Skip to main content
Log in

The introduction of topology into analytic philosophy: two movements and a coda

  • Original Research
  • Published:
Synthese Aims and scope Submit manuscript

Abstract

Both early analytic philosophy and the branch of mathematics now known as topology were gestated and born in the early part of the 20th century. It is not well recognized that there was early interaction between the communities practicing and developing these fields. We trace the history of how topological ideas entered into analytic philosophy through two migrations, an earlier one conceiving of topology geometrically and a later one conceiving of topology algebraically. This allows us to reassess the influence and significance of topological methods for philosophy, including the possible fruitfulness of a third conception of topology as a structure determining similarity.

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

Data availability

Not applicable.

Code availability

Not applicable.

Notes

  1. The first volume of Philosophy of Science had been published only the year before.

  2. See also Mormann (2013, pp. 423, 426). Although Mormann emphasizes the philosophy of science, he evidently understands this broadly, including within it the philosophy of geometry and space. This is why there is no harm in his unqualified reference to “philosophy,” and we shall often follow this practice.

  3. This is a common conception of topology among philosophers. For instance, Callender and Weingard write that “Topology is a kind of abstraction from metrical geometry” (1996, p. 21).

  4. See also Mormann (2007, 2008) respectively, for more on his view of Russell and Carnap’s use of topological and geometrical ideas.

  5. See also Mormann (2013, p. 431): “neither Russell nor any other philosophers of science ever took notice of.

    the path-breaking work of the American mathematician Marshall H. Stone.” Russell may not be as culpable as Mormann states, since by the time Stone had announced and published his results in 1934–1937, Russell was no longer actively working on these topics.

  6. Although this term finds its origin in the work of Leibiniz, by the time of Russell and Carnap’s writings, it had become synonymous with topology. Indeed, Henri Poincare’s influential early work in topology bore the title “Analysis Situs” (Poincaré, 1895). This work has been noted as an originator of modern topological concepts; mathematician and historian of mathematics Jean Dieudonné has claimed that before this work of Poincaré we can only speak of the pre-history of the discipline (Dieudonné, 2009, p. 15). Oswald Veblen’s textbook of topology was also titled Analysis Situs (Veblen, 1922). Additionally, it is clear from the context in Analysis of Matter that Russell took the terms to be synonymous; he first mentions analysis situs immediately prior to introducing Hausdorff’s definitions (Russell, 1927, pp. 295–96).

  7. Key examples include An Essay on the Foundations of Geometry (Russell, 1897), which Russell wrote on fellowship shortly after completing his BA in mathematics (Wahl, 2018: pp. x–xiv), and his earlier work “The Logic of Geometry” (Russell, 1896), which concerns the nature of the Euclidean axiom of congruence. He argues (against Helmholtz) that this axiom is strictly a priori and as such it cannot be proven through experience of objects in space. As a result, he concludes that rejecting this axiom is absurd from both a logical and philosophical standpoint. In support of these results, Russell provides argumentation that is both philosophical and mathematical.

  8. Here, Russell credits Whitehead with this approach to points without citing a specific work. But as noted by Mormann (2013, p. 430), Russell elsewhere credits a specific work by Whitehead in reference to his construction of points: the fourth volume of the Principia. This work never saw publication. As a result it is not possible to consult Whitehead’s point construction in his own words, at least not in that particular work. However, in his “On the Mathematical Concepts of the Material World” (Whitehead, 1906), we do find a strikingly similar construction to the one recited in OKEW. Whitehead’s construction of points (Whitehead, 1906, pp. 488–492) builds points out of a theory of interpoints, which are similar to the enclosures described by Russell. Whitehead’s techniques did not involve topological concepts. Incidentally, that work by Whitehead cites Veblen, who would later become well known for his work in topology. However, the citations are to Veblen’s dissertation which concerned geometry and predated his work as a topologist. There does not seem to be evidence that the topological notions employed by Russell later in The Analysis of Matter had their origin in Whitehead’s point construction.

  9. Technically speaking, this definition captures what we would now call a Hausdorff neighborhood basis. Condition (D) is the Hausdorff condition, while a neighborhood system is generated by closing the basis under the superset relation.

  10. Indeed, his most famous student was Kurt Gödel.

  11. For the most focused critical discussion of Russell’s scholarship regarding the construction of points, see Mormann (2008). Note also that Russell’s attempted construction of points bears resemblance to later approaches to mereotopology using (ultra)filters, for which see footnote 22.

