Skip to main content
Log in

On the Carroll–Chen Model

  • Article
  • Published:
Journal for General Philosophy of Science Aims and scope Submit manuscript

Abstract

I argue that the Carroll–Chen cosmogonic model does not provide a plausible scientific explanation of the past hypothesis (the thesis that our universe began in an extremely low-entropy state). I suggest that this counts as a welcomed result for those who adopt a Mill–Ramsey–Lewis best systems account of laws and maintain that the past hypothesis is a brute fact that is a non-dynamical law.

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

Fig. 1

Similar content being viewed by others

Notes

  1. Albert (2000, 96).

  2. Both Albert and Loewer would add another non-dynamical law that is “a probability distribution (or density) over possible initial conditions that assigns a value 1 to PH and is uniform over those micro-states that realize PH” Loewer (2008, 157), where by ‘PH’ Loewer means the past hypothesis.

  3. Lewis (1973b, 73).

  4. Callender (2004b, 249). For more on the Mill–Ramsey–Lewis best-systems account (BSA), see Cartwright et al. (2005, 797–799), Earman (1984, 196–199), Lewis (1973b, 72–77, 1983, 365–368, 1994, 478–482), Loewer (2001, 619, 2012) and Ramsey (1990, 150). For criticisms of the BSA see Armstrong (1983, 70–71), Belot (2011, 70–72), and van Fraassen (1989, 48–51).

    Penrose (1989a, 391–482) seemed to regard something like the past hypothesis as a law; it is just that he understood the initial low-entropy state in terms of Weyl curvature which vanishes as one approaches the beginning of the universe.

  5. See the comments in Sklar (1993, 311–312).

  6. See e.g., Carroll (2006, 1132, 2008a, 48, 50), Carroll and Chen (2004, 3–4), Cf. Carroll (2010, 288). For the implicit claim that the state is improbable see Penrose (2007, 729–731) and Price (2004, 230–231). The fact that a state is unnatural does not necessarily imply that that state is improbable.

  7. Even some of those who would insist that such a state is scientifically brute believe that it could be explained. See Callender (2004a, 199), though cf. his comments in (2004b, 241).

  8. Some of these alternative approaches assume different understandings of our universe’s low-entropy state.

  9. Two points about presentation: First, the discussion will be mostly informal. I delve into mathematical details only when absolutely necessary, and I discuss the physical significance of the underlying technical work, citing that work at every turn for those who might want to explore the underlying mathematical physics. I apologize for the abundance of footnotes, and note here that I stand by my interpretive work, and quote not a few theoretical physicists who agree with my interpretation of the underlying technical details.

    Second, while the analysis in general rests on results already in the literature, these results find novel applications in my critique of the CC-M. For example, no one has articulated just how especially problematic the measure problem and N-bound validity are for the CC-M. No one, so far as I’m aware, has used the EGS or BGV theorems to object to metauniverse nucleation and quantum tunneling. Moreover, no one has voiced the specific worries I will express about the incompleteness of the CC-M. My use of an argument from causation for the impossibility of the background space of the CC-M has heretofore never been supplemented in the way I justify several of the key premises, and some specific points about the internal inconsistency of the model are completely novel.

  10. As Roger Penrose stated, the “early spatial uniformity represents the universe’s extraordinarily low initial entropy”, Penrose (2012, 76). See also Penrose (2007, 706–707). Most writing on the subject agree with Penrose here. See the broader discussions in Albrecht (2004, 371–374), Greene (2004, 171–175), North (2011, 327), Penrose (1979, 611–617, 1989b, 251–260, 2012, 73–79), Price (1996, 79–85, 2004, 227–228) and Wald (1984, 416–418, 2006, 395). See also Callender (2010, 47–51). Earman (2006, 417–418, cf. the comments on 427) is very skeptical of the contemporary orthodoxy on these matters.

  11. Let me say here what I’m concerned with when I discuss or mention the arrow of time. First, I am not interested in the asymmetry of time itself. I am, however, concerned with the asymmetry of the contents of the cosmos (on this distinction see Price 1996, 16–17; North 2011, 312). There are, therefore, many arrows of time, though some maintain that these arrows can be reduced to the thermodynamic arrow. It is this supposed principal arrow with which I’m worried when I comment on the arrow of time below.

  12. “The low entropy starting point is the ultimate reason that the universe has an arrow of time, without which the second law would not make sense.” Dyson et al. (2002, 1). Cf. (Bousso (2012, 2–3; 26) for a different view. The discussion of these sorts of issues in North (2011) is first-rate.

  13. Below, I call the universes that help compose the multiverse “metauniverses”. These are spawned by the ‘Universe’ (capital-U), the background de Sitter spacetime.

  14. Carroll and Chen (2004, 27). I’m borrowing their wording here. The quotation in context is about something different, viz., the fact that the background space–time is never in an equilibrium state because metauniverses can always be generated resulting in the further increase of entropy.

  15. Given that r0 = 0 and that Λ = 3/R2; where R corresponds to a positive constant, and r is the Schwarzschild radius. The equation is from de Sitter (1918, 230); but see also the discussion in de Sitter (1917, 7) and Earman (1995, 7).

