Skip to main content
Log in

On the Boundary of the Cosmos

  • Published:
Foundations of Physics Aims and scope Submit manuscript

Abstract

Intuitively, the totality of physical reality—the Cosmos—has a beginning only if (i) all parts of the Cosmos agree on the direction of time (the Direction Condition) and (ii) there is a boundary to the past of all non-initial spacetime points such that there are no spacetime points to the past of the boundary (the Boundary Condition). Following a distinction previously introduced by J. Brian Pitts, the Boundary Condition can be conceived of in two distinct ways: either topologically, i.e., in terms of a closed boundary, or metrically, i.e., in terms of the Cosmos having a finite past. This article proposes that the Boundary Condition should be posed disjunctively, modifies and improves upon the metrical conception of the Cosmos’s beginning in light of a series of surprising yet simple thought experiments, and suggests that the Direction and Boundary Conditions should be thought of as more fundamental to the concept of the Cosmos’s beginning than classical Big Bang cosmology.

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

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Similar content being viewed by others

Notes

  1. I will often speak as though spatio-temporal regions are point sets, but friends of the Aristotelian conception of continua can interpret at least some of my point set talk as involving a kind of pragmatic fiction. Moreover, the Aristotelian account of continua lacks the resources to distinguish between closed and open sets and, for that reason, do not have the resources to develop the Topological Conception, since the Topological Conception requires the notion of a point set with a closed boundary. However, the Boundary Condition is defined disjunctively and the second disjunct – the Metrical Conception – is the only natural way for Aristotelians to develop a conception of the Cosmos’s beginning. For that reason, the Boundary Condition does not assume the truth of either the Aristotelian or Cantorian accounts.

  2. Although I will describe spacetime in terms of the pair (Mg), I should not be interpreted as siding with what has sometimes been called the “angle bracket school” [14] or indeed any other approach to the foundations of spacetime theories; instead, I only intend that spacetime can be mathematically represented by the pair (Mg).

  3. For defenses of metrical conventionalism, see [13, 15,16,17].

  4. My comments depend upon two facts: first, that there is no topological feature that distinguishes the intervals (0, 1] and \((-\infty , 1]\). Second, that there is a topological feature that distinguishes (0, 1] and [0, 1]. What does it mean to say that there is no topological feature that distinguishes two intervals? X and Y are said to be topologically equivalent just in case there exists a continuous transformation from X to Y and which has a continuous inverse. In order to prove that (0, 1] and \((-\infty , 1]\) are topologically equivalent, it suffices to construct a suitable continuous transformation and to show that the inverse of the transformation is continuous. I will assume that the topology on \((-\infty , 1]\) and (0, 1] is the subspace topology, that is, the topology inherited from the standard topology on \({\mathbb {R}}\). Consider the function \(f(x) = -1/x + 2\). Trivially, f(x) is a monotonically increasing function that maps the interval (0, 1] to \((-\infty , 1]\); moreover, f(x) trivially has a continuous inverse, at least on the interval in question. Having established the first result, I move to considering the second, namely, that there is a topological feature that distinguishes the intervals (0, 1] and [0, 1]. Here, I will draw upon a more general result, namely, that there is a topological distinction between an open and a closed boundary. One can prove that compact (closed and bounded) sets can be mapped by continuous functions only to compact sets. Consequently, there is no continuous function mapping the compact set [0, 1] to the non-compact set (0, 1]; hence, (0, 1] and [0, 1] are not topologically equivalent [18, p. 121]. An anonymous referee claimed that the intervals (0, 1] and \((-\infty , 1]\) are topologically distinct even though the two intervals are homeomorphic. To be sure, if we consider the two intervals as sub-intervals of \({\mathbb {R}}\), then there are topological features that the two intervals do not share. Here are three examples: (i) the interval \((-\infty , 1]\) is closed in the real numbers \({\mathbb {R}}\), whereas (0, 1] is not; (ii) the complement of \((-\infty , 1]\) in \({\mathbb {R}}\) is connected, whereas the complement of (0, 1] is not; (iii) the closure of (0, 1] in \({\mathbb {R}}\) is compact, while the closure of \((-\infty , 1]\) is not. (Thanks to Brian Harbourne and Nung Kwan Yip for helpful discussion of this point.) While one may use any one of these features to claim that \((-\infty , 1]\) and (0, 1] are not “topologically equivalent”, all three are extrinsic features that depend upon how the intervals \((-\infty , 1]\) and (0, 1] are embedded within \({\mathbb {R}}\). Due to the fact that \((-\infty , 1]\) and (0, 1] are homeomorphic, there is no intrinsic way to distinguish the two. Note that I am using the intervals \((-\infty , 1]\) and (0, 1] as “stand-ins” for the Cosmos’s history; since the Cosmos is the totality of physical reality, the only features that matter for my purposes are the features that are intrinsic to an interval, and so, for my purposes, there is no relevant topological distinction between \((-\infty , 1]\) and (0, 1].