  12. It is worthwhile to acknowledge that one of Carnap’s early philosophical influences came not from the analytic tradition as he explicitly invoked the Husserlian concept of Wesenserschauung in his characterization of intuitive space in “Der Raum” (Ryckman, 2007, p. 103; Friedman, 2019, p. 204). Nevertheless, historians also give good reason to understand the early Carnap as operating squarely in the early analytic tradition. Gottfried Gabriel points out that Carnap himself gave “greatest tribute” to Frege out of his influential teachers during his youthful time in Jena and Frieberg from 1910 to 1914 (Gabriel, 2007, pp. 65–66), among other nuances of the relationship between Carnap and the “grandfather of analytic philosophy.” Additionally, in his editorial notes that accompany a new translation of “Der Raum,” Michael Friedman cites the shadow cast by Russell and Whitehead’s Principia in Carnap’s own pointers to relevant literature in his account of formal space (Friedman, 2019, p. 174). If we take Russell to be a progenitor of analytic philosophy, then his influence on Carnap cited here is pertinent when attempting to categorize Carnap’s early work.

  13. Space: A Contribution to the Theory of Science.

  14. Foundations of Geometry.

  15. Carnap, like his predecessors, does not clearly distinguish between a manifold’s topological and differentiable structure, the distinction between which would only come about a decade later (Veblen and Whitehead, 1931, 1932). Nevertheless, even though formal space has the standard differentiable structure associated with products of the real line, Carnap does focus his attention on its topological properties.

  16. It is worth noting that Friedman argues that Carnap’s approach is substantially flawed. But, for our present purposes we need not weigh in on the efficacy of the structure of intuitive space in Der Raum (Friedman, 2019, pp. 187–188).

  17. This work was later translated as The Axiomatization of the Theory of Relativity (Reichenbach, 1965).

  18. For more on geometrical themes in Carnap’s early work, see Mormann (2007).

  19. We agree with Mormann’s comments insofar as they apply to geometric topology specifically. In this sense, his assertions apply in the context of the historical figures of interest in this section (Russell and Carnap). However, we will see in Sect. 3 that this view of the relationship between geometry and topology is not conceptually required by the features of topology. Starting in the 1930s and more so in the 1940s, the connections between topology and algebra became pronounced.

  20. Indeed, Poincaré’s famous namesake conjecture is a noteworthy early problem in topology.

  21. It was also in that year that Whitehead moved to the philosophy department at Harvard, and as noted in the previous subsection, Whitehead influenced Russell’s engagement with topological ideas. It would be worth investigating further whether Whitehead had any influence on Franklin or Wiener.

  22. For another account of topological regions that includes a study of maps between regions, see Roeper (1997).

  23. Roeper also cites an earlier predecessor, by Arnold Johanson, that takes a similar approach using category theory (Johanson, 1981). Johanson cites Whitehead and Russell’s TAM, among others, as inspiration.

  24. Stone himself had called them Boolean spaces for obvious reasons (Johnstone, 1982, p. xvi).

  25. Alternatively, one can proceed from the Boolean algebra’s prime ideals, which are dual to the ultrafilters and more faithful to Stone’s original presentation (Johnstone, 1982, pp. xv–xvi).

  26. For instance, they are exactly the spaces that are totally separated and compact. See Johnstone (1982, pp. 69–70) for a number of other, equivalent characterizations.

  27. According to Google Scholar, Stone’s article (Stone, 1937) has over 1800 citations, and each of these subsequent works (Jonsson and Tarski, 1951; Wallman, 1938) has hundreds of citations each.

  28. Grosholz (1985, Sec. 2; 2007, chap. 10.3) provides a comparison of Tarski and Stone’s work in this area and offers an account of the resulting significance of topology for logic as a branch of mathematics. In particular, she comments on the initial limitations of the analogy between topology and logic, which was initially restricted to quite limited logics and topological spaces, that had to be overcome for it to become widely fruitful.

  29. Specifically, theorem 4.11 in Tarski (1956) and theorem 7 in Stone (1938).

  30. What follows is not a comprehensive statement of the extensive publications by these authors and other scholars working on topological formal learning theory and epistemic logic. Rather, we intend merely to introduce the reader to one fruitful branch of intellectual development that has arisen from the introduction of topology into philosophy.

  31. Again, this just scratches the surface of the application of topology in mathematical logic. See also, for instance, Grosholz (2007, chap. 10.5) for an accessible treatment of some connections between topology and model theory, and Aiello et al. (2007) for a variety of surveys of spatial logics that employ topology.

