Hostname: page-component-8448b6f56d-xtgtn Total loading time: 0 Render date: 2024-04-19T21:06:11.642Z Has data issue: false hasContentIssue false

THE LOGIC AND TOPOLOGY OF KANT’S TEMPORAL CONTINUUM

Published online by Cambridge University Press:  29 January 2018

RICCARDO PINOSIO*
Affiliation:
ILLC and Department of Philosophy, University of Amsterdam
MICHIEL VAN LAMBALGEN*
Affiliation:
ILLC and Department of Philosophy, University of Amsterdam
*
*ILLC AND DEPARTMENT OF PHILOSOPHY UNIVERSITY OF AMSTERDAM P.O. BOX 94242, 1090 GE AMSTERDAM, THE NETHERLANDS E-mail: rpinosio@gmail.com
ILLC AND DEPARTMENT OF PHILOSOPHY UNIVERSITY OF AMSTERDAM P.O. BOX 94242, 1090 GE AMSTERDAM, THE NETHERLANDS E-mail: m.vanlambalgen@uva.nl

Abstract

In this paper we provide a mathematical model of Kant’s temporal continuum that yields formal correlates for Kant’s informal treatment of this concept in the Critique of Pure Reason and in other works of his critical period. We show that the formal model satisfies Kant’s synthetic a priori principles for time (whose consistence is not obvious) and that it even illuminates what “faculties and functions” must be in place, as “conditions for the possibility of experience”, for time to satisfy such principles. We then present a mathematically precise account of Kant’s transcendental theory of time—the most precise account to date.

Moreover, we show that the Kantian continuum which we obtain has some affinities with the Brouwerian continuum but that it also has “infinitesimal intervals” consisting of nilpotent infinitesimals; these allow us to capture Kant’s theory of rest and motion in the Metaphysical Foundations of Natural Science.

While our focus is on Kant’s theory of time the material in this paper is more generally relevant for the problem of developing a rigorous theory of the phenomenological continuum, in the tradition of Whitehead, Russell, and Weyl among others.