  5. Bimetric theories indistinguishable from standard General Relativity have been considered in [19,20,21,22,23, 24, pp. 335–336] A similar—though in principle observationally distinguishable—theory was considered in [21, 22, 25]; that theory approximates standard General Relativity arbitrarily well given a sufficiently small graviton mass. Bimetric theories have also been considered in [26,27,28].

  6. This claim is easy to motivate. Consider, for example, the case where we begin with Minkowski space and then construct a new spacetime \(S'\) using the procedure I’ve described. And now suppose, for reductio, that there exists a closed space-like boundary \({\mathcal {B}}\) shared by all of the time-like curves in \(S'\). There are two cases that we can consider: first, the case where \({\mathcal {B}}\) is “parallel” to \(\Sigma\), that is, the case where all of the time-like geodesics connecting \({\mathcal {B}}\) and \(\Sigma\) measure the same proper time at the point of intersection with \(\Sigma\), and, second, the case where \({\mathcal {B}}\) is not parallel to \(\Sigma\). In the first case, since all of the time-like geodesics measure the same proper time when they intersect \(\Sigma\), we are guaranteed that there are at least two time-like geodesics that measure the same age at the point at which they respectively intersect with \(\Sigma\). But, by construction, no two time-like geodesics have the same age at their respective points of intersection with \(\Sigma\). In the second case, \({\mathcal {B}}\) is tilted with respect to \(\Sigma\); in that case, there is some volume where \({\mathcal {B}}\) and \(\Sigma\) intersect. Past the volume where \({\mathcal {B}}\) and \(\Sigma\) intersect, \(\Sigma\) does not exist. For that reason, the spacetime will include at least one point that is not in the causal future or the causal past of some point in \(\Sigma\), so that \(\Sigma\) is not actually a spacetime-wide surface. Again, one of our assumptions has been violated. Thus, since the spacetime’s closed boundary cannot be a space-like surface in either case and the two cases are mutually exclusive and exhaustive, the spacetime’s closed boundary cannot be a space-like surface. While this result is much more difficult to establish when we allow for arbitrary spacetime curvature, a single example suffices to establish that one can construct spacetimes with the features I’ve described.

  7. That is, \(\varepsilon\) is a bijection from the positive real numbers to the set of time-like curves. We are guaranteed that this bijection exists because there is a bijection from n-dimensional space to the set of positive real numbers. Nonetheless, bijections between n-dimensional space and the positive real numbers are not generally smooth, thereby lending another reason why the boundary described is “jagged”.

  8. Skow cashes out his view in terms of absolute time, but indicates that he intends for his view to be generalizable to relativistic spacetimes.

  9. On some quantum gravity theories—such as causal set theory [45,46,47,48,49,50]—the spacetime metric appears only in the theory’s continuum limit, thereby allowing for the possibility that there are regions of the Cosmos where the spacetime metric is inapplicable. However, we should not necessarily think of those regions as amorphous in the sense discussed in this section. Consider, for example, Brightwell and Gregory’s [50] construction of the continuum limit for a spacetime interval spanned by a number of spacetime atoms “linked” together in a chain. When the chain is sufficiently long, the spacetime interval is proportional to the number of links in the chain. As causal set theorists like to say, in causal set theory, metrical facts are determined by counting. For that reason, supposing that there are only a small number of spacetime atoms in some region, so that the continuum limit does not apply in the region, we need only consider a larger region to recover relevant metrical facts. In any case, recall that the Boundary Condition for the Cosmos to have a beginning is disjunctive. If the initial portion of the Cosmos is correctly described by causal set theory, then, since causal sets always have closed boundaries, the Cosmos would satisfy the first disjunct—by having a topological beginning—and so would have a beginning.

