Skip to main content

The ‘Space’ at the Intersection of Physics, Metaphysics, and Mathematics

  • Chapter
  • First Online:
Book cover The Deep Metaphysics of Space

Part of the book series: European Studies in Philosophy of Science ((ESPS))

  • 423 Accesses

Abstract

As developed in Chap. 5, the attempts to fcapture the ontology of space and spacetime using the conceptual apparatus of substantivalism and relationism—that space is, respectively, either an independently existing entity or a relation among material substances—have proved to be quite problematic since the sophisticated versions of both substantivalism and relationism seem to be identical in the context of general relativity (GR), our best theory on the large scale structure of space. But, whether space is a substance or a relation is just one of the many conceptual distinctions that have entered the modern debate on the nature of space. In addition to substantivalism and relationism, one also finds references to realism versus anti-realism, platonism versus nominalism, and background independence versus background dependence. In this chapter, we will begin the examination of these new dichotomies, their impact on the traditional substantival/relational distinction, and their relationship to one of the least discussed, but quite central, components of the spatial ontology debate, namely, how the philosophy of mathematics factors into the philosophy of space and time.

This is a preview of subscription content, log in via an institution to check access.

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 69.99
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 89.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD 119.99
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

Notes

  1. 1.

    See, Chakravartty 2011, who raises problems for using the traditional mind-based, or linguistic-conventional, type of nominalism in conjunction with scientific realism (SR): i.e., if a body’s properties are mind-dependent, then scientific realism is undermined (since SR claims that a mind-independent reality explains the success of our best scientific theories). Chakravartty’s seems to advocate a version of a trope theory to overcome this difficulty and uphold SR; i.e., he claims that a thing’s properties can allow different mind-based interpretations without undermining the reality of the property (2011, 170–171). However, the traditional nominalist could simply accept that certain interpretations of an object’s properties are true, and others false (which thereby upholds SR as well), but without reifying the object’s properties as real particular things (a la trope theory). In fact, Chakravartty’s view may constitute a trope-based version of the type of (neo-Meinongian) nominalism advanced in this chapter. See, also, footnotes 2 and 3.

  2. 2.

    Immanent realism differs from a trope theory in that the former accepts abstract objects (so that, e.g., all square objects instantiate “squareness”, the latter being the very same abstract object in all square bodies), whereas trope theorists reject abstract objects (hence, all square bodies have a property of squareness, the latter being a particular that stands in relations of “resemblance” with other square particulars in other square material objects). In addition, Sepkoski (2007, 16) suggests a three part division of mathematical ontology that is close to the three part division offered in our investigation, although he equates the view that we have labeled “immanent realism” with Aristotle’s position as regards the existence of substantial forms.

  3. 3.

    We have employed the term “truth-based nominalism”, which is adapted from Pincock’s (2012) concept, “truth-value nominalism”, so that Pincock’s own usage is kept separate from our interpretation. Alternatively, truth-based nominalism can be called “liberal nominalism”, a designation used in Slowik (2005a). In Balaguer (2009), the view that we have defined as truth-based nominalism is categorized as “neo-Meinongianism”, and he includes a number of the authors discussed in this chapter as among its defenders, e.g., Salmon 1998, Azzouni 2004, Priest 2005, and Bueno 2005.

  4. 4.

    However, Newstead and Franklin’s view seems to be a mixture of our truth-based nominalism and immanent realism (2012, 83–84). Overall, there are many options for constructing an hypothesis that lies between the extremes of fictionalism and traditional platonism (where abstract objects are non-spatial in Plato’s sense). Examining all of these potential views, and how they might differ from one another, is beyond the scope of this chapter.

  5. 5.

    In particular, Pincock (in Balaguer et al. 2013, 271–272) has doubts concerning the hypothesis entitled, “heavy duty platonism” (see below), although our analysis treats this concept as equivalent to truth-based nominalism. The problem probably resides in the difficulty in determining just what Field intends by this concept, as will be further explained below.

  6. 6.

    The view of structure advocated in Brading and Landry (2006), called “minimal structuralism”, would seem to be in accord with both truth-based nominalism and the ESR view advocated in this monograph (and the ESR-L version that will be developed in Chap. 8). “What we call minimal structuralism is committed only to the claim that the kinds of objects that a theory talks about are presented through the shared structure of its theoretical models and that the theory applies to the phenomena just in case the theoretical models and the data models share the same kind of structure. No ontological commitment—nothing about the nature, individuality or modality of particular objects—is entailed” (2006, 577). If one interprets “objects” as including substantival space, bodies, fields, etc., then this characterization of structure could apply to the conception of spacetime structure advocated in our investigation.

