Skip to main content
Log in

Stabilization of phenomenon and meaning

On the London & London episode as a historical case in philosophy of science

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

Abstract

In recent years, the use of historical cases in philosophy of science has become a proper topic of reflection. In this article I will contribute to this research by means of a discussion of one very famous example of case-based philosophy of science, namely the debate on the London & London model of superconductivity between Cartwright, Suárez and Shomar on the one hand, and French, Ladyman, Bueno and Da Costa on the other. This debate has been going on for years, without any satisfactory resolution. I will argue here that this is because both sides impose on the historical case a particular philosophical conception of scientific representation that does not do justice to the historical facts. Both sides assume, more specifically, that the case concerns the discovery of a representational connection between a given experimental insight – the Meissner effect – and the diamagnetic meaning of London and London’s new equations of superconductivity. I will show, however, that at the time of the Londons’ publication, neither the experimental insight nor the meaning of the new equations was established: both were open for discussion and they were stabilized only later. On the basis of this historical discussion, I will then propose an alternative approach to the case study: the case should not be seen as a site of confrontation between pre-existing philosophical accounts, but rather as a way to historically elaborate and develop philosophical concepts. I will then show how approaching the historical episode in this way suggests an alternative approach to the philosophical study of representation, according to which it involves the establishment, over time, of a stable connection between constellations of different elements that, through discussion and engagement with alternative views and approaches, come to constitute phenomenon and meaning.

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.

Institutional subscriptions

Similar content being viewed by others

Notes

  1. Some of the central publications in this new debate are Burian (2001, 2002), Pitt (2001), Schickore (2011), Arabatzis and Schickore (2012) and Kinzel (2015) and the papers collected in Mauskopf and Schmaltz (2012) and Sauer and Scholl (2016).

  2. The London & London model has become a kind of topic in itself: not only CSS and FLBD, but also for example Landry (2007, p. 3), Schindler (2007), Morrison (2008, p. 75), Bailer-Jones (2009, p. 140), and Le Bihan (2012, p. 270) have discussed aspects of it in their work on scientific models.

  3. The original article consists of two parts. The first, by Cartwright, formulates their more general philosophical position, while the second, by Shomar and Suárez, concerns the case study (Cartwright et al. 1995, p. 137). Besides this article, I will also make use of later ones by Suárez (1999) and Cartwright and Suárez (2008). Even though not all of these articles are written by all three of them, I will still ascribe them all to CSS, following the tradition in the debate. The same holds for FLBD’s articles.

  4. On the syntactic view, theories are seen as collections of sentences, whereas on the semantic view, theories are described in terms of model-theoretic structures. Both offer a theory-driven view of science in the sense that both assume that it is theory that drives scientific research. For an overview of the debate between the syntactic and the semantic view, see Lutz (2017).

  5. Shomar and Suárez illustrate what they take to be theory-driven views of models through a discussion of Michael Redhead’s article ‘Models in Physics’ (Redhead 1980). In a later article, Suárez (1999) makes the same claim about Ernan McMullin’s article ‘Galilean Idealizations’ (McMullin 1985).

  6. As Michael Tinkham points out, “in many circumstances we expect absolutely no change in field or current to occur in times less than \(10^{10^{10}} years\)” (Tinkham 1996, p. 2).

  7. This was the first of a series of further articles (London 1935, 1936, 1937a, b; London and London 1935b).

  8. At the time, both the terms ‘superconductor’ and ‘supraconductor’ were used interchangeably.

  9. As examples of such proposals, the London brothers referred to articles by Becker et al. (1933), by Braunbek (1934), and by London (1934). The general intuition underlying these accounts was that “since superconductors do not obey Ohm’s law (steady-state current proportional to electric field) it is most natural to suppose that it is the acceleration of the current which is proportional to field” (Leggett 1995, p. 924). For a normal conductor we know that a current either induces an electric field or is supported by one. Ohm’s law then tells us that this current is directly proportional to the electric field: J = σE, where σ denotes the material’s conductivity. For superconducting materials this cannot be the case, for there we have a current (J) in the absense of an external field (E = 0). It was still commonly thought, however, that there was some kind of relation between the superconducting current and electric fields, an intuition that found its expression in the acceleration equation (Suárez 1999, p. 185).

  10. Λ is “the analog of specific resistance – i.e. a new characteristic constant depending on the material” (Dahl 1992, p. 229).