  10. Nothing crucial in this example hangs on whether time is absolute. The example can be reconstructed for relativistic spacetimes. To construct a relativistic spacetime without metrical structure, first consider a spacetime S with metric \(g_{\mu \nu }\). And now consider the metric \({\tilde{g}}_{\mu \nu }\) produced from \(g_{\mu \nu }\) by the conformal transformation \({\tilde{g}}_{\mu \nu } = \Omega ^2 g_{\mu \nu }\) where \(\Omega\) is a positive and smooth but otherwise arbitrary scalar function. For relativistic spacetimes, multiplication by \(\Omega ^2\) leaves the spacetime’s light cone structure unaltered. Call the resulting spacetime \({\tilde{S}}\). Two spacetimes that are related by such a transformation, e.g., S and \({\tilde{S}}\), are said to be conformally equivalent A spacetime without metrical structure can then be constructed by identifying all of the members of a given class of conformally equivalent spacetimes. Let’s call the spacetime that results from identifying all of the members of a given class of conformally equivalent spacetimes \(S_C\). Since the conformal transformation left the light cone structure unaltered, \(S_C\) is equipped with light cone structure but not metrical structure and so \(S_C\) is an example of a relativistic metrically amorphous spacetime. To construct a relativistic spacetime analogous to the spacetime inhabited by Pam and Jim, one can “glue” a metrically amorphous spacetime region R between two regions that are not metrically amorphous.

  11. A similar point has been previously made in various places, but, in particular, see [51, pp. 125–126, 131] As Earman argues, in order for the Cosmos to have an objectively finite past, there must be an objective best choice of global time function according to which the past is finite. In contrast, Hermann Weyl [52] maintained that the choice of time scale is, is to a certain degree, conventional. In more technical terms, Weyl argued that there is gauge freedom in one’s choice of metric tensor so that the metric tensor is determined only up to a conformal factor, as in footnote 10. In any case, were Weyl’s theory correct, time scale would not correspond to any objective physical fact [53, p. 451]. Additional technical details for Weyl’s theory can be found in [54].

  12. An analogous construction can be produced using relativistic physics.

  13. Ideally, the account should also be consistent with a future quantum gravity theory, but, given that we do not yet possess a successful quantum gravity theory, the account that I offer here will need to be provisional.

  14. What do I mean by the ‘objective’ or ‘fundamental’ qualifier? Recall that one way to motivate a bimetric theory takes inspiration from Poincaré’s thought experiments in which there is an apparent metric due to the presence of universal forces on our measuring devices. By an ‘objective’ or ‘fundamental’ metric I mean a metric that is not merely apparent and that truly has metaphysical significance. I do not adopt a stance in this article on what precisely is requird for a metric to truly have metaphysical significance.

  15. The first spacetime atom in each temporal series of spacetime atoms need not be numerically identical to the first spacetime atom in any other temporal series of spacetime atoms. Note that this allows for an interesting possibility. There could be a discrete spacetime with a jagged edge, i.e., a discrete spacetime such that there are only a finite number of spacetime atoms to the past of every non-initial spacetime atom and yet there is no upper bound to the number of spacetime atoms to the past of non-initial spacetime atoms. For example, for every spacetime atom with n previous atoms, there may be another with \(n+1\) previous atoms.

  16. Perhaps the reader will object that one way that a series can have a boundary involves the series having a first member and having a first member has to do with the ordinal structure of the series. Nonetheless, the conjunction of the Direction Condition and the Boundary Condition successfully captures the Cosmos’s ordinal structure; for example, the Cosmos might have a closed boundary—and so satisfy the Topological Conception—and that closed boundary might be the Cosmos’s first moment in virtue of satisfying the Direction Condition. Moreover, the conjunction of the Direction Condition and the Boundary Condition capture a broader range of cases than if we defined the Boundary Condition in terms of the ordinal structure of spacetime.

  17. For some of the problems involved with utilizing b-incompleteness as the definitive feature of singular spacetimes, see chapter 2 in [61]; also see [69].