  16. Penrose (2007, 747–748), Misner et al. (1973, 745); and for an extensive treatment of de Sitter and anti-de Sitter space–times see Hawking and Ellis (1973, 124–134); but see also the discussions in Bousso (1998, 2000, 19–21) and Ginsparg and Perry (1983, 245–251). I should add here that de Sitter space is also thought to have infinite volume. See Carroll and Chen (2004, 27), and see the nice illustration of the space in Carroll (2006, 1134).

  17. Figure 1 is based, in part, on Carroll’s own Fig. 87 in Carroll (2010, 363).

  18. Unless otherwise indicated, in this section just about everything I say about the CC-M holds for the MCC-M. Therefore, (again, unless I indicate otherwise) wherever one sees ‘CC-M’, read ‘MCC-M’ as well.

  19. Before I proceed, I should provide a bit of an apologetic for what I’m up to in this section. First, C&C are completely honest and humble about the CC-M’s incompleteness. I do not mean to mercilessly pile on their worries about how to complete the model. My contention below will be that given scientific realism and the fact that substantive portions of the CC-M are inconsistent and admittedly not well understood, one cannot plausibly maintain that the CC-M provides a legitimate explanation of the low-entropy state. That is an important academic and philosophical point. Second, subsequent sections of this paper criticize the model on the assumption that there are ways of providing the details. So even if one does not agree with the aforementioned contention, one will still have to respond to some damaging criticism.

  20. I have in mind the results of Steinhardt and Wesley (2009, 104026-4 ff.). Though cf. Koster and Postma (2011).

  21. Reminiscent of Chisholm’s understanding of events in his (1990, 419 see definition D11). See also Koons (2000, 31–43).

  22. See Bunzl (1979, 145 and 150).

  23. Kim defined “causal closure of the physical domain” as the thesis that “…any physical event that has a cause at time t has a physical cause at t.” Kim (1989b, 43, emphasis in the original). He says that explanatory exclusion is the principle that “[n]o event can be given more than one complete and independent explanation.” Kim (1989a, 79, emphasis in the original).

  24. De Muijnck does try to rescue a counterfactual account of causation from cases of overdetermination.

  25. Mackie’s term, from Mackie (1974, 43), cf. De Muijnck (2003, 65–66), Schaffer (2003, 28). Schaffer tells us that, “[q]uantitative overdetermination occurs whenever the cause is decomposable into distinct and independently sufficient parts” Schaffer (2003, 28).

  26. Paul’s term from Paul (2007).

  27. I should point out that Paul believes that the consequence of there being such prevalent constitutive overdetermination is “mysterious and problematic” Paul (2007, 277). Paul attempts to rid the world of such prevalence by arguing that fundamental causal relata are property instances shared by overlapping entities involved in obtaining causal relations. See Paul (2007, 282).

  28. Schaffer (2003, 42. n. 9) emphasis in the original, though I’ve reversed the order. The objection is tied to Kim (1989a). Some interpret Kim as suggesting that overdetermination just doesn’t involve the causation of an event by an object and the parts that compose it. See Sider (2003, 719. n. 2).

  29. In the discussion above, I assume a simple counterfactual analysis of causation, not unlike the one defended in Lewis (1973a) or the later more sophisticated analysis in Lewis (2004). My point could easily be adjusted so as to accommodate other, (perhaps) more plausible accounts of causation.

  30. Clarkson, Ghezelbash, and Mann stated, “…one implication of the N-bound (and the maximal mass conjecture) is that a quantum gravity theory with an infinite number of degrees of freedom (such as M-theory) cannot describe spacetimes with Λ = 0…” Clarkson et al. (2003, 360).

  31. His argument is explanatory: “It is hard to see what, other than the Λ–N correspondence, would offer a compelling explanation [of] why such disparate elements appear to join seamlessly to imply a simple and general result” (Bousso (2000, 18).

  32. Bousso (2000, 3). Even if the type of entropy in play is information-theoretic or Von Neumann entropy, that fact would be irrelevant. The main argument of this section still runs.

  33. Bousso (2000, 2). In subsequent discussion, I will sometimes speak of N-bound validity for a space–time. What I mean by such a judgment is that Eq. 2 (Proposition 5) holds for those space–times.

  34. Add to these assumptions the further two conditions that the system feature elements that are “spatially bounded”, and that the system has constant energy; Sklar (1993, 36).

  35. Throughout the remainder of the paper, one may read ‘MCC-M’ wherever one sees ‘CC-M’. All subsequent argumentation will be applicable to both.

  36. For some the following nagging objection will remain: Fields in QFT admit infinitely many degrees of freedom, therefore something is wrong with the above argumentation. The reasoning is out of touch with the contemporary state of the art in quantum cosmology. Numerous considerations suggest that QFT breaks down and most cosmologists (it seems) no longer believe that QFT will reside prominently in the endgame quantum theory of gravity. In fact, David J. Gross has said that “[t]he longstanding problem of quantizing gravity is probably impossible within the framework of quantum field theory… QFT has proved useless in incorporating quantum gravity into a consistent theory at the Planck scale. We need to go beyond QFT, to a theory of strings or to something else, to describe quantum gravity” Gross (1997, 10).