  32. For instance, one of Kelly’s main theses is that computability should play as important a role in formal epistemology and induction in the philosophy of science as probability enjoys (Kelly, 1996, p. 8; Suppes, 1998, p. 351).

  33. In this chapter on topology, Kelly also draws on theoretical resources found in the literature related to recursion and decidability (Gold, 1965; Kugel, 1977; Putnam, 1965). Although Kelly’s application of these ideas is interesting, our focus in this section remains on the use of topology as facilitated by the conceptual connections derived from Stone’s theorem. For more on the connections between recursion theory, topology, and the Borel hierarchy, see Grosholz (1985, Sec. 3; 2007, chap. 10.4).

  34. Additionally, van Benthem gestured towards the future directions that might be taken by combining aspects of his project in Logical Dynamics of Information and the formal learning theory of LRI (van Benthem, 2011, p. 249).

  35. Mormann also draws another contrast, that “20th century philosophy of science showed no interest in topology as an object of philosophical reflection. There has been no ‘philosophy of topology’ in analogy to disciplines such as ‘philosophy of physics’, ‘philosophy of biology’, or ‘philosophy of geometry’” (Mormann, 2013, p. 425). However, the analogy is not as strong as it might at first seem, since physics and biology are distinct sciences, while geometry historically has been a partly mathematical, partly empirical discipline. Topology, by contrast, is clearly a subdiscipline of mathematics.

  36. Mormann (2013, pp. 428, 431) does discuss Stone’s theorem, including how it connects logic with topology, but seemingly only in the context of a geometric conception of topology. See also Mormann (2020, Sec. 3).

  37. Mormann (2020, Sec. 5) also shows how to formulate different versions of Leibnizian identity principles for objects in topological terms (i.e., using topological separation axioms). We also believe that the topological structure employed in such cases should be interpreted in terms of similarity (of the objects involved) instead of in terms of geometry.

  38. We defer to Fletcher (forthcoming, Sec. 2) for the technical details.

  39. Mormann (2020, p. 144) acknowledges this particular interpretation: “Every conceptual space is endowed with a metric that measures the similarity between objects.” But he does not seem to extend that interpretation to topology more generally.

References

  • Abramov, B. W. (1973). Geometrical structures in science. PhD Dissertation, Bloomington, IN: Indiana University.

  • Abramsky, S. (1987). Domain theory and the logic of observable properties. PhD Dissertation, London: Queen Mary College, University of London.

  • Aiello, M., Pratt-Hartmann, I., & van Benthem, J. (2007). Handbook of Spatial Logics. Dordrecht: Springer.

    Book  Google Scholar 

  • Baltag, A., Bezhanishvili, N., & Fernández González, S. (2019). The McKinsey-Tarski theorem for topological evidence logics. In R. Iemhoff, M. Moortgat, R. de Queiroz, (Eds.), Logic, Language, Information, and Computation, (pp. 177–194). Lecture Notes in Computer Science 11541. Berlin, Heidelberg: Springer.

  • Baltag, A., Bezhanishvili, N., Özgün, A., & Smets, S. (2013). The topology of belief, belief revision and defeasible knowledge. In D. Grossi, O. Roy, & H. Huang (Eds.), Logic, Rationality, and Interaction, (pp. 27–40). Lecture Notes in Computer Science 8196. Berlin, Heidelberg: Springer.

  • Baltag, A., Bezhanishvili, N., Özgün, A., & Smets, S. (2016a). Justified belief and the topology of evidence. In J. Väänänen, Å. Hirvonen, & R. de Queiroz (Eds.), Logic, Language, Information, and Computation, (pp. 83–103). Lecture Notes in Computer Science 9803. Berlin, Heidelberg: Springer.

  • Baltag, A., Bezhanishvili, N., Özgün, A., & Smets, S. (2017). The topology of full and weak belief. In H. H. Hansen, S. E. Murray, M. Sadrzadeh, & H. Zeevat (Eds.), Logic, Language, and Computation, (pp. 205–228). Lecture Notes in Computer Science 10148. Berlin, Heidelberg: Springer.

  • Baltag, A., Gierasimczuk, N., & Smets, S. (2016b). On the solvability of inductive problems: A study in epistemic topology. Electronic Proceedings in Theoretical Computer Science, 215(June), 81–98.

    Article  Google Scholar 

  • Bartha, P. F. A. (2010). By Parallel Reasoning: The Construction and Evaluation of Analogical Arguments. Oxford University Press.

