Skip to main content

The Significance of Relativistic Computation for the Philosophy of Mathematics

  • Chapter
  • First Online:
Hajnal Andréka and István Németi on Unity of Science

Part of the book series: Outstanding Contributions to Logic ((OCTR,volume 19))

  • 322 Accesses

Abstract

In the paper I discuss the importance of relativistic hypercomputation for the philosophy of mathematics, in particular for our understanding of mathematical knowledge. I also discuss the problem of the explanatory role of mathematics in physics and argue that relativistic computation fits very well into the so-called programming account. Relativistic computation reveals an interesting interplay between the empirical realm and the realm of very abstract mathematical principles that even exceed standard mathematics and suggests, that such principles might play an explanatory role. I also argue that relativistic computation does not have some of the weaknesses of other hypercomputational models, thus it is particularly attractive for the philosophy of mathematics.

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 84.99
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 109.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD 109.99
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.

    These ideas were also presented by D. Malament. For a brief history, see Andréka et al. [2]. For a general discussion, see for instance also Earman and Norton [29], Etesi and Németi [32], Hogarth [47, 48], Shagrir and Pitowsky [87].

  2. 2.

    The best-known and most discussed example is the four-color theorem, proved in 1976 [7, 8]. A more contemporary example is Kepler’s conjecture concerning optimal packing of spheres in space, which was proved by Hales [39, 40]. There is an intense discussion concerning the status of computer proofs, in particular concerning the status of “empirical input”, which is present as we have to rely on strong and sophisticated physical theories (like quantum mechanics) in order to believe that our computer performs what it is supposed to so that it is reliable. This shows that the problem of the “empirical ingredient” in mathematics cannot be simply dismissed.

  3. 3.

    In some cases, quantum computing provides a speed-up effect, the most spectacular example being Shor’s quantum factoring algorithm [88, 89]. So, if quantum computers existed, they could perhaps speedup at least some of the proofs. The importance of quantum computation for the philosophy of mathematics is discussed in Wójtowicz [106, 108]. For an introduction to quantum computing, see for example Nielsen and Chuang [70]; a popular survey on quantum algorithms is Montanaro [68]. An important paper also discussing the philosophical issues is Deutsch et al. [28].

  4. 4.

    Cf. e.g. the special issues of Minds and Machines (12, 2002 13, 2003), Applied Mathematics and Computation (178, 2006), Theoretical Computer Science (317, 2004), Parallel Processing Letters (3, 2012). Syropoulos [97] is a survey of the hypercomputational models; a recent survey of analogue models of computation is given in Bournez and Pouly [14]; a general survey of physical models is Piccinini [75]; for a general discussion on the physical Church–Turing thesis see e.g. Cotogno [21] or Piccinini [74].

  5. 5.

    A truly clear description is given e.g. in Andréka et al. [2], Németi and Dávid [71], Etesi and Németi [32]. I follow these papers, therefore the presentation of the RTM model given here can be viewed as an extended (and modified) quotation.

  6. 6.

    A computer performing 1010000 computational steps needs energy, it might break down, the black hole might evaporate etc. Another difficulty is the communication problem addressed in Németi and Dávid [71]. Wüthrich [109] discusses it in the context of quantum mechanics, considering viable solutions exploiting quantum entanglement. As getting into a black hole is not particularly attractive, in Andréka et al. [6] a scenario is discussed that is more optimistic for the observer and is based on the wormhole hypothesis. It is also interesting to observe that the possibility of performing a hypercomputational task (like checking the consistency of ZFC) within Special Relativity is equivalent to the existence of superluminal signals [73].

  7. 7.

    In general, this is a subset of Δ2: equality holds under the assumption that there is no fixed upper bound on the number of signals sent on a finite path. Theoretical results of a different type (concerning ordinal computation times) are given in Hamkins and Lewis [44] (Hamkins [42] is a more popular presentation). It is known Stannett [95] that exactly the countable ordinal times (i.e. times < ω1) can be imbedded into real numbers. This opens space for investigations concerning the relationships between general notions of analogue computability and ordinal computability.

  8. 8.

    We might also think of two RTMs, one of which tries to prove TPC, while the other tries to prove its negation. Then we will either obtain (i) no signal (TPC is independent of ZFC), (ii) one signal (we will learn whether it is TPC or its negation which is provable in ZFC), or (iii) two signals, which means that ZFC is inconsistent. So, this scenario involves an implicit consistency check.

  9. 9.

    A clear expression of the thesis (in the context of computer-assisted proofs) is given by Rota: “Verification is proof, but verification may not give the reason. What, then, are we to mean by a reason” [84, 187]. An eminent mathematician, Mordell, claims that “Even when a proof has been mastered, there may be a feeling of dissatisfaction with it, though it may be strictly logical and convincing: such as, for example, the proof of a proposition in Euclid. The reader may feel that something is missing. The argument may have been presented in such a way as to throw no light on the why and wherefore of the procedure or on the origin of the proof or why it succeeds” [69, 11] (quotation from Mancosu [66, 142]). For an introduction to the problem of explanation within mathematics, see Mancosu [6567].