  37. The Copernican principle says, roughly, that our causal past and position in space–time is not unique or distinctive. See Stoeger et al. (1995, 1).

  38. Borrowing some wording from Smeenk (2013, 630–631). See also Stoeger et al. (1995, 1). There is a interesting discussion of these matters in Clarkson and Maartens (2010), Maartens (2011). It is important to add that the result from Ehlers, Geren, and Sachs does not extend to times prior to the decoupling era. They remarked, “the result presented cannot be taken to mean that the universe in its earliest stages was necessarily a Friedmann model…” Ehlers, Geren, and Sachs (1968, 1349, emphasis mine). The beginning of the era in question is the moment(s) at or during which radiation and matter decoupled (about 240,000 or so years after the initial singularity).

  39. (Clifton et al. (2012, 051303-1 to 051303-2). For more on the Sunyaev–Zel’dovich effect see Weinberg (2008, 132–135).

  40. It may be that in order to alleviate worries about fine-tuning and the cosmological constant, one should appropriate a scalar field model of dark energy. In addition, it seems that the best way of understanding dark energy via quintessence is to posit a scalar field model of dark energy. As Weinberg remarked, “[t]he natural way to introduce a varying vacuum energy is to assume the existence of one or more scalar fields, on which the vacuum energy depends, and whose cosmic expectation values change with time” (Weinberg 2008, 89). For more on dark energy and scalar field models of such energy, see Sahni (2002, 3439–3441).

  41. Clarkson and Maartens (2010, 2) stated,

    “Isotropy is directly observable in principle, and indeed we have excellent data to show that the CMB is isotropic about us to within one part in ~105 (once the dipole is interpreted as due to our motion relative to the cosmic frame, and removed by a boost).”

    Weinberg (2008, 129) confesses that treating the CMBR as perfectly isotropic and homogeneous is “a good approximation”. Weinberg went on to affirm that “the one thing that enabled Penzias and Wilson to distinguish the background radiation from radiation emitted by earth’s atmosphere was that the microwave background did not seem to vary with direction in the sky.” ([ibid.]).

  42. Clifton, Clarkson, and Bull embrace their idealized assumptions in Clifton et al. (2012, 051303-4).

  43. See Hawking and Ellis (1973, 353-354). For a discussion of the CMBR anisotropies, see Lyth and Liddle (2009, 152–169) and Weinberg (2008, 129–148).

  44. Ehlers et al. (1968) also ignored the cosmological constant.

  45. Maartens and Matravers (1994, 2701). These galactic observations correspond to propositions (O1)–(O4) in Maartens and Matravers (1994, 2694). They are not observations of isotropic background blackbody radiation. See also (Maartens (2011) and Hasse and Perlick (1999).

  46. Maartens’ discussion of the specific result I am interested in is an expansion on earlier work with Chris Clarkson in Clarkson and Maartens (2010).

  47. Maartens (2011, 5131).

  48. Such that Fμ = Fμν  = Fμνα = 0 holds (from Eq. 3.24 of Maartens (2011, 5125).

  49. See Maartens (2011, 5125, 5131).

  50. “FLRW models with ordinary matter have a singularity at a finite time in the past.” (Smeenk 2013, 612). Hawking and Ellis stated, “…there are singularities in any Robertson–Walker space–time in which μ > 0, p ≥ 0 and Λ is not too large.” Hawking and Ellis (1973, 142). See also Wald (1984, 213–214); and the discussion of FLRW models in Penrose (2007, 717–723).

  51. See the review of many of these theorems in Hawking and Ellis (1973, 261–275).

  52. (Hawking and Penrose 1970, 529). This paper also provides an excellent review of both Hawking and Penrose’s previous work on singularity theorems, see especially (ibid., 529–533).

  53. Hawking (1996, 19).

  54. The quoted bit is from (Hawking and Penrose 1970, 531). Of course, they were not concerned with inflationary cosmology in 1970. Here is the broader context of the quote, “…we shall require the slightly stronger energy condition given in (3.4), than that used in I. This means that our theorem cannot be directly applied when a positive cosmological constant λ is present.” (Hawking and Penrose 1970, 531, emphasis in the original). Many authors have noted that inflationary cosmological models violate the strong energy condition of the Hawking–Penrose theorem. See, for example Wall (2013, 20. n. 14); and Borde and Vilenkin (1996, 824. n. 17), inter alios.

  55. The three conditions are stated in Borde and Vilenkin (1996, 819).

  56. See Borde and Vilenkin (1997, 718–719).

  57. Borde et al. (2003).

  58. “The result depends on just one assumption: The Hubble parameter H has a positive value when averaged over the affine parameter of a past-directed null or noncomoving timelike geodesic.” (Borde et al. 2003, 151301-4). See also Mithani and Vilenkin (2012, 1) “…it [the BGV] states simply that past geodesics are incomplete provided that the expansion rate averaged along the geodesic is positive: H av  > 0.” (ibid.); and Vilenkin (2013, 043516-1).