  18. Here, \({\mathcal {B}}\) need not be a space-like surface. Instead, I will understand \({\mathcal {B}}\) as the set of closed initial points for all of the time-like and light-like curves in the spacetime. In the case of a spacetime with a “jagged” boundary, there may be no simple relationship – and possibly no continuity – between the points in the set.

  19. This result will not necessarily follow for any curve that is not located in \({\mathcal {B}}\). For example, suppose that spacetime has a closed boundary but that the initial portion of the Cosmos has the “fractal” metrical properties discussed above. In that case, any time-like or light-like curve not located in \({\mathcal {B}}\) will have infinite backwards extension.

References

  1. Pitts, J.B.: Why the Big Bang singularity does not help the Kalām cosmological argument for theism. Br. J. Philos. Sci. 59(4), 675–708 (2008)

    Google Scholar 

  2. Matthews, G.: Time’s arrow and the structure of space-time. Philos. Sci. 46(1), 82–97 (1979)

    MathSciNet  Google Scholar 

  3. Castagnino, M., Lombardi, O., Lara, L.: The global arrow of time as a geometrical property of the universe. Found. Phys. 33(6), 877–912 (2003)

    MathSciNet  Google Scholar 

  4. Craig, W.L., Sinclair, J.: The Kal\(\overline{a}\)m cosmological argument. In: Craig, W.L., Moreland, J. (eds.) The Blackwell Companion to Natural Theology, pp. 101–201. Wiley-Blackwell, West Sussex (2009)

    Google Scholar 

  5. Craig, W.L.: God and real time. Relig. Stud. 26, 335–347 (1990)

    Google Scholar 

  6. Craig, W.L.: Creation and Divine Action. In: Routledge Companion to Philosophy of Religion, pp. 318–328. Routledge, London (2007)

    Google Scholar 

  7. Godfrey-Smith, W.: Beginning and Ceasing to Exist. Philos. Stud. 32(4), 393–402 (1977)

    Google Scholar 

  8. Monton, B.: Seeking God in Science: An Atheist Defends Intelligent Design. Broadview Press, Peterborough (2009)

    Google Scholar 

  9. Oderberg, D.: The Beginning of Existence. Int. Philos. Quart. 43(2), 145–158 (2003)

    Google Scholar 

  10. Mullins, R.: The End of the Timeless God. Oxford University Press, Oxford (2016)

    Google Scholar 

  11. Mullins, R.: Time and the everlasting god. Pittsbg. Theol. J. 3, 38–56 (2011)

    Google Scholar 

  12. Leon, F.: On Finitude, Topology, and Arbitrariness. In: Is God the Best Explanation of Things? A Dialogue, pp. 53–70. Palgrave Macmillan, Cham, Switzerland (2019)

    Google Scholar 

  13. Reichenbach, H.: The Direction of Time. University of California Press, Berkeley (1971). Edited by Maria Reichenbach

  14. Brown, H., Pooley, O.: The dynamical approach to spacetime theories. In: Knox, E., Wilson, A. (eds.) The Routledge Companion to Philosophy of Physics. Routledge, New York (2022)

    Google Scholar 

  15. Poincaré, H.: Science and hypothesis. In: Gould, S.J. (ed.) The Value of Science: Essential Writings of Henri Poincaré, pp. 2–178. Random House, New York (2001) . (Originally published in 1905)

    Google Scholar 

  16. Reichenbach, H.: The Philosophy of Space and Time. Dover, New York (1958)

    MATH  Google Scholar 

  17. Grünbaum, A.: Geometry and Chronometry. University of Minnesota Press, Minneapolis, MN (1968)

    MATH  Google Scholar 

  18. Wapner, L.: The Pea and the Sun: A Mathematical Paradox. CRC Press, Wellesley, MA (2005)

    MATH  Google Scholar 

  19. Pitts, J.B., Schieve, W.C.: Nonsingularity of flat Robertson-Walker models in the special relativistic approach to Einstein’s equations. Found. Phys. 33(9), 1315–1321 (2003)

    ADS  MathSciNet  Google Scholar 

  20. Pitts, J.B., Schieve, W.C.: Null cones and Einstein’s equations in Minkowski spacetime. Found. Phys. 34(2), 211–238 (2004)