    Book  Google Scholar 

  • Beklemishev, L., & Gabelaia, D. (2014). Topological interpretations of provability logic. In G. Bezhanishvili (Ed.), Leo Esakia on Duality in Modal and Intuitionistic Logics, (pp. 257–290). Outstanding Contributions to Logic 4. Dordrecht: Springer Netherlands.

  • Belot, G. (2013). Bayesian orgulity. Philosophy of Science, 80(4), 483–503.

    Article  Google Scholar 

  • Bezhanishvili, G., Esakia, L., & Gabelaia, D. (2010). The modal logic of stone spaces: Diamond as derivative. The Review of Symbolic Logic, 3(1), 26–40.

    Article  Google Scholar 

  • Button, T., & Walsh, S. (2018). Philosophy and Model Theory. Oxford University Press.

    Book  Google Scholar 

  • Callender, C., & Weingard, R. (1996). An introduction to topology. The Monist, 79(1), 21–33.

    Article  Google Scholar 

  • Carnap, R. (1920). Grundlegung Der Geometrie. MA Thesis. Box 3, Folder 12. UCLA Library Special Collections.

  • Carnap, R. (1922) 2019. Space: A contribution to the theory of science. In A.W. Carus, M. Friedman, W. Kienzler, A. Richardson, S. Schlotter (Eds.), The Collected Words of Rudolf Carnap, (vol. 1, pp. 21–208). Oxford University Press.

  • Carnap, R. (1925) 2019. On the dependence of the properties of space on those of time. In A.W. Carus, M. Friedman, W. Kienzler, A. Richardson, S. Schlotter (Eds.), The Collected Works of Rudolf Carnap, (vol. 1, pp. 297–325). Oxford University Press.

  • Carus, A.W., & Friedman M. (2019). Introduction. In A.W. Carus, M. Friedman, W. Kienzler, A. Richardson, & S. Schlotter (Eds.), The Collected Works of Rudolf Carnap, (vol. 1, pp. xxiii–xli). Oxford University Press.

  • Dieudonné, J. (2009). A History of Algebraic and Differential Topology, 1900–1960. Modern Birkhäuser Classics. Boston: Springer.

  • Dummett, M. (1978). Can analytical philosophy be systematic, and ought it to be? In Truth and Other Enigmas, (pp. 437–461). Cambridge, MA: Harvard University Press.

  • Elga, A. (2016). Bayesian humility. Philosophy of Science, 83(3), 305–323.

    Article  Google Scholar 

  • Esakia, L. (1974). Topological Kripke models. Doklady Akademii Nauk, 214(2), 298–301.

    Google Scholar 

  • Esakia, L. (2004). Intuitionistic logic and modality via topology. Annals of Pure and Applied Logic, 127(1–3), 155–170.

    Article  Google Scholar 

  • Feferman, S. (2004). Tarski’s conception of logic. Annals of Pure and Applied Logic, 126(1–3), 5–13.

    Article  Google Scholar 

  • Fletcher, S. C. (2020a). Similarity structure and emergent properties. Philosophy of Science, 87(2), 281–301.

    Article  Google Scholar 

  • Fletcher, S. C. (2020b). The principle of stability. Philosophers’ Imprint, 20(3), 1–22.

    Google Scholar 

  • Fletcher, S. C. (2021a). An invitation to approximate symmetry, with three applications to intertheoretic relations. Synthese, 198(5), 4811–4831.

    Article  Google Scholar 

  • Fletcher, S. C. (2021b). Similarity structure and diachronic emergence. Synthese, 198(9), 8873–8900.

    Article  Google Scholar 

  • Fletcher, S. C. (forthcoming). “Similarity Structure on Scientific Theories.” In B. Skowron (Ed.), Topological Philosophy. De Gruyter. http://philsci-archive.pitt.edu/17222/.

  • Forrest, P. (1996). From ontology to topology in the theory of regions. The Monist, 79(1), 34–50.

    Article  Google Scholar 

  • Forrest, P. (2010). Mereotopology without mereology. Journal of Philosophical Logic, 39(3), 229–254.

    Article  Google Scholar 

  • Franklin, P. (1922a). The four color problem. American Journal of Mathematics, 44(3), 225–236.

    Article  Google Scholar 

  • Franklin, P. (1922b). The meaning of rotation in the special theory of relativity. Proceedings of the National Academy of Sciences, 8(9), 265–268.

    Article  Google Scholar 

  • Franklin, P. (1934). A six color problem. Journal of Mathematics and Physics, 13(1–4), 363–369.

    Article  Google Scholar 

  • Franklin, P. (1935). What is topology? Philosophy of Science, 2(1), 39–47.