  59. Vilenkin (2013, 043516-1).

  60. Borde et al. (2003, 151301-4), cf. Guth (2004, 49). See also Mithani and Vilenkin (2012, 1–2).

  61. Farhi et al. (1990, 419).

  62. Carroll and Chen (2004, 27. n. 6).

  63. Vilenkin stated, “[e]ven though the BGV theorem is sometimes called a ‘singularity theorem’, it does not imply the existence of spacetime singularities. All it says is that an expanding region of spacetime cannot be extended to the past beyond some boundary \(\mathcal{B}\)” Vilenkin (2013, 043516-1).

  64. See Carroll (2008b, 4, 2010, 50, 349–350, particularly 408. n. 277), but cf. Penrose (1996, 36) for a different view.

  65. Vilenkin (2006, 175) emphasis mine. Ashtekar (2009, 9) acknowledged that the BGV does not rely on Einstein’s field equations.

  66. It seems that C&C go in for a generalized second law. In their discussion of black hole entropy and Hawking radiation, they stated that “[o]ne can prove [69], [70], [71], [72] certain versions of the Generalized Second Law, which guarantees that the radiation itself, free to escape to infinity, does have a larger entropy than the original black hole.” Carroll and Chen (2004, 18).

  67. See Wall (2013). Cf. Jacobson (1994).

  68. Carroll and Chen (2004, 25). There is also the separate question of how likely or natural it is that our metauniverse’s large scale structure is due to some prior inflationary era. Carroll and Tam address this question to some degree in their (2010).

  69. That work also required a quantum theory with an infinitely dimensional Hilbert space.

References

  • Aguirre, A. Carroll, S. M., & Johnson, M. C. (2011). Out of equilibrium: Understanding cosmological evolution to lower-entropy states. arXiv:1108.0417v1.

  • Albert, D. Z. (2000). Time and chance. Cambridge, MA: Harvard University Press.

    Google Scholar 

  • Albert, D. Z. (2010). Probability in the Everett picture. In S. Saunders, J. Barrett, A. Kent, & D. Wallace (Eds.), Many worlds? Everett, quantum theory, & reality (pp. 355–368). Oxford: Oxford University Press.

    Chapter  Google Scholar 

  • Albert, D. Z. (2015). After physics. Cambridge, MA: Harvard University Press.

    Book  Google Scholar 

  • Albrecht, A. (2004). Cosmic inflation and the arrow of time. In J. D. Barrow, P. C. W. Davies, & C. L. Harper Jr (Eds.), Science and ultimate reality: Quantum theory, cosmology, and complexity (pp. 363–401). New York, NY: Cambridge University Press.

    Chapter  Google Scholar 

  • Armstrong, D. M. (1983). What is a law of nature?. Cambridge: Cambridge University Press.

    Book  Google Scholar 

  • Ashtekar, A. (2009). Singularity resolution in loop quantum cosmology: A brief overview. Journal of Physics: Conference Series, 189, 012003.

    Google Scholar 

  • Baker, D. (2007). Measurement outcomes and probability in Everettian quantum mechanics. Studies in History and Philosophy of Modern Physics, 38, 153–169.

    Article  Google Scholar 

  • Banks, T. (2000). Cosmological breaking of supersymmetry or little Lambda goes back to the future II. arXiv:hep-th/0007146v1.

  • Banks, T. (2007). Entropy and initial conditions in cosmology. arXiv:hep-th/0701146v1.

  • Banks, T. (2015). Holographic inflation and the low entropy of the early universe. arXiv:1501.02681v1 [hep-th].

  • Belot, G. (2011). Geometric possibility. New York: Oxford University Press.

    Book  Google Scholar 

  • Boddy, K. K., Carroll, S. M., & Pollack, J. (2014). De sitter space without quantum fluctuations. arxiv:1405.0298v1 [hep-th].

  • Borde, A., & Vilenkin, A. (1996). Singularities in inflationary cosmology: A review. International Journal of Modern Physics D, 5, 813–824.

    Article  Google Scholar 

  • Borde, A., & Vilenkin, A. (1997). Violation of the weak energy condition in inflating spacetimes. Physical Review D, 56, 717–723.

    Article  Google Scholar 

  • Borde, A., Guth, A. H., & Vilenkin, A. (2003). Inflationary spacetimes are incomplete in past directions. Physical Review Letters, 90, 151301.

    Article  Google Scholar 

  • Bousso, R. (1998). Proliferation of de Sitter space. Physical Review D, 58, 083511.

    Article  Google Scholar 

  • Bousso, R. (2000). Positive vacuum energy and the N-bound. Journal of High Energy Physics, 11(2000), 038.

    Article  Google Scholar 

  • Bousso, R. (2012). Vacuum structure and the arrow of time. Physical Review D, 86, 123509.

    Article  Google Scholar 

  • Bousso, R., DeWolfe, O., & Myers, R. C. (2003). Unbounded entropy in spacetimes with positive cosmological constant. Foundations of Physics, 33, 297–321.

    Article  Google Scholar 

  • Brand, M. (1977). Identity conditions for events. American Philosophical Quarterly, 14, 329–337.

    Google Scholar 

  • Bucher, M. (2002). A braneworld universe from colliding bubbles. Physics Letters B, 530, 1–9.