    ADS  MathSciNet  MATH  Google Scholar 

  21. Pitts, J.B., Schieve, W.C.: Universally coupled massive gravity. Theor. Math. Phys. 151(2), 700–717 (2007)

    MATH  Google Scholar 

  22. Pitts, J.B.: Space-time constructivism vs. modal provincialism: or, how special relativistic theories needn’t show Minkowski chronogeometry. Stud. Hist. Philos. Sci. Part B 67, 191–198 (2019)

    MathSciNet  MATH  Google Scholar 

  23. Feynman, R., Moringo, F., Wagner, W.: Feynman Lectures on Gravitation. CRC Press, Boca Raton (2003). Edited by Brian Hatfield

  24. Lockwood, M.: The Labyrinth of Time: Introducing the Universe. Oxford University Press, Oxford (2007)

    Google Scholar 

  25. Pitts, J.B.: Kant, Schlick and Friedman on Space, Time and Gravity in Light of Three Lessons from Particle Physics. Erkenntnis 83, 135–161 (2018)

    MathSciNet  Google Scholar 

  26. Moffat, J.: Bimetric gravity theory, varying speed of light and the dimming of supernovae. Int. J. Mod. Phys. D. (2003). https://doi.org/10.1142/S0218271803002366

    Article  MATH  Google Scholar 

  27. Hossenfelder, S.: Bimetric theory with exchange symmetry. Phys. Rev. (2008). https://doi.org/10.1103/PhysRevD.78.044015

    Article  MathSciNet  MATH  Google Scholar 

  28. Hossenfelder, S.: Static scalar field solutions in symmetric gravity. Classical and Quantum Gravity 33, 185008 (2016)

    ADS  MathSciNet  MATH  Google Scholar 

  29. Brown, H.: Correspondence, invariance and heuristics in the emergence of special relativity. In: French, S., Kamminga, H. (eds.) Correspondence. Invariance and Heuristics. Kluwer Academic Publishers, Dordrecht (1993)

    Google Scholar 

  30. Brown, H.: On the role of special relativity in general relativity. Int. Stud. Philos. Sci. 11, 67–81 (1997)

    MathSciNet  MATH  Google Scholar 

  31. Brown, H.: Physical Relativity: Space-Time Structure from a Dynamical Perspective. Oxford University Press, Oxford (2005)

    MATH  Google Scholar 

  32. Brown, H., Pooley, O.: Minkowski space-time: a glorious non-entity. In: French, S., Kamminga, H. (eds.) The Ontology of Spacetime, pp. 67–89. Elsevier, Amsterdam (2006)

    Google Scholar 

  33. Brown, H., Pooley, O.: The Origins of the Space-Time Metric: Bell’s Lorentzian Pedagogy and its Significance in General Relativity. In: Callender, C., Huggett, N. (eds.) Physics Meets Philosophy at the Plank Scale (2001)

  34. DiSalle, R.: Understanding Space-time: The Philosophical Development of Physics from Newton to Einstein. Cambridge University Press, Cambridge (2006)

    MATH  Google Scholar 

  35. Knox, E.: Physical relativity from a functionalist perspective. Stud. Hist. Philos. Sci. Part B 67, 118–124 (2019)

    MathSciNet  MATH  Google Scholar 

  36. Swinburne, R., Bird, J.H.: Symposium: the beginning of the universe. Proc. Aristot. Soc. Suppl. Vol. 40, 125–150 (1966)

    Google Scholar 

  37. Halvorson, H., Kragh, H.: Cosmology and theology. In: Zalta, E.N. (ed.) The Stanford Encyclopedia of Philosophy, Spring, 2019th edn. Stanford University, n.p, Metaphysics Research Lab (2019)

  38. Milne, E.: Kinematic Relativity. Oxford University Press, Oxford (1948)

    MATH  Google Scholar 

  39. Misner, C.: Absolute zero of time. Phys. Rev. 186(5), 1328–1333 (1969)

    ADS  MATH  Google Scholar 

  40. Roser, P.: Gravitation and Cosmology with York Time, (2016). Doctoral dissertation from Clemson University

  41. Roser, P., Valentini, A.: Cosmological history in York time: inflation and perturbations. General Relativity and Gravitation 49(13) (2017)

  42. Malament, D.: The class of continuous timelike curves determines the topology of spacetime. J. Math. Phys. 18, 1399–1404 (1977)