    Article  Google Scholar 

  • Franklin, P., & Wiener, N. (1926). Analytic approximations to topological transformations. Transactions of the American Mathematical Society, 28(4), 762–785.

    Article  Google Scholar 

  • Friedman, M. (2019). Editorial notes to “Space: A contribution to the theory of science”. In A.W. Carus, M. Friedman, W. Kienzler, A. Richardson, & S. Schlotter (Eds.), The Collected Works of Rudolf Carnap. (vol. 1, pp. 172–208). Oxford University Press.

  • Gabriel, G. (2007). Carnap and Frege. In M. Friedman & R. Creath (Eds.), The Cambridge Companion to Carnap (pp. 65–80). Cambridge University Press.

    Chapter  Google Scholar 

  • Gärdenfors, P. (2000). Conceptual Spaces: The Geometry of Thought. Cambridge, MA: MIT Press.

    Book  Google Scholar 

  • Gärdenfors, P. (2014). The Geometry of Meaning: Semantics Based on Conceptual Spaces. Cambridge, MA: MIT Press.

    Book  Google Scholar 

  • Genin, K., & Kelly, K. T. (2017). The topology of statistical verifiability. Electronic Proceedings in Theoretical Computer Science, 251(July), 236–250.

    Article  Google Scholar 

  • Giovanelli, M. (2021). Nothing but coincidences: the point-coincidence argument and einstein’s struggle with the meaning of coordinates in physics. European Journal for Philosophy of Science, 11(45), 1–64.

    Google Scholar 

  • Gold, E. M. (1965). Limiting recursion. Journal of Symbolic Logic, 30(1), 28–48.

    Article  Google Scholar 

  • Goldblatt, R. (2003). Mathematical modal logic: a view of its evolution. Journal of Applied Logic, 1(5–6), 309–392.

    Article  Google Scholar 

  • Grattan-Guinness, I. (2012). Logic, topology and physics: points of contact between Bertrand Russell and Max Newman. Russell: The Journal of Bertrand Russell Studies, 32(1), 5–29.

    Article  Google Scholar 

  • Grosholz, E. R. (1985). Two episodes in the unification of logic and topology. The British Journal for the Philosophy of Science, 36(2), 147–157.

    Article  Google Scholar 

  • Grosholz, E. R. (2007). Representation and Productive Ambiguity in Mathematics and the Sciences. Oxford University Press.

    Google Scholar 

  • Grünbaum, A. (1951). The philosophy of continuity: a philosophical interpretation of the metrical continuum of physical events in the light of contemporary mathematical conceptions. PhD Dissertation, New Haven, CT: Yale University.

  • Grünbaum, A. (1973). Philosophical Problems of Space and Time. 2nd ed. Dordrecht: D. Reidel.

  • Hausdorff, F. (1914). Grundzüge Der Mengenlehre. Leipzig: Veit.

    Google Scholar 

  • Hellman, G., & Shapiro, S. (2018). Varieties of Continua: From Regions to Points and Back. Oxford University Press.

    Book  Google Scholar 

  • Johanson, A. A. (1981). Topology without points. Quaestiones Mathematicae, 4, 185–200.

    Article  Google Scholar 

  • Johnstone, P. (1982). Stone Spaces. Cambridge Studies in Advanced Mathematics 3. Cambridge University Press.

  • Jonsson, B., & Tarski, A. (1951). Boolean algebras with operators. American Journal of Mathematics, 73(4), 891–939.

    Article  Google Scholar 

  • Kelly, K. T. (1996). The Logic of Reliable Inquiry. Logic and computation in philosophy. Oxford University Press.

  • Kugel, P. (1977). Induction, pure and simple. Information and Control, 35(4), 276–336.

    Article  Google Scholar 

  • Lemhoff, R. (2020). Intuitionism in the philosophy of mathematics. In E. N. Zalta (Ed.), The Stanford Encyclopedia of Philosophy, Fall 2020. https://plato.stanford.edu/archives/fall2020/entries/intuitionism/.

  • Lewin, K. (1923). Die Zeitliche Geneseordnung. Zeitschrift Für Physik, 13, 62–81.

    Article  Google Scholar 

  • Lewis, D. (1973). Counterfactuals. Blackwell.

    Google Scholar 

  • MacLane, S. (1939). Review of M.H. Stone’s ‘Applications of the theory of boolean rings to general topology.’ Journal of Symbolic Logic, 4(2), 88–89.