  11. A short discussion of the experiment by Kamerlingh Onnes and Willem Tuyn that was taken to confirm this belief can be found on page 15.

  12. Feynman (1964, chpt. 36) and Kittel (2005, chpt. 12) offer insightful accounts of what ferromagnetism comes down to.

  13. Feynman (1964, chpt. 34) and Kittel (2005, chpt. 11) offer good discussions of what diamagnetism comes down to.

  14. By local or particular knowledge, CSS mean knowledge that is not part of a larger theoretical whole: “a mediating model mediates between high level theory and the world by conveying some particular or local knowledge specific to the effect or phenomenon that is being modelled” (Suárez 1999, p. 170).

  15. The notion of a mediating model was first presented in the Models as Mediators-volume (Morgan and Morrison 1999), of which Suárez’s article is a part. While different authors stressed different aspects of the concept, all of them conceived of models as autonomous entities that mediate between theory and phenomena. The volume’s general aim was to collect instances of, and reflection on, models playing essential roles in scientific practice that were separate from any role they play in constituting theory (Cartwright and Suárez 2008, p. 64).

  16. Both sides also state this explicitly. Thus, CSS point out that “[m]ost of the original documents referred to by French and Ladyman are the same as those that we used in the 4-year long study that led to our account; and their reading of the history is the same as ours and as Gavroglu’s and Dahls on the points at issue” (Cartwright and Suárez 2008, p. 69). And in their latest articles, FLBD state that the main difference is maybe primarily of a philosophical nature (Bueno et al. 2012a, p. 46); (Bueno et al. 2012b, p. 103).

  17. An interesting side remark is that Suárez (1999, p. 188), in quoting this passage from the London brothers, stops right before the emphasized part.

  18. I prefer to use the more neutral terms ‘insight’ or ‘conception’ when talking about the diamagnetic idea, rather than ‘analogy’ or ‘knowledge’, since the question of how to characterize the insight is exactly what is at issue between CSS and FLBD. As will become clear, I do not think that approaching the results of Meissner and Ochsenfeld’s experiments in this way, i.e. in terms of the knowledge it provides, does justice to the case. I would like to thank an anonymous reviewer for pressing me on this point.

  19. The original, in German, goes as follows: “Durch die Gleichung (8) [∇×ΛJ = −H/c] wird der Supraleiter gekennzeichnet als ein einziges grosses diamagnetisches Atom” (London and London 1935b, p. 348).

  20. A few years later Fritz London did draw a direct link between the Meissner effect and diamagnetism: “Meissner and Ochsenfeld found in 1933, that a supraconductor behaves not only like an ideal conductor, but in addition also like a very strong diamagnetic metal” (London 1937a, p. 793). This is something that will be discussed in footnote 42 on page 22. I would like to thank an anonymous reviewer for pushing me to specify this point.

  21. Proposals similar to Schickore’s are made by Chang (2012), Kinzel (2015, 2016), Pietsch (2016), Scholl and Räz (2016) and Knuuttila and Loetgers (2016). Some of them argue that we should completely abandon the confrontational model, and use history merely to develop our philosophical concepts, not as evidence for philosophical theories. Others claim that the distinction should not be seen in such absolute terms: the use of case studies in philosophy is, on their view, both evidential and hermeneutic, i.e. it can serve both to argue for general philosophical claims and to elaborate a better understanding of them. My aim here is not to argue for either of these two options. The only point I wish to make is that, as it stands, the debate between CSS and FLBD has been conducted primarily in terms of the confrontational model, and that therefore, employing the alternative hermeneutic approach could help us in furthering the debate.

  22. Dahl (1992, p. 164 – 167) offers an extensive account of the irreversibility-idea, with very insightful illustrations. He also contrasts it, in a very clear way, with the idea of reversibility (Dahl 1992, p. 179 – 181). See also Matricon and Waysand (2003, p. 55).

  23. Dahl is referring here to Gabriel Jonas Lippmann, who is famous for Lippmann’s rule, a rule that was long thought to govern the behaviour of all superconductors: “The rule, published by Lippmann in 1919, was actually stimulated by [Onnes’] 1914 experiments and is simply the consequence of Maxwell’s electrodynamics for perfect conductors. It states that the magnetic flux linking any closed circuit within a body of zero resistance […] cannot change; circulating currents are induced on the surface so as to create a magnetic flux density in the interior that cancels the flux density due to the applied field” (Dahl 1992, p. 102). For a technical discussion of Lippmann’s theorem, see (Essén and Fiolhais 2012, p. 165 – 166).