  7. 7.

    Another question concerns what counts as a spacetime theory. If any use of geometric relations were to qualify as the type of structuralist spacetime theory considered above, then nearly all physical theories would trivially count given an actual or potential geometric formulation. A case in point is Fresnel’s wave optics, cited in Chap. 5, which utilizes a trigonometric relationship to relate the intensities of reflected and refracted light that pass through mediums of different optical density (Worrall 1989, 119). One could classify such theories as spacetime theories, of course, but the designation seems more apt for the most basic dynamical theories that correlate inertial motion and force against a particular spacetime backdrop (e.g., Newtonian, Leibnizian, etc.).

  8. 8.

    The Quine-Putnam Indispensability thesis is the best known philosophical debate related to the interface of mathematics and empirical reality, although the thesis does not really concern the interaction between the two realms; rather, the contention is that one should be a realist about mathematics if one is also a realist about scientific laws, objects, etc.

  9. 9.

    It should be noted that the problem of incongruent counterparts (Kant’s handed objects) could stand as a counter-argument against the claim that spacetime lacks causal powers. On the other hand, one could always invoke a maneuver similar to that advocated in Sklar (1985, 234–248), which accepts an “intrinsic” property of handedness (i.e., a primitive property defined via continuous rigid motions) to avoid the commitment to substantival space. On the whole, there are a variety of strategies that the relationist can offer to undermine the inference that space “causes” the change in handedness. That is, while the background spatial structure is obviously relevant to determining an object’s handedness, the claim that space causes a change in handedness is as dubious as the claim that inertial structure is the cause of the observed non-inertial forces of a (non-uniformly) accelerating body.

  10. 10.

    See, Teller (1991), Sklar (1990), Bricker (1990), and Azzouni (2004, 196–212), for similar arguments about the causal irrelevance of spacetime structures for explaining accelerated motions and effects. An important early critique is Einstein (1923, 112–113), who labels absolute space as “a fictitious cause” since rotation with respect to absolute space is not “an observable fact of experience”. Furthermore, with respect to the interrelationship between the metric and matter fields in GR, this interrelationship does not explain why non-inertial forces are associated with accelerated motion; rather, the interrelationship is relevant to explaining the metric curvature.

  11. 11.

    Since truth-based nominalists reject platonism, they do not claim, of course, that mathematical entities and their relationship enter into spacetime theories. But, spacetime structures are not just the physical relationships between the physical objects, either, since (as argued above) causation is not a spacetime relationship (whereas causation is probably the most prevalent physical relationship among physical objects). Spacetime structures, like all other mathematical structures, are unique in this regard: systems can exemplify the structures, but the structures are neither identical to the systems, nor (for the nominalists) can they exists in the absence of systems.

  12. 12.

    Some of Newton’s descriptions might challenge this conclusion, such as his contention that God “constitutes” space (N 91), if one takes that term as synonymous with “is”. Likewise, one may strive to read such terms as “effects”, “affection”, “attribute”, “modes”, and “consequences”, as Newton’s somewhat clumsy way of describing a fictionalist elimination of space in favor of God’s being. Yet, Clarke’s straightforward denial that “space is consequence of God” equates with “space is God” poses a significant challenge, as does the tenor of Newton’s Des Maizeaux drafts, which rejects Clarke’s use of the term “property” in favor of “modes” and “consequences”; likewise, he does not object to Clarke’s claim that “space is not God”. In short, a fictionalist interpretation faces severe obstacles given this evidence.

  13. 13.

    One of the most straightforward declarations of nominalism can be found in the New Essays, where numbers and extension are compared (see also Chap. 3): “[I]n conceiving several things at once one conceives something in addition to the number, namely the things numbered; and yet there are not two pluralities, one of them abstract (for the number) and the other concrete (for the things numbered). In the same way, there is no need to postulate two extensions, one abstract (for space) and the other concrete (for body)” (NE II.iv.5).

  14. 14.

    In more detail, although a monad’s extended secondary matter would not arise if God prevented it, with secondary matter equating with bodily extension at the macrolevel, the unextended primary matter would remain at the microlevel. See Chap. 4 for more on the primary/secondary, and primitive/derivative force, distinctions.

  15. 15.