    Article  Google Scholar 

  • Bunzl, M. (1979). Causal overdetermination. The Journal of Philosophy, 76, 134–150.

    Article  Google Scholar 

  • Callender, C. (2004a). Measures, explanations and the past: Should ‘special’ initial conditions be explained? The British Journal for the Philosophy of Science, 55, 195–217.

    Article  Google Scholar 

  • Callender, C. (2004b). There is no puzzle about the low-entropy past. In C. Hitchcock (Ed.), Contemporary debates in philosophy of science (Contemporary Debates in Philosophy) (pp. 240–255). Malden: Blackwell.

  • Callender, C. (2010). The past hypothesis meets gravity. In G. Ernst & A. Hüttemann (Eds.), Time, chance and reduction: Philosophical aspects of statistical mechanics (pp. 34–58). Cambridge: Cambridge University Press.

    Chapter  Google Scholar 

  • Carroll, S. M. (2006). Is our universe natural? Nature, 440, 1132–1136.

    Article  Google Scholar 

  • Carroll, S. M. (2008a). The cosmic origins of time’s arrow. Scientific American, 298, 48–57.

    Article  Google Scholar 

  • Carroll, S. M. (2008b). What if time really exists? arXiv:0811.3772v1 [gr-qc].

  • Carroll, S. M. (2010). From eternity to here: The quest for the ultimate theory of time. New York, NY: Dutton.

    Google Scholar 

  • Carroll, S. M. (2012). Does the universe need god? In J. B. Stump & A. G. Padgett (Eds.), The blackwell companion to science and christianity (pp. 185–197). Malden: Blackwell.

    Chapter  Google Scholar 

  • Carroll, S. M., & Chen, J. (2004). Spontaneous inflation and the origin of the arrow of time. arXiv:hep-th/0410270v1.

  • Carroll, S. M., & Chen, J. (2005). Does inflation provide natural initial conditions for the universe. General Relativity and Gravitation, 37, 1671–1674.

    Article  Google Scholar 

  • Carroll, S. M., & Tam, H. (2010). Unitary evolution and cosmological fine-tuning. arXiv:1007.1417v1 [hep-th].

  • Carroll, S. M., Johnson, M. C., & Randall, L. (2009). Dynamical compactification from de Sitter space. Journal of High Energy Physics, 11(2009), 094.

    Article  Google Scholar 

  • Cartwright, N., Alexandrova, A., Efstathiou, S., Hamilton, A., & Muntean, I. (2005). Laws. In F. Jackson & M. Smith (Eds.), The oxford handbook of contemporary philosophy (pp. 792–818). New York: Oxford University Press.

    Google Scholar 

  • Chisholm, R. M. (1990). Events without times an essay on ontology. Noûs, 24, 413–427.

    Article  Google Scholar 

  • Clarkson, C. (2012). Establishing homogeneity of the universe in the shadow of dark energy. Comptes Rendus Physique, 13, 682–718.

    Article  Google Scholar 

  • Clarkson, C., & Maartens, R. (2010). Inhomogeneity and the foundations of concordance cosmology. Classical and Quantum Gravity, 27, 124008.

    Article  Google Scholar 

  • Clarkson, R., Ghezelbash, A. M., & Mann, R. B. (2003). Entropic N-bound and maximal mass conjecture violations in four-dimensional Taub–Bolt (NUT)-dS spacetimes. Nuclear Physics B, 674, 329–364.

    Article  Google Scholar 

  • Clifton, T., Clarkson, C., & Bull, P. (2012). Isotropic blackbody cosmic microwave background radiation as evidence for a homogenous universe. Physical Review Letters, 109, 051303.

    Article  Google Scholar 

  • Davidson, D. (1967). Causal relations. The Journal of Philosophy, 64, 691–703.

    Article  Google Scholar 

  • Davidson, D. (1969). The individuation of events. In N. Rescher (Ed.), Essays in honor of Carl G. Hempel (pp. 216–234). Dordrecht: Reidel.

    Chapter  Google Scholar 

  • Davidson, D. (2001). The individuation of events. In D. Davidson (Ed.), Essays on actions and events (2nd ed., pp. 163–180). Oxford: Clarendon Press.

    Chapter  Google Scholar 

  • Davies, P. C. W. (1983). Inflation and time asymmetry: Or what wound up the universe. Nature, 301, 398–400.

    Article  Google Scholar 

  • De Muijnck, W. (2003). Dependences, connections, and other relations: A Theory of mental causation. (Philosophical Studies Series 93) Dordrecht: Kluwer.

  • de Sitter, W. (1917). On Einstein’s theory of gravitation, and its astronomical consequences (Third Paper). Monthly Notices of the Royal Astronomical Society, 78, 3–28.

    Article  Google Scholar 

  • de Sitter, W. (1918). On the curvature of space. In Koninklijke Akademie van Wetenschappen KNAW, Proceedings of the section of sciences, Amsterdam (pp. 229–243). 20, part 1.

  • Dyson, L., Kleban, M., & Susskind, L. (2002). Disturbing implications of a cosmological constant. Journal of High Energy Physics, 10(2002), 011.