  24. Dahl offers an extensive account of these experiments, and the discussion that followed when they were presented in 1924 (Dahl 1992, p. 106 – 110). It turned out, however, that these experiments were misleading, because of the hollowness of the sphere (Dahl 1992, p. 164), which made it difficult to carry out precise measurements (Matricon and Waysand 2003, p. 55).

  25. Dahl (1992, p. 177 – 181) offers an extensive discussion of the context and the material set-up of these experiments.

  26. Gorter and Casimir’s motivation for studying the phenomenon thermodynamically derived from different theoretical and experimental investigations of superconductivity: they referred to work by Willem Hendrik Keesom and J. N. van den Ende, Wander Johannes de Haas and collaborators, Meissner and Ochsenfeld, and de Haas and Josina M. Casimir (Gorter and Casimir 1934, p. 306 – 307).

  27. One thermodynamic characteristic of the phenomenon of superconductivity is that the transition does not involve any change of latent heat (i.e. energy required to change the state — gas, liquid, solid — of the material) or of volume, but only of specific heat (i.e. energy required to change a unit mass of the material by one unit temperature). In 1934, Arend Joan Rutgers presented an equation which relates this specific heat change to the critical magnetic field (Smith and Wilhelm 1935, p. 261); (Dahl 1992, p. 155).

  28. The original in German goes as follows: “wir [haben] jetzt erfahren, dass ausserdem die magnetische Induktion im Supraleiter verschwindet, wenigstens sofern wir den noch sehr unsicheren experimentellen Befund so auf einen idealen Gehalt hin interpretieren dürfen, wie es durch die thermodynamische Diskussion durch Gorter (1933a, b) und Gorter und Casimir (1934) nahegelegt wird”.

  29. The main reason for this uncertainty about the precise interpretation of the experiments was, in part, the “extreme brevity, seeming contradictions, and obvious importance” of Meissner and Ochsenfeld’s paper. (Dahl 1992, p. 182). It also was a consequence, however, of the results of the first experimental attempts to replicate it. Babbitt and Mendelssohn’s experiments, for example, “neither disproved the classical theory nor confirmed the Meissner effect with any certitude” (Dahl 1992, p. 189), a claim that was repeated by researchers in Toronto carrying out similar experiments (Dahl 1992, p. 190). And Keesom and Johannes Antonie Kok stated that Meissner and Ochsenfeld’s results entailed primarily that “one must be cautious as to the appreciation of the value of the magnetic field in the neighbourhood of a supraconductor” (Keesom and Kok 1934, p. 503). It was only with Rjabinin and Schubnikow (1935), who “reported perhaps the clearest confirmation of the Meissner effect”, that the results started to get accepted (Dahl 1992, p. 193).

  30. A perfect conductor is an idealized material that is characterized in terms of Maxwell’s equations with infinite conductivity, i.e. zero resistance (R = 0) (Tinkham 1996, p. 2).

  31. The precise workings of the theorem are discussed by, among others, Smith and Wilhelm (1935, p. 266 – 267), Dahl (1992, 151 – 153), Leggett (1995, p. 917 – 918) and Schmalian (2011, p. 47). For a broader historical discussion of the theoretical study of conductivity, see Hoddeson and Baym (1980) and Hoddeson et al. (1987).

  32. Fritz London himself formulated Bloch’s result as follows: “it is rigorously demonstrable that, on the basis of the recognized conceptions of the electron theory of metals, a theory of supraconductivity is impossible” (London 1935, p. 24).

  33. In their review article of 1935, Smith and Wilhelm ((Smith and Wilhelm 1935), p. 260 – 261) offered a discussion of both approaches. For more contemporary discussions, see Hoddeson et al. (1987, p. 30 – 31), Gavroglu and Goudaroulis (1989, p. 75 – 77), Dahl (1992, p. 151 – 153), Leggett (1995, p. 917), Matricon and Waysand (2003, p. 42) and Schmalian (2011, p. 45 – 48).

  34. There was quite an interest in the penetration depth because it promised to offer more insight into the nature of superconductivity: “the penetration depth is the region where the magnetic contribution to the free energy density changes from its “normal” or “vacuum” value to that appropriate for the superconductor, and its dependence on various physical parameters can provide a sensitive test of the quantitative aspects of any theory of superconductivity” (Chandrasekhar 1969, p. 13). One important study carried out in order to investigate the penetration depth was in fact Meissner and Ochsenfeld’s experiment, since Meissner believed that this could lead to further theoretical insight (Dahl 1992, p. 174); (Leggett 1995, p. 916).