    While the origins of Leibniz’ view remain uncertain, it is highly likely that this particular dispute between Leibniz and Clarke stems, at least in part, from a more basic dichotomy concerning the definition of ratios in (Euclid’s) geometry, a dichotomy that found many important advocates in the seventeenth century. One school of thought, represented by Wallis, accepts that ratios are quantities that can be explained purely algebraically using an intricate exponent method; whereas others, such as Barrow, argue that ratios are not quantities, hence algebraic relationships fail in this regard (i.e., a geometric conception of quantity is primary). What is intriguing is that, like Wallis, Leibniz defends the primacy of algebra over geometry (e.g., AG 251–252), while Clarke and Newton side with Barrow in defending the primacy of geometry over algebra. In short, the foundations debate in the philosophy of mathematics, i.e., whether algebra or geometry is more fundamental, or at least superior, may be the origins of this particular aspect of the Leibniz-Clarke correspondence—and, in turn, it may ultimately reflect a more basic difference in their respective visions of the deep metaphysics of spatial ontology: Newton and Clarke’s spatially extended God conception versus Leibniz’ non-spatial God/monad hypothesis. More historical work is required on the origin of Leibniz’ view, but see Jesseph (2016) for an analysis of the ratio dispute between Wallis and Barrow.

  16. 16.

    In the L.III.5 quote above, the contrast between “space as absolute being/reality” and “space as a real truth” could be taken to correspond to, respectively, a commitment to abstract geometric objects versus the truths of geometry. This inference gains support given Leibniz’ rejection of abstract objects in general (see footnote 13). Yet, leaving aside abstract objects, Newton’s conception of space is actually similar to Leibniz’ theory: both accept nominalism, with the main difference being the ontological foundation of that nominalist spatial structure, God (Newton/Clarke) versus matter and God (Leibniz). Since God is central to the ontology of space for both Newton and Leibniz, the L.III.5 quote can also be viewed as the difference between positing a conception of space that is either entirely independent of matter (Newton/Clarke) or dependent on matter for instantiating the God-grounded truths of spatial geometry (Leibniz).

  17. 17.

    The analogy developed here is a bit loose, as a commentator has pointed out in an earlier version of this material. Since a wavefunction is an element of a Hilbert space, and is thus not defined on a Hilbert space in the same way that a metric is defined on a manifold, a more correct analogy would be between a wavefunction as an element of a Hilbert space and a manifold point as an element of a manifold. However, substituting an analogy between the wavefunction/Hilbert space and a point/manifold, rather than employing an analogy between the wavefunction/Hilbert space and a metric/manifold, works just as well for our purposes. On a related note, Dieks correctly points out that, as regard non-relativistic QM, “the Hilbert space formalism does not start from a space-time manifold in which particles are located. The quantum state is given by a vector in Hilbert space, and has in general no special relation to specific space-time points. Rather, ‘position’ is treated in the same way as ‘spin’ or other quantities that are direct particle properties: all the quantities are ‘observables’, represented by Hermetian operators in Hilbert space” (Dieks 2001a, 16).

References

  • Adams, R. 1994. Leibniz: Determinist, theist, idealist. Oxford: Oxford University Press.

    Google Scholar 

  • Armstrong, D.M. 1988. Can a naturalist believe in universals. In Science in reflection, ed. E. Ullmann-Margalit, 103–115. Dordrecht: Kluwer.

    Google Scholar 

  • Armstrong, D.M. 1997. A world of states of affairs. Cambridge: Cambridge University Press.

    Book  Google Scholar 

  • Arntzenius, F. 2012. Space, time, and stuff. Oxford: Oxford University Press.

    Book  Google Scholar 

  • Arthur, R. 2013. Leibniz’s theory of space. Foundations of Science 18: 449–528.

    Article  Google Scholar 

  • Azzouni, J. 2004. Deflating existential consequence. Oxford: Oxford University Press.

    Book  Google Scholar 

  • Balaguer, M. 2009. Platonism in metaphysics. In The Stanford encyclopedia of philosophy, ed. E. Zalta. http://plato.stanford.edu/archives/sum2009/entries/platonism/.

  • Balaguer, M., E. Landry, S. Bangu, and C. Pincock. 2013. Structures, fictions, and the explanatory epistemology of mathematics in science. Metascience 22: 247–273.

    Article  Google Scholar 

  • Barbour, J. 1982. Relational concepts of space and time. British Journal for the Philosophy of Science 33: 251–274.