    Article  Google Scholar 

  • Earman, J. (1984). Laws of nature: The empiricist challenge. In R. J. Bogdan (Ed.), D.M. Armstrong (Profiles Vol. 4, pp 191–223). Dordrecht: Reidel.

  • Earman, J. (1995). Bangs, crunches, whimpers, and shrieks: Singularities and acausalities in relativistic spacetimes. New York: Oxford University Press.

    Google Scholar 

  • Earman, J. (2006). The “past hypothesis”: Not even false. Studies in History and Philosophy of Modern Physics, 37, 399–430.

    Article  Google Scholar 

  • Ehlers, J., Geren, P., & Sachs, R. K. (1968). Isotropic solutions of the Einstein–Liouville equations. Journal of Mathematical Physics, 9, 1344–1349.

    Article  Google Scholar 

  • Farhi, E., Guth, A., & Guven, J. (1990). Is it possible to create a universe in the laboratory by quantum tunneling? Nuclear Physics B, 339, 417–490.

    Article  Google Scholar 

  • Fischler, W., Morgan, D., & Polchinski, J. (1990). Quantization of false-vacuum bubbles: A hamiltonian treatment of gravitational tunneling. Physical Review D, 42, 4042–4055.

    Article  Google Scholar 

  • Geroch, R. (1966). Singularities in closed universes. Physical Review Letters, 17, 445–447.

    Article  Google Scholar 

  • Gibbons, G. W., & Hawking, S. W. (1977). Cosmological event horizons, thermodynamics, and particle creation. Physical Review D, 15, 2738–2751.

    Article  Google Scholar 

  • Ginsparg, P., & Perry, M. J. (1983). Semiclassical perdurance of de Sitter space. Nuclear Physics B, 222, 245–268.

    Article  Google Scholar 

  • Greene, B. (2004). The fabric of the cosmos: Space, time, and the texture of reality. New York: Alfred A. Knopf.

    Google Scholar 

  • Gross, D. J. (1997). The triumph and limitations of quantum field theory. arXiv:hep-th/9704139v1.

  • Guth, A. (2004). Inflation. In W. L. Freedman (Ed.), Measuring and modeling the universe (Carnegie Observatories Astrophysics Series Vol. 2, pp. 31–52) Cambridge: Cambridge University Press.

  • Hasse, W., & Perlick, V. (1999). On spacetime models with an isotropic hubble law. Classical and Quantum Gravity, 16, 2559–2576.

    Article  Google Scholar 

  • Hawking, S. W. (1965). Occurrence of singularities in open universes. Physical Review Letters, 15, 689–690.

    Article  Google Scholar 

  • Hawking, S. W. (1966a). The occurrence of singularities in cosmology. In Proceedings of the royal society of London. Series A, mathematical and physical sciences, Vol. 294, pp. 511–521.

  • Hawking, S. W. (1966b). The occurrence of singularities in cosmology. II. In Proceedings of the royal society of London. Series A, mathematical and physical sciences, Vol. 295, pp. 490–493.

  • Hawking, S. W. (1967). The occurrence of singularities in cosmology. III. Causality and singularities. In Proceedings of the royal society of London. Series A, mathematical and physical sciences, Vol. 300, pp. 187–201.

  • Hawking, S. W. (1996). Classical theory. In S. Hawking, & R. Penrose (Eds.), The nature of space and time. (Princeton Science Library) (pp. 3–26). Princeton: Princeton University Press.

  • Hawking, S. W., & Ellis, G. F. R. (1973). The large scale structure of space-time. New York: Cambridge University Press.

    Book  Google Scholar 

  • Hawking, S. W., & Penrose, R. (1970). The singularities of gravitational collapse and cosmology. In Proceedings of the royal society of London. Series A, mathematical and physical sciences, Vol. 314, pp. 529–548.

  • Jacobson, T. (1994). Black hole entropy and induced gravity. arXiv:gr-qc/9404039v1.

  • Kim, J. (1989a). Mechanism, purpose, and explanatory exclusion. In Philosophical perspectives. (Philosophy of Mind and Action Theory) (Vol. 3, pp. 77–108). Atascadero: Ridgeview Publishing Company.

  • Kim, J. (1989b). The myth of nonreductive materialism. Proceedings and Addresses of the American Philosophical Association, 63, 31–47.

    Article  Google Scholar 

  • Koons, R. C. (2000). Realism regained: An exact theory of causation, teleology, and the mind. New York: Oxford University Press.

    Google Scholar 

  • Koster, R., & Postma, M. (2011). A no-go for no-go theorems prohibiting cosmic acceleration in extra dimensional models. Journal of Cosmology and Astroparticle Physics, 12(2011), 015. doi:10.1088/1475-7516/2011/12/015.

    Article  Google Scholar 

  • Lewis, D. (1973a). Causation. The Journal of Philosophy, 70, 556–567.

    Article  Google Scholar 

  • Lewis, D. (1973b). Counterfactuals. Malden: Blackwell.

    Google Scholar 

  • Lewis, D. (1983). New work for a theory of universals. Australasian Journal of Philosophy, 61, 343–377.