  35. See e.g. Casimir (1977, p. 488), Dahl (1992, p. 174 – 175), Leggett (1995, p. 924) or Matricon and Waysand (2003, p. 68 – 70) for discussions of their work.

  36. The original German goes as follows: “Wir wollen in diesem Abschnitt untersuchen, wie die oben gefundene Lösung für Supraleiter sich aus derjenigen für gewöhnliche Leiter mit endlicher Leitfähigkeit σ durch den Grenzübergang σ ergibt”.

  37. Heinz London already had quite some understanding of the equation, because in his PhD thesis, he elaborated “a notion of finite penetration depth in superconductors due to the inertia of electrons, similar in concept to that of Becker, Heller, and Sauter” (Dahl 1992, p. 226). Fritz was very enthousiastic about this work (Dahl 1992, p. 228). See also Matricon and Waysand (2003, p. 68 – 70) for a discussion of the relation between Heinz’s PhD and the work by Becker, Heller, and Sauter.

  38. A similar reading of the Londons’ work can also be found in Wilson’s (1936) review article. He described the Londons’ work as offering “the most promising line of attack” for understanding the mechanism of superconductivity (Wilson 1936, p. 269), but he also noted that “[t]he great defect in the theory is that it is purely formal, and it is probable that no great advance can be made without a microscopic derivation of Eq. 8 or an understanding of what happens during the transition to the superconducting state” (Wilson 1936, p. 270).

  39. The original German goes as follows: “Im Gegensatz zu der üblichen Vorstellung, dass im Supraleiter Ströme ohne Mitwirkung eines Feldes bestehen können, ist in der hier vorgeschlagenen Formulierung der Suprastrom als eine Art diamagnetischer Volumenstrom gekennzeichnet. Der Suprastrom wird durch ein Magnetfeld aufrechterhalten, welches seinerseits von dem Strome selbst erzeugt sein kann.”

  40. The original French goes as follows: “On sait que le diamagnétisme est une particularité de la mécanique quantique. On ne peut expliquer ce phénomène par aucun mécanisme de la mécanique classique.”

  41. The original German goes as follows: “Betrachten wir nun die Gleichungen [(9)], so ist zu bemerken, dass die Analogie mit dem diamagnetischen Atom nicht zu wörtlich genommen werden kann. Zur Begründung von [(9)] wäre es notwendig zu zeigen, dass die Eigenfunktion des Supraleiters auch durch ein elektrostatisches Feld in erster Näherung keine Störung erfährt. Bei einem gewöhnlichen diamagnetischen Atom ist dies nicht zutreffend. Die Eigenfunktion eines solchen erfährt vielmehr eine Störung, welche proportional der angreifenden elektrischen Feldstärke ist.”

  42. This may now also explain the fact, pointed out in footnote 20, that Fritz London, in his London (1937a) article, drew a much closer connection between Meissner and Ochsenfeld’s experiment and diamagnetism than in their earlier work, since by that time his work on the diamagnetic programme had already progressed much further.

  43. See especially Cartwright and Suárez (2008, p. 64 – 68) and Bueno et al. (2012b, p. 102) for a thorough discussion of how both sides interpret the idea of independence from theory.

  44. Meissner repeats this claim during a presentation of his later results at the Royal Society (Meissner 1935, p. 15).

  45. A similar claim can be found in Smith and Wilhelm’s (1935, p. 270) review article, in which they presented their vortex current theory. This theory, they claimed, seems to be experimentally indistinguishable from the Londons’ theory, even though it retains the local validity of the acceleration equation.

  46. Feest’s conception of phenomena here is influenced by Bogen and Woodward’s (1988) distinction between data and phenomena. On Feest’s view, surface phenomena are to be found in the data collected in a particular experiment, while hidden phenomena are the phenomena that are instantiated by these data and explained by theory.