    Article  Google Scholar 

  • Bird, A. 2007. Nature’s metaphysics. Oxford: Oxford University Press.

    Book  Google Scholar 

  • Brading, K., and A. Skiles. 2012. Underdetermination as a Path to Structural Realism. In Structural realism: Structure, object, and causality, ed. E. Landry and D. Rickles, 99–116. Dordrecht: Springer.

    Google Scholar 

  • Brading, K., and E. Landry. 2006. A minimal construal of scientific structuralism. Philosophy of Science 73: 571–581.

    Article  Google Scholar 

  • Bricker, P. 1990. Absolute time versus absolute motion: Comments on Lawrence Sklar. In Philosophical perspectives on Newtonian science, ed. P. Bricker and R.I.G. Hughes, 77–91. Cambridge, MA: MIT Press.

    Google Scholar 

  • Brown, H. 2005. Physical relativity. Oxford: Oxford University Press.

    Book  Google Scholar 

  • Bueno, O. 2005. Dirac and the dispensability of mathematics. Studies in History and Philosophy of Modern Physics 36: 465–490.

    Article  Google Scholar 

  • Chakravartty, A. 2011. Scientific realism and ontological relativity. The Monist 94: 157–180.

    Article  Google Scholar 

  • Chihara, C. 2004. A structural account of mathematics. Oxford: Oxford University Press.

    Google Scholar 

  • Colyvan, M. 2001. The indispensability of mathematics. Oxford: Oxford University Press.

    Book  Google Scholar 

  • Dieks, D. 2001a. Space-time relationism in Newtonian and relativistic physics. International Studies in the Philosophy of Science 15: 5–17.

    Article  Google Scholar 

  • Dieks, D. 2001b. Space and time in particle and field physics. Studies in History and Philosophy of Modern Physics 32: 217–241.

    Article  Google Scholar 

  • DiSalle, R. 1994. On dynamics, indiscernibility, and spacetime ontology. British Journal for the Philosophy of Science 45: 265–287.

    Article  Google Scholar 

  • DiSalle, R. 1995. Spacetime theory as physical geometry. Erkenntnis 42: 317–337.

    Article  Google Scholar 

  • Earman, J. 1989. World enough and space-time. Cambridge: MIT Press.

    Google Scholar 

  • Earman, J., and J. Norton. 1987. What price space-time substantivalism?: The hole story. British Journal for the Philosophy of Science 38: 515–525.

    Article  Google Scholar 

  • Einstein, A. 1923. The foundations of the general theory of relativity. In The Principle of Relativity. Trans. W. Perrett and G.B. Jeffery, 109–164. New York: Dover Publications.

    Google Scholar 

  • Field, H. 1980. Science without numbers. Princeton: Princeton University Press.

    Google Scholar 

  • Friedman, M. 1981. Theoretical explanation. In Reduction, time and reality, ed. R. Healey. Cambridge: University of Cambridge Press.

    Google Scholar 

  • Hellman, G. 1989. Mathematics without numbers. Oxford: Oxford University Press.

    Google Scholar 

  • Huggett, N. 2006b. The regularity account of relational spacetime, Mind 114: 41–73.

    Google Scholar 

  • Jesseph, D. 2016. Ratios, quotients, and the language of nature. In The language of nature: Reassessing the mathematization of natural philosophy, Minnesota studies in the philosophy of science, 160-177. Minneapolis: University of Minnesota Press.

    Google Scholar 

  • Koyré, A., and I.B. Cohen. 1962. Newton and the Leibniz-Clarke correspondence. Archives Internationales d’Histoire des Sciences 15: 63–126.

    Google Scholar 

  • Lewis, D. 1993. Mathematics is megethology. Philosophia Mathematica 1: 3–23.

    Article  Google Scholar 

  • Liston, M. 2000. Review of M. Steiner. The Applicability of Mathematics as a Philosophical Problem Philosophia Mathematica 8: 190–207.

    Article  Google Scholar 

  • Loux, M. 2006. Metaphysics, 3rd ed. Routledge: New York.

    Google Scholar 

  • Lowe, E.J. 2002. A survey of metaphysics. Oxford: Oxford University Press.

    Google Scholar 

  • Magalhães, E. 2006. Armstrong on the spatio-temporality of universals. Australasian Journal of Philosophy 84: 301–308.

    Article  Google Scholar 

  • Mundy, B. 1983. Relational theories of Euclidean space and Minkowski spacetime. Philosophy of Science 50: 205–226.