    Article  Google Scholar 

  • Malament, D. (2019). Editorial notes to ‘On the dependence of the properties of space on those of time.’ In A.W. Carus, M. Friedman, W. Kienzler, A. Richardson, S. Schlotter (Eds.), The Collected Works of Rudolf Carnap, (vol. 1, pp. 326–338). Oxford University Press.

  • Malaterre, C., Chartier, J.-F., & Pulizzotto, D. (2019). What is this thing called philosophy of science? A computational topic-modeling perspective, 1934–2015. HOPOS: The Journal of the International Society for the History of Philosophy of Science, 9(2), 215–249.

    Google Scholar 

  • McKinsey, J. C. C. (1941). A solution of the decision problem for the Lewis systems S2 and S4, with an application to topology. Journal of Symbolic Logic, 6(4), 117–124.

    Article  Google Scholar 

  • McKinsey, J. C. C., & Tarski, A. (1944). The algebra of topology. The Annals of Mathematics, 45(1), 141–191.

    Article  Google Scholar 

  • McKinsey, J. C. C., & Tarski, A. (1946). On closed elements in closure algebras. The Annals of Mathematics, 47(1), 122–162.

    Article  Google Scholar 

  • Mormann, T. (1993). Natural predicates and topological structures of conceptual spaces. Synthese, 95(2), 219–240.

    Article  Google Scholar 

  • Mormann, T. (2007). Geometrical Leitmotifs in Carnap’s early philosophy. In M. Friedman & R. Creath (Eds.), The Cambridge Companion to Carnap (pp. 43–64). Cambridge University Press.

    Chapter  Google Scholar 

  • Mormann, T. (2008). Russell’s many points. In A. Hieke & H. Leitgeb (Eds.), Proceedings of the 31th International Ludwig Wittgenstein-Symposium in Kirchberg, (pp. 239–258). Kirchberg: De Gruyter.

  • Mormann, T. (2013). Topology as an issue for history of philosophy of science. In H. Andersen, D. Dieks, W. J. Gonzalez, T. Uebel, & G. Wheeler, New Challenges to Philosophy of Science, (pp. 423–434). Philosophy of Science in a European Perspective 4. Dordrecht: Springer.

  • Mormann, T. (2020). Topological aspects of epistemology and metaphysics. In A. Peruzzi & S. Zipoli Caiani, Structures Mères: Semantics, Mathematics, and Cognitive Science, (pp. 135–152). Studies in Applied Philosophy, Epistemology and Rational Ethics 57. Cham: Springer.

  • Mormann, T. (2021). Prototypes, poles, and tessellations: Towards a topological theory of conceptual spaces. Synthese, 199(1–2), 3675–3710.

    Article  Google Scholar 

  • Newman, M. H. A. (1923). The foundations of mathematics from the standpoint of physics. Fellowship Dissertation. Item F33.1. St John’s College Archive, Cambridge.

  • Newman, M. H. A. (1939). The Elements of the Topology of Plane Sets of Points. 2nd ed. Cambridge: Cambridge University Press.

  • Norton, J. D. (2011). Challenges to Bayesian confirmation theory. In P. S. Bandyopadhyay & M. R. Forster (Eds.), Philosophy of Statistics (pp. 391–439). Handbook of the Philosophy of Science 7. Amsterdam: North-Holland.

  • O’Connor, J. J., & Robertson E. F. (2010). “Philip Franklin.” In MacTutor History of Mathematics Archive. Scotland: School of Mathematics and Statistics, University of St Andrews. https://mathshistory.st-andrews.ac.uk/Biographies/Franklin/.

  • O’Connor, J. J., & Robertson E. F. (2003). “Norbert Wiener.” In MacTutor History of Mathematics Archive. Scotland: School of Mathematics and Statistics, University of St Andrews. https://mathshistory.st-andrews.ac.uk/Biographies/Wiener_Norbert/.

  • Padovani, F. (2013). Genidentity and topology of time: Kurt Lewin and Hans Reichenbach. In N. Milkov & V. Peckhaus (Eds.), The Berlin Group and the Philosophy of Logical Empiricism, (pp. 91–122). Boston Studies in the Philosophy and History of Science 273. Dordrecht: Springer.

  • Poincaré, H. (1902) 2017. Science and Hypothesis: The Complete Text. London: Bloomsbury.

  • Pianesi, F., & Varzi, A. C. (1996). Events, topology and temporal relations. The Monist, 79(1), 89–116.

    Article  Google Scholar 

  • Poincaré, H. (1895). Analysis situs. Journal De L’école Polytechnique, 2(1), 1–123.