  47. This approach suggests, for example, the following topics of further historical-philosophical research: the way in which Meissner himself, in collaboration with Heidenreich after Ochsenfeld moved to a new academic position, attempted to replicate his own experiments, and how these were received within the community at the time (see Dahl 1992, [p. 195 – 208]); the way in which the Londons’ new account led to a prediction for the precise penetration depth of the superconducting current, and how this was tested experimentally by David Shoenberg (Matricon and Waysand 2003, p. 64); or how the Londons’ (London and London 1935a) work led Heinz to elaborate a theory that provided a way to study the differences between superconducting pure metals and superconducting alloys (Dahl 1992, p. 230 – 231). All these topics allow us to study, more specifically, how both skill and validation contributed to the emergence of a stable fit between experimental results and theoretical interpretations over time.

References

  • Arabatzis, T., & Schickore, J. (2012). Ways of integrating history and philosophy of science. Perspectives on Science, 20(4), 395–408.

    Article  Google Scholar 

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

  • Becker, R., Sauter, F., Heller, G. (1933). Uber die Stromverteiling einer Supraleitenden Kugel. Zeitschrift für Physik, 85(11), 772–787.

    Article  Google Scholar 

  • Bogen, J., & Woodward, J. (1988). Saving the phenomena. The Philosophical Review, XCVII(3), 305–352.

    Google Scholar 

  • Braunbek, W. (1934). Die Ausbreiting elektromagnetischer Wellen in einem Supraleiter. Zeitschrift für Physik, 87(7), 470–483.

    Article  Google Scholar 

  • Bueno, O., & French, S. (2011). How theories represent. British Journal for Philosophy of Science, 62(4), 857–894.

    Article  Google Scholar 

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

    Article  Google Scholar 

  • Bueno, O., French, S., Ladyman, J. (2012a). Models and structures: phenomenological and partial. Studies in History and Philosophy of Modern Physics, 43, 43–46.

    Article  Google Scholar 

  • Bueno, O., French, S., Ladyman, J. (2012b). Empirical factors and structure transference: returning to the London account. Studies in History and Philosophy of Modern Physics, 43, 95–104.

    Article  Google Scholar 

  • Burian, R.M. (2001). The dilemma of case studies resolved: the virtues of using case studies in the history and philosophy of science. Perspectives on Science, 9(4), 383–404.

    Article  Google Scholar 

  • Burian, R.M. (2002). Comments on the precarious relationship between history and philosophy of science. Perspectives on Science, 10(4), 398–407.

    Article  Google Scholar 

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

    Article  Google Scholar 

  • Cartwright, N., Shomar, T., Suárez, M. (1995). The tool box of science: tools for the building of models with a superconductivity example. Poznán Studies in the Philosophy of the Sciences and the Humanities, 44, 137–149.

    Google Scholar 

  • Casimir, H. (1977). Superconductivity and superfluidity. In Mehra, J. (Ed.) The physicist’s conception of nature. Springer.

  • Chandrasekhar, B.S. (1969). Early experiments and phenomenological theories. In Parks, R. (Ed.) Superconductivity, Vol. 1. Marcel Dekker, Inc.

  • Chang, H. (2012). Beyond case-studies: history as philosophy. In: (Mauskopf and Schmaltz 2012).

  • Da Costa, N., & French, S. (2000). Models, theories, and structures: thirty years on. Philosophy of Science, 57, 116–127.

    Article  Google Scholar 

  • Dahl, P.F. (1992). Superconductivity: its historical roots and development from mercury to the ceramic oxides. American Institute of Physics.

  • Essén, H., & Fiolhais, M.C.N. (2012). Meissner effect, diamagnetism, and classical physics – a review. American Journal of Physics, 80, 164–169.

    Article  Google Scholar 

  • Feest, U. (2011). What exactly is stabilized when phenomena are stabilized? Synthese, 182, 57–71.

    Article  Google Scholar 

  • Feynman, R. (1964). The Feynman lectures on physics volume II: mainly electromagnetism and matter. Adison-Wesley.

  • French, S. (1999). The phenomenological approach to physics. Studies in History and Philosophy of Modern Physics, 30, 267–281.

    Article  Google Scholar 

  • French, S., & Ladyman, J. (1997). Superconductivity and structures: revisiting the London account. Studies in History and Philosophy of Modern Physics, 28(3), 363–393.

    Article  Google Scholar 

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

    Article  Google Scholar 

  • French, S. (2010). Keeping quiet on the ontology of models. Synthese, 172(2), 231–249.

    Article  Google Scholar 

  • Gavroglu, K. (1995). Fritz London: a scientific biography. Cambridge University Press.

  • Gavroglu, K., & Goudaroulis, Y. (1989). Methodological aspects of the development of low temperature physics 1881–1956: concepts out of context(s). Kluwer Academic Publishers.