    ADS  MathSciNet  MATH  Google Scholar 

  43. Skow, B.: Extrinsic temporal metrics. In: Zimmerman, D. (ed.) Oxford Studies in Metaphysics. Oxford University Press, Oxford (2010)

  44. Penrose, R.: Cycles of Time: An Extraordinary New View of the Universe. Vintage, New York (2012)

    MATH  Google Scholar 

  45. Bombelli, L., Lee, J., Meyer, D., Sorkin, R.: Space-time as a causal set. Phys. Rev. Lett. 59(5), 521–524 (1987)

    ADS  MathSciNet  Google Scholar 

  46. Dowker, F.: Causal sets as discrete spacetime. Contemp. Phys. 47(1), 1–9 (2006)

    ADS  MATH  Google Scholar 

  47. Dowker, F.: Introduction to causal sets and their phenomenology. Gen. Relativ. Gravit. 45, 1651–1667 (2013)

    ADS  MathSciNet  MATH  Google Scholar 

  48. Dowker, F.: Evolution of universes in causal set cosmology. C. R. Phys. 18, 246–253 (2017)

    ADS  Google Scholar 

  49. Dowker, F.: Being and becoming on the road to quantum gravity; or, the birth of a baby is not a baby. In: Huggett, N., Matsubara, K., Wüthrich, C. (eds.) Beyond Spacetime: The Foundations of Quantum Gravity, pp. 133–142. Cambridge University Press, Cambridge (2020)

    MATH  Google Scholar 

  50. Brightwell, G., Gregory, R.: Structure of random discrete spacetime. Phys. Rev. Lett. 66(3), 260–263 (1991)

    ADS  MathSciNet  MATH  Google Scholar 

  51. Earman, J.: Till the end of time. In: Earman, J., Glymour, C., Stachel, J. (eds.) Foundations of Space-Time Theories, pp. 109–133. University of Minnesota Press, Minneapolis, MN (1977)

    Google Scholar 

  52. Weyl, H.: Gravitation and electricity. In: O’Raifeartaigh, L. (ed.) The Dawning of Gauge Theory, pp. 24–37. Princeton University Press, Princeton (1997). Originally published in 1918

  53. Penrose, R.: The Road to Reality: A Complete Guide to the Laws of the Universe. Vintage, New York (2004)

    MATH  Google Scholar 

  54. Bell, J.L., Korté, H.: Hermann Weyl. In: Zalta, E.N. (ed.) The Stanford Encyclopedia of Philosophy, Winter, 2016th edn. Stanford University, n.p, Metaphysics Research Lab (2016)

  55. von Koch, H.: On a continuous curve without tangent constructible from elementary geometry. In: Classics on Fractals, pp. 25–46. Routledge, London (2018). Originally published in 1904

  56. Mandelbrot, B.: How Long is the Coastline of Britain? Statistical Self-Similarity and Fractional Dimension. In: Classics on Fractals, pp. 351–358. Routledge, London (2018). Originally published in 1967

  57. Hogarth, M.: Does general relativity allow an observer to view an eternity in finite time? Found. Phys. Lett. 5, 173–181 (1966)

    MathSciNet  Google Scholar 

  58. Earman, J., Norton, J.: Forever is a day: supertasks in Pitowsky and Malament-Hogarth spacetimes. Philos. Sci. 60, 22–42 (1993)

    MathSciNet  Google Scholar 

  59. Manchak, J.B., Roberts, B.W.: Supertasks. In: Zalta, E.N. (ed.) The Stanford Encyclopedia of Philosophy, Winter, 2016th edn. Stanford University, n.p, Metaphysics Research Lab (2016)

  60. Etesi, G., Németi, I.: Non-turing computations via Malament-Hogarth space-times. Int. J. Theor. Phys. 41, 341–370 (2002)

    MathSciNet  MATH  Google Scholar 

  61. Earman, J.: Bangs, Crunches, Whimpers, and Shrieks. Oxford University Press, Oxford (1995)

    Google Scholar 

  62. Norton, J.: Geometries in Collision: Einstein, Klein and Riemann. In: The Symbolic Universe: Geometry and Physics 1890-1930, pp. 128–144. Oxford University Press, Oxford (1999)

  63. Isham, C.: Prima facie questions in quantum gravity. Lect. Notes Phys. 434, 1–21 (1994). https://doi.org/10.1007/3-540-58339-4_13.