    Google Scholar 

  • Pratt-Hartmann, I. (2007). First-order mereotopology. In M. Aiello, I. Pratt-Hartmann, & J. Van Benthem (Eds.), Handbook of Spatial Logics, (pp. 13–97). Dordrecht: Springer Netherlands.

  • Putnam, H. (1965). Trial and error predicates and the solution to a problem of Mostowski. Journal of Symbolic Logic, 30(1), 49–57.

    Article  Google Scholar 

  • Reichenbach, H. (1924). Axiomatik der relativistischen Raum-Zeit-Lehre. Braunschweig: Fried. Vieweg & Sohn.

  • Reichenbach, H. (1965). The Axiomatization of the Theory of Relativity. Los Angeles: University of California Press.

    Google Scholar 

  • Riemann, B. (1864) 2016. On the Hypotheses Which Lie at the Bases of Geometry. In J. Jürgen (Ed.) Classic Texts in the Sciences (trans. William Clifford). Switzerland: Springer International Publishing.

  • Robb, A. A. (1914). A Theory of Time and Space. Cambridge University Press.

    Google Scholar 

  • Robb, A. A. (1921). The Absolute Relations of Time and Space. Cambridge University Press.

    Google Scholar 

  • Robb, A. A. (1936). The Geometry of Space and Time. Cambridge University Press.

    Google Scholar 

  • Roeper, P. (1997). Region-based topology. Journal of Philosophical Logic, 26(3), 251–309.

    Article  Google Scholar 

  • Russell, B. (1896). The logic of geometry. Mind, 5(17), 1–23.

    Article  Google Scholar 

  • Russell, B. (1897). An Essay on the Foundations of Geometry. Cambridge University Press.

    Google Scholar 

  • Russell, B. (1914). Our Knowledge of the External World. Chicago: Open Court Publishing Company.

    Google Scholar 

  • Russell, B. (1927). The Analysis of Matter. London: Kegan Paul.

    Google Scholar 

  • Ryckman, T. (2007). Carnap and Husserl. In M. Friedman & R. Creath (Eds.), The Cambridge Companion to Carnap (pp. 81–105). Cambridge University Press.

    Chapter  Google Scholar 

  • Ryckman, T. A. (2018). Early philosophical interpretations of general relativity. In E. N. Zalta (Ed.), Stanford Encyclopedia of Philosophy, Spring 2018. https://plato.stanford.edu/archives/spr2018/entries/genrel-early/.

  • Scott, D. (1968). Extending the topological interpretation to intuitionistic analysis. Compositio Mathematica, 20, 194–210.

  • Schulte, O. (1999a). Means-ends epistemology. The British Journal for the Philosophy of Science, 50(1), 1–31.

    Article  Google Scholar 

  • Schulte, O. (1999b). The logic of reliable and efficient inquiry. Journal of Philosophical Logic, 28(4), 399–438.

    Article  Google Scholar 

  • Schulte, O., & Juhl, C. (1996). Topology as epistemology. The Monist, 79(1), 141–147.

    Article  Google Scholar 

  • Shehtman, V. (1983). Modal logics of domains on the real plane. Studia Logica, 42, 63–80.

    Article  Google Scholar 

  • Smyth, M. B. (1983). Power domains and predicate transformers: A topological view. In Automata, Languages and Programming, (pp. 662–675). Lecture Notes in Computer Science 154. Berlin: Springer.

  • Starr, W. (2021). “Counterfactuals.” In E. N. Zalta (Ed.), Stanford Encyclopedia of Philosophy, Spring 2021. https://plato.stanford.edu/archives/spr2021/entries/counterfactuals/.

  • Steinsvold, C. (2008). A note on logics of ignorance and borders. Notre Dame Journal of Formal Logic, 49(4), 385–392.

    Article  Google Scholar 

  • Stone, M. H. (1934). Boolean algebras and their application to topology. Proceedings of the National Academy of Sciences, 20(3), 197–202.

    Article  Google Scholar 

  • Stone, M. H. (1936). The theory of representation for Boolean algebras. Transactions of the American Mathematical Society, 40(1), 37–111.

    Google Scholar 

  • Stone, M. H. (1937). Applications of the theory of Boolean rings to general topology. Transactions of the American Mathematical Society, 41(3), 375–481.

    Article  Google Scholar 

  • Stone, M. H. (1938). Topological representations of distributive lattices and Brouwerian logics. Časopis pro Pěstování Matematiky a Fysiky, 67(1), 1–25.