  • Gorter, C. (1933a). Some remarks on the thermodynamics of supraconductivity. Archives du Musèe Teyler, 7, 378–386.

    Google Scholar 

  • Gorter, C. (1933b). Theory of supraconductivity. Nature, 132, 931.

    Article  Google Scholar 

  • Gorter, C.J. (1964). Superconductivity until 1940 in Leiden and as seen from there. Reviews of Modern Physics, 36(1), 3–7.

    Article  Google Scholar 

  • Gorter, C., & Casimir, H. (1934). On supraconductivity I. Physica, 1, 306–320.

    Article  Google Scholar 

  • Hoddeson, L.H., & Baym, G. (1980). The development of the quantum mechanical electron theory of metals: 1900-28. Proceedings of the Royal Society of London A, 371 (1744), 8–23.

    Article  Google Scholar 

  • Hoddeson, L.H., Baym, G., Eckert, M. (1987). The development of the quantum mechanical electron theory of metals: 1928–1933. Reviews of Modern Physics, 59(1), 287–327.

    Article  Google Scholar 

  • Kamerlingh Onnes, H. (1913). Commun. Kamerlingh Onnes Lab., Tech. Rep. 34b, Univ. Leiden.

  • Kamerlingh Onnes, H., & Tuyn, W. (1924). Nouvelles expériences avec les supraconducteurs. In Institut international de physique Solvay, (Ed.) Conductibilité électrique des métaux et problèmes connexes: rapports et discussions du quatrième Conseil de physique tenu à Bruxelles du 24 au 29 avril 1924. Gauthier-Villars.

  • Keesom, W.H., & Kok, J.A. (1934). Measurement of the latent heat of thallium connected with the transition, in a constant external magnetic field, from the supraconductive to the non-supraconductive state. Commun. Kamerlingh Onnes Lab., Tech. Rep. 203e, Univ. Leiden.

  • 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, 49, 48–57.

    Article  Google Scholar 

  • Kinzel, K. (2016). Pluralism in historiography: a case study of case studies. In (Sauer and Scholl 2016).

  • Kittel, C. (2005). Introduction to solid state physics, 8th Edn. Wiley.

  • Knuuttila, T., & Loetgers, A. (2016). Contrasting cases: the Lotka-Volterra model times three. In (Sauer and Scholl 2016).

  • Ladyman, J. (2002). Understanding philosophy of science. Routledge.

  • Landry, E. (2007). Shared structure need not be shared set-structure. Synthese, 158, 1–17.

    Article  Google Scholar 

  • Le Bihan, S. (2012). Defending the semantic view: what it takes. European Journal for Philosophy of Science, 2, 249–274.

    Article  Google Scholar 

  • Leggett, A.J. (1995). Superfluids and superconductors. In Brown, L., Pippard, B., Pais, A. (Eds.) Twentieth century physics, Vol. 2. CRC Press.

  • London, H. (1934). Production of heat in supraconductors by alternating currents. Nature, 133, 497–498.

    Article  Google Scholar 

  • London, F. (1935). Macroscopical interpretation of supraconductivity. Proceedings of the Royal Society A, 152(875), 24–34.

    Google Scholar 

  • London, F. (1936). Electrodynamics of macroscopic fields in supraconductors. Nature, 137, 991–992.

    Article  Google Scholar 

  • London, F. (1937a). A new conception of supraconductivity. Nature, 140, 793–796, 834–836.

    Article  Google Scholar 

  • London, F. (1937b). Une conception nouvelle de la supra-conductibilité: conférences faites à l’Institut Poincaré, Paris. Hermann.

  • London, F., & London, H. (1935a). The electromagnetic equations of the supraconductor. Proceedings of the Royal Society, A149, 71–88.

    Google Scholar 

  • London, F., & London, H. (1935b). Supraleitung und Diamagnetismus. Physica, 2(1), 341–354.

    Article  Google Scholar 

  • Lorentz, H.A. (1924). Application de la théorie des électrons aux propriétés des métaux. In Institut international de physique Solvay (Ed.), Conductibilité électrique des métaux et problèmes connexes: rapports et discussions du quatrième Conseil de physique tenu à Bruxelles du 24 au 29 avril 1924. Gauthier-Villars.

  • Lutz, S. (2017). What was the syntax-semantics debate in the philosophy of science about? Philosophy and Phenomenological Research, 95(2), 319–352.