    Article  Google Scholar 

  • Newstead, A. and J. Franklin. 2012. Indispensability without Platonism. In Properties, powers and structures, ed. A. Bird, B. Ellis, and H. Sankey, 81–97. New York: Routledge.

    Google Scholar 

  • Ney, A., and D. Albert. 2013. The wave function: Essays on the metaphysics of quantum mechanics. Oxford: Oxford University Press.

    Book  Google Scholar 

  • Pasnau, R. 2011. Metaphysical themes: 1274–1671. Oxford: Oxford University Press.

    Google Scholar 

  • Pincock, C. 2012. Mathematics and scientific representation. Oxford: Oxford University Press.

    Book  Google Scholar 

  • Pooley, O. 2006. Points, particles, and structural realism. In The structural foundations of quantum gravity, ed. D. Rickles, S. French, and J. Saatsi, 83–120. Oxford: Oxford University Press.

    Google Scholar 

  • Priest, G. 2005. Towards non-being. Oxford: Oxford University Press.

    Book  Google Scholar 

  • Psillos, S. 2010. Scientific realism: Between platonism and nominalism. Philosophy of Science 77: 947–958.

    Article  Google Scholar 

  • Psillos, S. 2011. Living with the abstract: Realism and models. Synthese 180: 3–17.

    Article  Google Scholar 

  • Psillos, S. 2012. Anti-nominalistic scientific realism: A defense. In Properties, powers and structures, ed. A. Bird, B. Ellis, and H. Sankey, 63–80. New York: Routledge.

    Google Scholar 

  • Resnik, M. 1985. How nominalist is Hartry Field’s nominalism. Philosophical Studies 47: 163–181.

    Article  Google Scholar 

  • Salmon, N. 1998. Nonexistence. Noûs 32: 277–319.

    Article  Google Scholar 

  • Sepkoski, D. 2007. Nominalism and constructivism in seventeenth-century mathematical philosophy. New York: Routledge.

    Google Scholar 

  • Shapiro, S. 1997. Philosophy of mathematics: Structure and ontology. Oxford: Oxford University Press.

    Google Scholar 

  • Shapiro, S. 2000. Thinking about mathematics. Oxford: Oxford University Press.

    Google Scholar 

  • Sklar, L. 1974. Space, time, and spacetime. Berkeley: University of California Press.

    Google Scholar 

  • Sklar, L. 1985. Philosophy and spacetime physics. Berkeley: University of California Press.

    Google Scholar 

  • Sklar, L. 1990. Real Quantities and Their Sensible Measures. In Philosophical perspectives on Newtonian science, ed. P. Bricker and R.I.G. Hughes, 57–76. Cambridge, MA: MIT Press.

    Google Scholar 

  • Slowik, E. 2005a. Spacetime, ontology, and structural realism. International Studies in the Philosophy of Science 19: 147–166.

    Article  Google Scholar 

  • Stein, H. 1977b. On space-time and ontology: Extract from a letter to Adolf Grünbaum. In Foundations of space-time theories, Minnesota studies in the philosophy of science, vol. 8, ed. J. Earman, C. Glymour, and J. Stachel, 374–403. Minneapolis: University of Minnesota Press.

    Google Scholar 

  • Sylla, E. 2002. Space and spirit in the transition from Aristotelian to Newtonian space. In The dynamics of Aristotelian natural philosophy from antiquity to the seventeenth century, ed. C. Leijenhorst, C. Lüthy, and J. Thijssen, 249–287. Leiden: Brill.

    Google Scholar 

  • Teller, P. 1991. Substance, relations, and arguments about the nature of space-time. Philosophical Review 100: 362–397.

    Article  Google Scholar 

  • Wallace, D., and C. Timpson. 2010. Quantum mechanics on spacetime I: Spacetime state realism. British Journal for the Philosophy of Science 61: 697–727.

    Article  Google Scholar 

  • Worrall, J. 1989. Structural realism: The best of both worlds? Dialectica 43: 99–124.

    Article  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Rights and permissions

Reprints and permissions

Copyright information

© 2016 Springer International Publishing Switzerland

About this chapter

Cite this chapter

Slowik, E. (2016). The ‘Space’ at the Intersection of Physics, Metaphysics, and Mathematics. In: The Deep Metaphysics of Space. European Studies in Philosophy of Science. Springer, Cham. https://doi.org/10.1007/978-3-319-44868-8_7

Download citation

Publish with us

Policies and ethics