    Article  Google Scholar 

  • Lewis, D. (1994). Humean supervenience debugged. Mind, 103, 473–490.

    Article  Google Scholar 

  • Lewis, D. (2004). Causation as influence. In J. Collins, N. Hall, & L. A. Paul (Eds.), Causation and counterfactuals (pp. 75–106). Cambridge, MA: MIT Press.

    Google Scholar 

  • Linde, A. (2004). Inflation, quantum cosmology, and the anthropic principle. In J. D. Barrow, P. C. W. Davies, & C. L. Harper (Eds.), Science and ultimate reality: Quantum theory, cosmology, and complexity (pp. 426–458). New York: Cambridge University Press.

    Chapter  Google Scholar 

  • Loewer, B. (2001). Determinism and chance. Studies in History and Philosophy of Modern Physics, 32, 609–620.

    Article  Google Scholar 

  • Loewer, B. (2008). Why there is anything except physics. In J. Hohwy & J. Kallestrup (Eds.), Being reduced: New essays on reduction, explanation, and causation (pp. 149–163). Oxford: Oxford University Press.

    Chapter  Google Scholar 

  • Loewer, B. (2012). Two accounts of laws and time. Philosophical Studies, 160, 115–137.

    Article  Google Scholar 

  • Lyth, D. H., & Liddle, A. R. (2009). The primordial density perturbation: Cosmology, inflation and the origin of structure. Cambridge: Cambridge University Press.

    Book  Google Scholar 

  • Maartens, R. (2011). Is the universe homogeneous? Philosophical Transactions of the Royal Society A: Mathematical, Physical & Engineering Sciences, 369, 5115–5137.

    Article  Google Scholar 

  • Maartens, R., & Matravers, D. R. (1994). Isotropic and semi-isotropic observations in cosmology. Classical and Quantum Gravity, 11, 2693–2704.

    Article  Google Scholar 

  • Mackie, J. L. (1974). The cement of the universe: A study of causation. Oxford: Oxford University Press.

    Google Scholar 

  • Maldacena, J., & Nuñez, C. (2001). Supergravity description of field theories on curved manifolds and a no go theorem. International Journal of Modern Physics A, 16, 822–855.

    Google Scholar 

  • Maudlin, T. (2007). The metaphysics within physics. New York: Oxford University Press.

    Book  Google Scholar 

  • McInnes, B. (2007). The arrow of time in the landscape. arXiv:0711.1656v2 [hep-th].

  • Misner, C. W., Thorne, K. S., & Wheeler, J. A. (1973). Gravitation. San Francisco: W.H. Freeman and Company.

    Google Scholar 

  • Mithani, A., & Vilenkin, A. (2012). Did the universe have a beginning?. arXiv:1204.4658v1 [hep-th].

  • Nikolić, H. (2004) Comment on “spontaneous inflation and the origin of the arrow of time”. arXiv:hep-th/0411115v1.

  • North, J. (2011). Time in thermodynamics. In C. Callender (Ed.), The oxford handbook of philosophy of time (pp. 312–350). New York: Oxford University Press.

    Google Scholar 

  • Page, D. N. (2008). Return of the Boltzmann brains. Physical Review D, 78, 063536.

    Article  Google Scholar 

  • Paul, L. A. (2007). Constitutive overdetermination. In J. K. Campbell, M. O’Rourke, & H. Silverstein (Eds.), Causation and explanation (pp. 265–290). Cambridge, MA: MIT Press.

    Google Scholar 

  • Penrose, R. (1965). Gravitational collapse and space-time singularities. Physical Review Letters, 14, 57–59.

    Article  Google Scholar 

  • Penrose, R. (1979). Singularities and time-asymmetry. In S. W. Hawking & W. Israel (Eds.), General relativity: An Einstein centenary survey (pp. 581–638). New York: Cambridge University Press.

    Google Scholar 

  • Penrose, R. (1989a). The emperor’s new mind. New York: Oxford University Press.

    Google Scholar 

  • Penrose, R. (1989b). Difficulties with inflationary cosmology. Annals of the New York Academy of Sciences, 571, 249–264.

    Article  Google Scholar 

  • Penrose, R. (1996). Structure of spacetime singularities. In S. Hawking, & R. Penrose (Eds.), The nature of space and time. (Princeton Science Library) (pp. 27–36). Princeton: Princeton University Press.

  • Penrose, R. (2007). The road to reality: A complete guide to the laws of the universe. New York: Vintage Books.

    Google Scholar 

  • Penrose, R. (2012). Cycles of time: An extraordinary new view of the universe. New York: Vintage Books.

    Google Scholar 

  • Price, H. (1996). Time’s arrow and archimedes’ point: New directions for the physics of time. New York: Oxford University Press.

    Google Scholar 

  • Price, H. (2004). On the origins of the arrow of time: Why there is still a puzzle about the low-entropy past. In C. Hitchcock (Ed.), Contemporary debates in philosophy of science (Contemporary Debates in Philosophy) (pp. 219–239). Malden: Blackwell.

  • Ramsey, F. P. (1990). Law and causality. In D. H. Mellor (Ed.), Philosophical papers (pp. 140–163). Cambridge: Cambridge University Press.