    Article  Google Scholar 

  • Struik, D. J. (1989). The MIT department of mathematics during its first seventy-five years: some recollections. In P. L. Duren, U. C. Merzbach, & H. Edwards (Eds.), A Century of Mathematics in America, (vol. 3, pp. 163–177). Providence: American Mathematical Society.

  • Suppes, P. (1998). Review: The logic of reliable inquiry, by Kevin Kelly. The British Journal for the Philosophy of Science, 49(2), 351–354.

    Article  Google Scholar 

  • Tarski, A. (1956). Sentential calculus and topology. In Logic, Semantics, Metamathematics: Papers from 1923 to 1938, translated by J.H. Woodger, pp. 421–454. Oxford: Clarendon Press.

  • van Benthem, J. (2006a). Epistemic logic and epistemology: The state of their affairs. Philosophical Studies, 128, 49–76.

    Article  Google Scholar 

  • Varzi, A. C. (1996). Parts, wholes, and whole-part relations: The prospects of mereotopology. Data & Knowledge Engineering, 20, 259–286.

    Article  Google Scholar 

  • Varzi, A. C. (2007). Spatial reasoning and ontology: parts, wholes, and locations. In M. Aiello, I. Pratt-Hartmann, & J. Van Benthem (Eds.), Handbook of Spatial Logics, (pp. 945–1038). Dordrecht: Springer Netherlands.

  • Veblen, O. (1922). Analysis Situs. American Mathematical Society Colloquium Lectures. New York: American Mathematical Society.

  • Veblen, O., & Whitehead, J. H. C. (1931). A set of axioms for differential geometry. Proceedings of the National Academy of Sciences, 17(10), 551–561.

    Article  Google Scholar 

  • Veblen, O., & Whitehead, J. H. C. (1932). The Foundations of Differential Geometry. Cambridge University Press.

    Google Scholar 

  • Vickers, Steven. 1989. Topology via Logic. Cambridge Tracts in Theoretical Computer Science. Cambridge University Press.

  • Vietoris, L. (1921). Stetige Mengen. Monatshefte für Mathematik und Physik, 31, 173–204.

    Article  Google Scholar 

  • Wahl, R. (Ed.). (2018). The Bloomsbury Companion to Bertrand Russell. London: Bloomsbury.

  • Wallman, H. (1938). Lattices and topological spaces. Annals of Mathematics, 39(1), 112–126.

    Article  Google Scholar 

  • Whitehead, A. N. (1906). On mathematical concepts of the material world. Philosophical Transactions of the Royal Society of London, 205, 465–525.

    Article  Google Scholar 

  • Whitehead, A. N. (1929). Process and Reality. Macmillian.

    Google Scholar 

  • Winnie, J. A. (1977). The causal theory of space-time. In J. S. Earman, C. N. Glymour, & J. J. Stachel (Eds.), Foundations of Space-Time Theories, (pp. 134–205). Minnesota Studies in the Philosophy of Science 8. Minneapolis: University of Minnesota Press.

  • van Benthem, J. (2006b). Open problems in logical dynamics. In D. M. Gabbay, S. S. Goncharov, & M. Zakharyaschev (Eds.), Mathematical Problems from Applied Logic I: Logics for the XXIst Century, (pp. 136–192). International Mathematical Series 4. New York: Springer.

  • van Benthem, J. (2011). Logical Dynamics of Information. Cambridge: Cambridge University Press.

  • van Benthem, J., Bezhanishvili, G. (2007). Modal logics of space. In M. Aiello, I. Pratt-Hartmann, & J. Van Benthem (Eds.), Handbook of Spatial Logics, (pp. 217–298). Dordrecht: Springer Netherlands.

  • van Benthem, J., & Martinez, M. (2008). The stories of logic and information. In P. Adriaans & J. Benthem (Eds.), Philosophy of Information (pp. 217–80). Handbook of the Philosophy of Science 8. Amsterdam: North-Holland.

Download references

Funding

The authors acknowledge support from the Office of the Vice President for Research, University of Minnesota, Twin Cities.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Samuel C. Fletcher.

Ethics declarations

Conflict of interest

The author declares that they have no conflict of interest.

Additional information

Publisher's Note

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

This article belongs to the topical collection “Metaphilosophy of Formal Methods”, edited by Samuel C. Fletcher.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Fletcher, S.C., Lackey, N. The introduction of topology into analytic philosophy: two movements and a coda. Synthese 200, 197 (2022). https://doi.org/10.1007/s11229-022-03689-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s11229-022-03689-9

Keywords

Navigation