    Article  ADS  MathSciNet  Google Scholar 

  64. Maudlin, T.: Time, topology and physical geometry. Proc. Aristot. Soc. Suppl. 84, 63–78 (2010)

    Google Scholar 

  65. Maudlin, T.: Philosophy of Physics: Space and Time. Princeton University Press, Princeton (2012)

    MATH  Google Scholar 

  66. DeLanda, M.: Intensive Science and Virtual Philosophy. Bloomsbury Academic, London (2013)

    Google Scholar 

  67. North, J.: Physics, Structure, and Reality. Oxford University Press, Oxford (2021)

    MATH  Google Scholar 

  68. Curiel, E.: The analysis of singular spacetimes. Philos. Sci. 66, 119–145 (1999)

    MathSciNet  Google Scholar 

  69. Curiel, E.: Singularities and black holes. In: Zalta, E.N. (ed.) The Stanford Encyclopedia of Philosophy, Spring, 2021st edn. Stanford University, n.p, Metaphysics Research Lab (2021)

  70. Joshi, P.: Spacetime Singularities. In: Ashtekar, A., Petkov, V. (eds.) Springer Handbook of Spacetime, pp. 409–436. Springer, Berlin (2014)

    Google Scholar 

  71. Lam, V.: The singular nature of spacetime. Philos. Sci. 74(5), 712–723 (2007)

    MathSciNet  Google Scholar 

  72. Manchak, J.B.: On Feyerabend, General Relativity, and ‘Unreasonable’ Universes. In: Bschir, K., Shaw, J. (eds.) Interpreting Feyerabend: Critical Essays. Cambridge University Press, Cambridge (2021)

  73. Glymour, C.: Topology, Cosmology and Convention. Synthese 24(1/2), 195–218 (1972)

    MATH  Google Scholar 

  74. Glymour, C.: Indistinguishable Space-Times and the Fundamental Group. In: Earman, J., Glymour, C., Stachel, J. (eds.) Foundations of Space Time Theories, pp. 50–60. University of Minnesota Press, Minneapolis, MN (1977)

    Google Scholar 

  75. Malament, D.: Observationally indistinguishable space-times. In: Earman, J., Glymour, C., Stachel, J. (eds.) Foundations of Space Time Theories, pp. 61–80. University of Minnesota Press, Minneapolis, MN (1977)

    Google Scholar 

  76. Manchak, J.B.: Can we know the global structure of spacetime? Stud. Hist. Philos. Mod. Phys. 40, 53–56 (2009)

    MathSciNet  MATH  Google Scholar 

  77. Manchak, J.B.: What is a physically reasonable space-time? Philosophy of Science 78(3), 410–420 (2011)

    MathSciNet  Google Scholar 

  78. Manchak, J.B.: General Relativity as a Collection of Collections of Models. In: Madarász, J., Székely, G. (eds.) Hajnal Andréka and István Németi on Unity of Science: from Computing to Relativity Theory Through Algebraic logic. Springer, Cham, Switzerland (2021)

  79. Beisbart, C.: Can we justifiably assume the cosmological principle in order to break model underdetermination in cosmology? J. Gen. Philos. Sci. 40(2), 175–205 (2009)

    Google Scholar 

  80. Beisbart, C.: What is the spatiotemporal extension of the universe? Underdetermination according to Kant’s first antinomy and in present-day cosmology. HOPOS: J. Int. Soc. Hist. Philos. Sci. (2022)

  81. Norton, J.: Observationally indistinguishable spacetimes: a challenge for any inductivist. In: Morgan, G. (ed.) Philosophy of Science Matters: The Philosophy of Peter Achinstein. Oxford University Press, Oxford (2011)

  82. Butterfield, J.: On under-determination in cosmology. Stud. Hist. Philos. Sci. Part B46(A), 57–69 (2014)

    MATH  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Contributions

DL wrote and reviewed the entire manuscript. (This is a single author paper)

Corresponding author

Correspondence to Daniel Linford.

Ethics declarations

Competing Interests

The authors declare no competing interests.

Additional information

Publisher's Note

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

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Linford, D. On the Boundary of the Cosmos. Found Phys 53, 76 (2023). https://doi.org/10.1007/s10701-023-00718-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s10701-023-00718-6

Keywords

Navigation