    Google Scholar 

  • Rea, M. (1998). In defense of mereological universalism. Philosophy and Phenomenological Research, 58, 347–360.

    Article  Google Scholar 

  • Sahni, V. (2002). The cosmological constant problem and quintessence. Classical and Quantum Gravity, 19, 3435–3448.

    Article  Google Scholar 

  • Schaffer, J. (2003). Overdetermining causes. Philosophical Studies, 114, 23–45.

    Article  Google Scholar 

  • Sider, T. (2003). What’s so bad about overdetermination? Philosophy and Phenomenological Research, 67, 719–726.

    Article  Google Scholar 

  • Simons, P. (2003). Events. In Michael J. Loux & Dean W. Zimmerman (Eds.), The oxford handbook of metaphysics (pp. 357–385). New York: Oxford University Press.

    Google Scholar 

  • Sklar, L. (1993). Physics and chance: Philosophical issues in the foundations of statistical mechanics. New York: Cambridge University Press.

    Book  Google Scholar 

  • Smeenk, C. (2013). Philosophy of cosmology. In Robert Batterman (Ed.), The oxford handbook of philosophy of physics (pp. 607–652). New York: Oxford University Press.

    Google Scholar 

  • Smolin, L. (2002). Quantum gravity with a positive cosmological constant. arXiv:hep-th/0209079v1.

  • Steinhardt, P. J. (2011). The inflation debate: Is the theory at the heart of modern cosmology deeply flawed? Scientific American, 304(4), 36–43.

    Article  Google Scholar 

  • Steinhardt, P. J., & Turok, N. (2002a). Cosmic evolution in a cyclic universe. Physical Review D, 65, 126003.

    Article  Google Scholar 

  • Steinhardt, P. J., & Turok, N. (2002b). A cyclic model of the universe. Science, 296, 1436–1439.

    Article  Google Scholar 

  • Steinhardt, P. J., & Turok, N. (2005). The cyclic model simplified. New Astronomy Reviews, 49, 43–57.

    Article  Google Scholar 

  • Steinhardt, P. J., & Wesley, D. (2009). Dark energy, inflation, and extra dimensions. Physical Review D, 79, 104026.

    Article  Google Scholar 

  • Stoeger, W. R., Maartens, R., & Ellis, G. F. R. (1995). Proving almost-homogeneity of the universe: An almost Ehlers–Geren–Sachs theorem. The Astrophysical Journal, 443, 1–5.

    Article  Google Scholar 

  • Strominger, A. (2001). The ds/CFT correspondence. Journal of High Energy Physics, 10(2001), 029.

    Google Scholar 

  • Vachaspati, T. (2007). On constructing baby universes and black holes. arXiv:0705.2048v1 [gr-qc].

  • Van Cleve, J. (2008). The moon and sixpence: A defense of mereological universalism. In T. Sider, J. Hawthorne, & D. W. Zimmerman (Eds.), Contemporary Debates in metaphysics (Contemporary Debates in Philosophy) (pp. 321–340). Malden: Blackwell.

  • van Fraassen, B. C. (1989). Laws and symmetry. New York: Oxford University Press.

    Book  Google Scholar 

  • van Inwagen, P. (1990). Material beings. Ithaca: Cornell University Press.

    Google Scholar 

  • Van Riet, T. (2012). On classical de Sitter solutions in higher dimensions. Classical and Quantum Gravity, 29, 055001.

    Article  Google Scholar 

  • Vilenkin, A. (2006). Many worlds in one: The search for other universes. New York: Hill and Wang.

    Google Scholar 

  • Vilenkin, A. (2013). Arrows of time and the beginning of the universe. Physical Review D, 88, 043516.

    Article  Google Scholar 

  • Wald, R. M. (1984). General relativity. Chicago: University of Chicago Press.

    Book  Google Scholar 

  • Wald, R. M. (2006). The arrow of time and the initial conditions of the universe. Studies in History and Philosophy of Modern Physics, 37, 394–398.

    Article  Google Scholar 

  • Wall, A. C. (2012). Proof of the generalized second law for rapidly changing fields and arbitrary horizon slices. Physical Review D, 85, 104049.

    Article  Google Scholar 

  • Wall, A. C. (2013). The generalized second law implies a quantum singularity theorem. Classical and Quantum Gravity, 30, 165003.

    Article  Google Scholar 

  • Wallace, D. (2012). The emergent multiverse: Quantum theory according to the everett interpretation. Oxford: Oxford University Press.

    Book  Google Scholar 

  • Weaver, C. (2012). What could be caused must actually be caused. In Synthese. 184, 299–317 with ‘Erratum to: What could be caused must actually be caused’. Synthese. 183, 279.

  • Weaver, C. (2013). A church-fitch proof for the universality of causation. Synthese, 190, 2749–2772.

    Article  Google Scholar 

  • Weinberg, S. (2008). Cosmology. New York: Oxford University Press.

    Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Christopher Gregory Weaver.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Weaver, C.G. On the Carroll–Chen Model. J Gen Philos Sci 48, 97–124 (2017). https://doi.org/10.1007/s10838-016-9337-9

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10838-016-9337-9

Keywords

Navigation