Type
Research Article
Copyright
Copyright © Association for Symbolic Logic 2018 

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

BIBLIOGRAPHY

Achourioti, T. & van Lambalgen, M. (2011). A formalization of Kant’s transcendental logic. The Review of Symbolic Logic, 4(02), 254289.Google Scholar
Alexandroff, P. (1937). Diskrete Räume. Matematicheskii Sbornik, 2(3), 501519.Google Scholar
Allen, J. F. (1983). Maintaining knowledge about temporal intervals. Communications of the ACM, 26(11), 832843.CrossRefGoogle Scholar
Arthur, R. T. (1995). Newton’s fluxions and equably flowing time. Studies in History and Philosophy of Science Part A, 26(2), 323351.Google Scholar
Banaschewski, B. & Pultr, A. (2010). Approximate maps, filter monad, and a representation of localic maps. Archivum Mathematicum, 46(4), 285298.Google Scholar
Barrow, I. (1976). Lectiones Geometricae. Hildesheim: Georg Olms Verlag.Google Scholar
Carelli, M. G. & Forman, H. (2012). Representation of multiple durations in children and adults. Child Development Research, 2011, 19.Google Scholar
Casasanto, D. & Boroditsky, L. (2008). Time in the mind: Using space to think about time. Cognition, 106(2), 579593.Google Scholar
Casasanto, D., Fotakopoulou, O., & Boroditsky, L. (2010). Space and time in the child’s mind: Evidence for a cross-dimensional asymmetry. Cognitive Science, 34(3), 387405.Google Scholar
Coquand, T. (2002). A Completness Proof for Geometric Logic. Gothenburg: Computer Science and Engineering Department, University of Gothenburg.Google Scholar
Dehaene, S. & Brannon, E. (2011). Space, Time and Number in the Brain: Searching for the Foundations of Mathematical Thought. Cambridge, MA: Academic Press.Google Scholar
Friedman, M. (2012). Kant on geometry and spatial intuition. Synthese, 186(1), 231255.Google Scholar
Friedman, M. (2013). Kant’s Construction of Nature: A Reading of the Metaphysical Foundations of Natural Science. Cambridge: Cambridge University Press.Google Scholar
Hellman, G. & Shapiro, S. (2013). The classical continuum without points. The Review of Symbolic Logic, 6(03), 488512.Google Scholar
Hodges, W. (1997). A Shorter Model Theory. Cambridge: Cambridge University Press.Google Scholar
Johnstone, P. T. (1986). Stone Spaces. Cambridge: Cambridge University Press.Google Scholar
Kant, I., Holger, K., Gerresheim, E., Heidemann, I., & Martin, G. (1908). Gesammelte Schriften. Berlin: G. Reimer.Google Scholar
Khalimsky, E., Kopperman, R., & Meyer, P. R. (1990). Computer graphics and connected topologies on finite ordered sets. Topology and its Applications, 36(1), 117.Google Scholar
Kock, A. (2006). Synthetic Differential Geometry. Cambridge: Cambridge University Press.Google Scholar
Kopperman, R., Kronheimer, E., & Wilson, R. (1998). Topologies on totally ordered sets. Topology and its Applications, 90(1), 165185.Google Scholar
Longuenesse, B. (1998). Kant and the Capacity to Judge. Princeton, NJ: Princeton University Press.Google Scholar
Nicod, J. (2014). Foundations of Geometry and Induction. Abingdon: Routledge.CrossRefGoogle Scholar
Northoff, G. (2012). Immanuel Kant’s mind and the brain’s resting state. Trends in Cognitive Sciences, 16(7), 356359.Google Scholar
Onof, C. & Schulting, D. (2015). Space as form of intuition and as formal intuition: On the note to B160 in Kant’s Critique of Pure Reason. Philosophical Review, 124(1), 158.Google Scholar
Ostaszewski, A. J. (1974). A characterization of compact, separable, ordered spaces. Journal of the London Mathematical Society, 2(4), 758760.CrossRefGoogle Scholar
Palmer, L. (2008). Kant and the brain: A new empirical hypothesis. Review of General Psychology, 12(2), 105.CrossRefGoogle Scholar
Palmer, L. & Lynch, G. (2010). A Kantian view of space. Science, 328(5985), 14871488.CrossRefGoogle ScholarPubMed
Pinosio, R. (2017). The Logic of Kant’s Temporal Continuum. Doctoral Dissertation, Institute for Logic, Language and Computation, University of Amsterdam.Google Scholar
Roeper, P. (2006). The Aristotelian continuum. A formal characterization. Notre Dame Journal of Formal Logic, 47(2), 211232.Google Scholar
Rota, G.-C. (1991). The pernicious influence of mathematics upon philosophy. Synthese, 88(2), 165178.Google Scholar
Russell, B. (1936). On order in time. Proceedings of the Cambridge Philosophical Society, 32, 216228.Google Scholar
Sambin, G. (2003). Some points in formal topology. Theoretical Computer Science, 305(1), 347408.Google Scholar
Tarski, A. (1959). What is elementary geometry? In Henkin, L., Suppes, P., and Tarski, A., editors. The Axiomatic Method with Special Reference to Geometry and Physics. Amsterdam: Elsevier, pp. 1629.Google Scholar
Thomason, S. (1984). On constructing instants from events. Journal of Philosophical Logic, 13(1), 8596.Google Scholar
Thomason, S. (1989). Free construction of time from events. Journal of Philosophical Logic, 18(1), 4367.Google Scholar
van Benthem, J. (2013). The Logic of Time: A Model-Theoretic Investigation into the Varieties of Temporal Ontology and Temporal Discourse. Dordrecht: Springer Science & Business Media.Google Scholar
van Dalen, D. (2009). The return of the flowing continuum. Intellectica, 1, 51.Google Scholar
Walker, A. G. (1947). Durées et instants. Revue Scientifique, 85, 131134.Google Scholar