Explaining science. Chicago: University of Chicago Press.

    Google Scholar 

  • Giere, R. (1999). Science without laws. Chicago: University of Chicago Press.

    Google Scholar 

  • Hempel, C. (1963). Implications of Carnap’s work for the philosophy of science. In P. A. Schilpp (Ed.), The philosophy of Rudolf Carnap (pp. 685–709). LaSalle: Open Court.

    Google Scholar 

  • Krause, D., & Bueno, O. (2008). Scientific theories, models, and the semantic approach. Preprint: http://philsci-archive.pitt.edu/archive/00003958/. Forthcoming in Principia.

  • Ladyman, J., & Ross, D. (with Spurrett, D. and Collier, J.) (2007). Every thing must go: Metaphysics naturalised. Oxford: Oxford University Press.

  • Löwenheim, L. (1915). Über Möglichkeiten im Relativkalkül”. Mathematische Annalen, 76, 447–470. Eng. transl.: On possibilities in the calculus of relatives. In J. van Heijenoort (Ed.), From Frege to Gödel: A source book in mathematical logic, 1879–1931 (pp. 228–251). Cambridge: Harvard University Press.

  • Lloyd, E. A. (1994) [1988]. The structure and confirmation of evolutionary theory. Princeton: Princeton University Press.

    Google Scholar 

  • Lutz, S. (2010). On a straw man in the philosophy of science—a defense of the received view. Paper presented at the British Society for the Philosophy of Science Annual Conference 2010 (Dublin; 8–9 July 2010). Available in a draft version at http://philsci-archive.pitt.edu/5497/.

  • Maudlin, T. (2007). The metaphysics within physics. Oxford: Oxford University Press.

    Book  Google Scholar 

  • McMullin, E. (1985). Galilean idealization. Studies in History and Philosophy of Science, 16, 247–273.

    Article  Google Scholar 

  • Morgan, M. S., & Morrison, M. (1999). Models as mediators. Cambridge: Cambridge University Press.

    Book  Google Scholar 

  • Morrison, M. (1999). Models as autonomous agents. In M. S. Morgan, & M. Morrison (Eds.), Models as mediators (pp. 38–65). Cambridge: Cambridge University Press.

    Chapter  Google Scholar 

  • Morrison, M. (2007). Where have all the theories gone? Philosophy of Science, 74, 195–228.

    Article  Google Scholar 

  • Moulines, C. U. (1975a). Zur logischen Rekonstruktion der Thermodynamik. Eine wissenschafts-theoretische Analyse. Tesis doctoral, Universidad de Munich, Munich.

  • Moulines, C. U. (1975b). A logical reconstruction of simple equilibrium thermodynamics. Erkenntnis, 9, 101–130.

    Article  Google Scholar 

  • Moulines, C. U. (1982). Exploraciones metacient’ficas. Madrid: Alianza.

    Google Scholar 

  • Moulines, C. U. (1991). Pluralidad y recursión. Madrid: Alianza.

    Google Scholar 

  • Moulines, C. U. (2002). Structuralism as a program for modelling theoretical science. Synthese (Structuralism), 130(1), 1–11.

    Article  Google Scholar 

  • Moulines, C. U., & Sneed, J. (1979). Suppes’ philosophy of physics. In R. J. Bogdan (Ed.), Patrick Suppes (pp. 59–91). Dortrecht: Reidel.

    Chapter  Google Scholar 

  • Niegergall, K.-G. (2002). Structuralism, model theory and reduction. Synthese, 130(1), 135–162.

    Article  Google Scholar 

  • Parrini, P., Salmon, M., & Salmon, W. C. (Eds.) (2003). Logical empiricism: Historical and contemporary perspectives. Pittsburgh: University of Pittsburgh Press.

    Google Scholar 

  • Poincaré, H. (1952) [1905]. Science and hypothesis. New York: Dover. Translated by G. B. Halsted (1905). Reprinted (1958).

    Google Scholar 

  • Poincaré, H. (1958) [1906] [1914]. The value of science. New York: Dover. Translated by G. B. Halsted (1914). Reprinted (1958).

    Google Scholar 

  • Putnam, H. (1962) [1979]. What theories are not. In E. Nagel, P. Suppes, & A. Tarski (Eds.), Logic, methodology and philosophy of science. Proceedings of the 1960 international congress. Stanford: Stanford University Press. Reprinted in H. Putnam (1979). Mathematics, matter and method: Philosophical papers (vol. 1, pp. 215–227). Cambridge: Cambridge University Press.

    Google Scholar 

  • Redhead, M. (1980). Models in physics. British Journal for Philosophy of Science, 31, 145–163.

    Article  Google Scholar 

  • Schmidt, H.-J. (2008). Structuralism in physics. In E. N. Zalta (Ed.), The Stanford encyclopedia of philosophy (Fall 2008 ed.). http://plato.stanford.edu/archives/fall2008/entries/physics-structuralism/.

  • Sneed, J. D. (1971). The logical structure of mathematical physics. Dordrecht: Reidel.

    Book  Google Scholar 

  • Stegmüller, W. (1976). The structure and dynamics of theories. New York: Springer.

    Google Scholar 

  • Stegmüller, W. (1979a). The structuralist view: Survey, recent developments and answers to some criticisms. In I. Niiniluoto, & R. Tuomela (Eds.), The logic and epistemology of scientific change. Amsterdam: North Holland.

    Google Scholar 

  • Stegmüller, W. (1979b). The structuralist view of theories. New York: Springer.