    Article  Google Scholar 

  • Matricon, J., & Waysand, G. (2003). The cold wars: a history of superconductivity. Translated from French by Charles Glashausser. Rutgers University Press.

  • Mauskopf, S., & Schmaltz, T. (2012). Integrating history and philosophy of science: problems and prospects. Springer.

  • Meissner, W. (1935). The magnetic effects occurring on transition to the supraconducting state. Proceedings of the Royal Society A, 152(875), 13–15.

    Google Scholar 

  • Meissner, W., & Ochsenfeld, R. (1933). Ein neuer Effekt bei Eintritt der Supraleitfähigkeit. Die Naturwissenschaften, 21, 787–788. In Forrest, A. (1987). Meissner and Ochsenfeld Revisited. European Journal of Physics 4 (2): 117-120.

    Article  Google Scholar 

  • Mendelssohn, K., & Babbitt, J.D. (1935). Magnetic behaviour of supraconducting tin spheres. Proceedings of the Royal Society of London, 151(873), 316–333.

    Google Scholar 

  • McMullin, E. (1985). Galilean idealizations. Studies in History and Philosophy of Science, 16(3), 247–273.

    Article  Google Scholar 

  • Monaldi, D. (2017). Fritz London and the scale of quantum mechanisms. Studies in History and Philosophy of Modern Physics, 60, 35–45.

    Article  Google Scholar 

  • Morgan, M., & Morrison, M. (1999). Models as mediators: perspectives on natural and social science. Cambridge University Press.

  • Morrison, M. (2008). Models as representational structures. In Hartmann, S. et al. (Eds.) Nancy Cartwright’s philosophy of science. Routledge.

  • Pietsch, W. (2016). Two modes of reasoning with case studies. In (Sauer and Scholl 2016).

  • Pitt, J.C. (2001). The dilemma of case studies: towards a heraclitian philosophy of science. Perspectives on Science, 9(4), 373–382.

    Article  Google Scholar 

  • Redhead, M. (1980). Models in physics. The British Journal for the Philosophy of Science, 31(2), 145–163.

    Article  Google Scholar 

  • Rjabinin, J.N., & Schubnikow, L. (1935). Über die Abhängigkeit der magnetischen Induktion des supraleitenden Blei vom Feld. Physikalischer Zeitschrift der Sowjet Union, 6, 557–568.

    Google Scholar 

  • Sauer, T., & Scholl, R. (2016). The philosophy of historical case studies. Springer.

  • Schickore, J. (2011). More thoughts on HPS: another 20 years later. Perspectives on Science, 19(4), 453–481.

    Article  Google Scholar 

  • Schindler, S. (2007). Rehabilitating theory: refusal of the ‘bottom-up’ construction of scientific phenomena. Studies in History and Philosophy of Science, 38, 160–184.

    Article  Google Scholar 

  • Schmalian, J. (2011). Failed theories of superconductivity. In Cooper, L.N., Feldman, D, Schmalian, J (Eds.) 50 Years: World Scientific Publishing Co.

  • Scholl, R, & Räz, T. (2016). Towards a methodology for integrated history and philosophy of science. In (Sauer and Scholl 2016).

  • Smith, H.G., & Wilhelm, J.O. (1935). Superconductivity. Reviews of Modern Physics, 7(4), 237–271.

    Article  Google Scholar 

  • Suárez, M. (1999). The role of models in the application of scientific theories: epistemological implications. In (Morgan and Morrison 1999).

  • Tinkham, M. (1996). Introduction to superconductivity, 2nd Edn. McGraw-Hill.

  • Wilson, A.H. (1936). Superconductivity and the theory of metals. Reports on Prgoress in Physics, 3, 262–271.

    Article  Google Scholar 

Download references

Acknowledgements

The author would like to thank Bert Leuridan, Mieke Boon, Brandon Boesch, Laura Georgescu, and the audiences at the OZSW Graduate Conference in Theoretical Philosophy 2016 (University of Twente), the GRAT Workshop 2016 (University of Antwerp), and the SPSP Conference 2016 (Rowan University).

The author would like to acknowledge the Research Foundation – Flanders (FWO) as funding institution.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Jan Potters.

Additional information

Publisher’s note

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

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Potters, J. Stabilization of phenomenon and meaning. Euro Jnl Phil Sci 9, 23 (2019). https://doi.org/10.1007/s13194-019-0247-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s13194-019-0247-7

Keywords

Navigation