    Book  Google Scholar 

  • Stegmüller, W. (1986). Die Entwicklung des neuen Strukturalismus seit 1973. Berlin: Springer.

    Google Scholar 

  • Suárez, M. (1999). The role of models in the application of scientific theories: Epistemological implications. In M. S. Morgan, & M. Morrison (Eds.), Models as mediators (pp. 168–195). Cambridge: Cambridge University Press.

    Chapter  Google Scholar 

  • Suárez, M. (2003). Scientific representation: Against similarity and isomorphism. International Studies in Philosophy of Science, 17, 225–244.

    Article  Google Scholar 

  • Suárez, M., & Cartwright, N. (2008). Theories: Tools versus models, Studies in History and Philosophy of Modern Physics, 39(1), 62–81.

    Article  Google Scholar 

  • Suppe, F. (1974). Some philosophical problems in biological speciation and taxonomy. In J. A. Wojcieckowski (Ed.) (1974). Conceptual basis of the classification of knowledge (pp. 190–243). Munich: Verlag Dokumentation.

    Google Scholar 

  • Suppe, F. (1977). The search for philosophical understanding of scientific theories. In F. Suppe (Ed.) (1977b). The structure of scientific theories (pp. 3–232). Champain: University of Illinois Press.

    Google Scholar 

  • Suppe, F. (1979). Theory structure. In Current research in philosophy of science (pp. 317–338). East Leansing: Philosophy of Science Association.

    Google Scholar 

  • Suppe, F. (1989). The semantic conception of theories and scientific realism. Urbana: University of Illinois Press.

    Google Scholar 

  • Suppes, P. (1960). A comparison of the meaning and uses of models in mathematics and the empirical sciences. Synthese, 12, 287–301. Reprinted in Suppes (1969), pp. 10–23.

    Article  Google Scholar 

  • Suppes, P. (1962). Models of data. In E. Nagel, P. Suppes, & Tarski, A. (Eds.), Logic, methodology and philosophy of science. Proceedings of the 1960 international congress. Stanford: Stanford University Press. Reprinted in P. Suppes (1969). Studies in the methodology and foundations of science: Selected papers from 1951 to 1969 (pp. 24–35). Dordrecht: Reidel.

    Google Scholar 

  • Suppes, P. (1967). What is a scientific theory? In S. Morgenbesser (Ed.), Philosophy of science today (pp. 55–67). New York: Basic Books.

    Google Scholar 

  • Suppes, P. (1969). Studies in the methodology and foundations of science: Selected papers from 1951 to 1969. Dordrecht: Reidel.

    Google Scholar 

  • Suppes, P. (1994). Answer to Wójcicki. In P. Humphreys (Ed.), Patrick Suppes: Scientific philosopher, II, theory, structure and measurement theory. Dordrecht: Kluwer Academic.

    Google Scholar 

  • Suppes, P. (2002). Representation and invariance of scientific structures, CSLI lecture notes. Stanford: Center for the Study of Language and Information.

    Google Scholar 

  • Tarski, A. (1933). The concept of truth in the languages of the deductive sciences. Expanded English translation in J. Corcoran (Ed.) (1983). Tarski: Logic, semantics, metamathematics, papers from 1923 to 1938 (pp. 152–278). Indianapolis: Hackett Publishing Company.

  • Thomson-Jones, M. (2006). Models and the Semantic View. Philosophy of Science, 73, 524–535.

    Article  Google Scholar 

  • van Fraassen, B. C. (1972). A formal approach to the philosophy of science. In R. G. Colodny (Ed.), Paradigms and paradoxes: The philosophical challenge of the quantum domain. Pittsburgh: Pittsburgh University Press.

    Google Scholar 

  • van Fraassen, B. C. (1980). Scientific image, Oxford: Clarendon Press.

    Book  Google Scholar 

  • van Fraassen, B. C. (1985). Empiricism in the philosophy of science. In P. M. Churchland, & C. A. Hooker (Eds.), Images of science: Essays on realism and empiricism, with a reply from Bas C. van Fraassen (pp. 245–308). Chicago: University of Chicago Press.

    Google Scholar 

  • van Fraassen, B. C. (1989). Laws and symmetry. Oxford: Clarendon Press.

    Book  Google Scholar 

  • van Fraassen, B. C. (1991). Quantum mechanics, an empiricist view. Oxford: Oxford University Press.

    Google Scholar 

  • van Fraassen, B. C. (2006). Representation: The problem for structuralism. Philosophy of Science, 73, 536–547.

    Article  Google Scholar 

  • van Fraassen, B. C. (2008). Scientific representation. Oxford: Oxford University Press.

    Book  Google Scholar 

  • Worrall, J. (1984). An unreal image. The British Journal of Philosophy of Science, 35(1), 65–80.

    Article  Google Scholar 

Download references

Acknowledgements

I am extremely grateful to Armond Duwell for extensive discussions on earlier versions of this paper. I also wish to thank Michael Dickson for perceptive comments, as well as Leah McClimans for pressing on the notion of adequacy in the last section. Finally, I wish to express my gratitude to two anonymous referees for pointing out the affinities between the views of the structuralist program and my own.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Soazig Le Bihan.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Le Bihan, S. Defending the Semantic View: what it takes. Euro Jnl Phil Sci 2, 249–274 (2012). https://doi.org/10.1007/s13194-011-0026-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s13194-011-0026-6

Keywords

Navigation