  10. 10.

    This claim might be considered controversial (I am indebted to the anonymous referee for the comments which inspired this footnote). For sufficiently short proofs understanding would not pose a problem—but I think that our understanding of even a short, one-page long formal proof in ZFC would consist in appealing to semantics. We would not treat the string “∀x,y∃z∀t (t∈z⇔(t∈x ∨ t∈y))” as meaningless, but rather would recognize it as the Union Axiom. Similarly we would try to identify the meaning of the formulas in the proof—and we would grasp the essence of the formal manipulations because we would interpret them in some way. Even understanding of hypothetical 1000-pages long formal proofs would consist in “translating” them into the “everyday mathematical concepts” (i.e. for instance we would think of lines 3235-5298 of the proof as of the application of the Central Limit Theorem—and this would constitute our understanding). However the relationship between the expertise in manipulating symbols and understanding is intricate as the never-ending discussions concerning Chinese room and its numerous variants show.

  11. 11.

    An oracle, “Pythiagora”, is described in Rav [82]; the author claims that it would have a destructive impact on mathematics as proof lies at the heart of mathematics, not theorems. A comparable situation is also considered in Tymoczko [101] (the first philosophical paper on the proof of the four-color theorem): Simon is a mathematical genius who solves the most difficult open problems in mathematics. But after some time, Simon refuses to present any proofs as he considers them to be too complex for us to understand. Should we trust him, and replace the standard proof (being a warrant for the truth of mathematical claims) by the phrase “Simon says”? The difference between Pythiagora, Simon, and the RTM is substantial: we know how the RTM works, so perhaps the oracle metaphor is misguided? However, from the purely mathematical point of view, the difference seems to be not so big: we only get the output, without any information concerning the proof.

  12. 12.

    The relationships between real and formal proofs in mathematics has been receiving much attention in recent years, see for instance the recent Hamami [41] and the extensive bibliography within. For an illuminating discussion concerning the role of proofs see Rav [82], Dawson [27].

  13. 13.

    See for instance Hauser [45] for a philosophical discussion and arguments in favor of the thesis that CH can be considered a legitimate, meaningful mathematical question.

  14. 14.

    Large cardinals are not a remedy: Levy and Solovay showed that assuming the existence of measurable, compact, Ramsey cardinals is consistent with both CH and ¬CH [59]. Moreover‚ Easton [30] proves that the value of the continuum can be almost arbitrary as the κ → 2κ function can behave in hugely different ways.

  15. 15.

    Such a discussion is important in modern set theory; cf. the discussion Feferman [33], Friedman [34], Maddy [62], Steel [96] (or more recent Maddy [63]). The recent discussion on the multiverse and hyperuniverse approach in set theory or the inner model program has a similar flavor: the aim is to find some axioms that give a true description of the set-theoretic world. See Hamkins [43], Arrigoni and Friedman [9].

  16. 16.

    In Szabó [98] a radical view is defended according to which formal systems can be treated as (embodied in) physical objects. This goes with a formalistic and physicalistic philosophy of mathematics, therefore my claim concerning the unnaturalness of (2) could not be accepted from his point of view.

  17. 17.

    Results concerning the necessary properties of numbers needed in physical theories fit this picture exactly. Such results are presented in Andréka et al. [5]; a detailed study of the properties needed to model accelerating observers is given in Székely [100].

  18. 18.

    The literature is already so vast that it would not make much sense to cite it here; I only mention [10, 11], which are important early contributions.

  19. 19.

    There are many examples discussed in the literature; one of the simplest is Euler’s theorem, which explains why it is not possible to cross the famous system of bridges in Königsberg (crossing every bridge exactly once). No physical details of this system are relevant for the explanation. There are of course much more advanced examples, to mention a few: (i) The Borsuk-Ulam theorem explains why there must be two antipodal points on the surface of the Earth where two physical parameters (say: temperature and pressure) are equal (cf. Baker [10, 11], Baker and Colyvan [12]); (ii) Lyon and Colyvan [60] discuss the explanatory role of phase spaces in physics, in which theorems on these spaces explain some phenomena; (iii) Baron [13] discusses a theorem on stochastic processes which can be considered to provide an explanation of the behavior of predators.

  20. 20.

    A recent survey of non-causal (in particular: mathematical) explanations is for instance [83].

  21. 21.

    The anonymous referee pointed out that this is really a case-by-case check and the result of this procedure is that it is not the case that we can prove 0 = 1 in PA. In order to justify Con(PA) we need one more step, i.e. the double-negation elimination. It is therefore non-constructive—and indirect.

  22. 22.

    This is true under the assumption of consistency of ZFC (if ZFC is inconsistent, we can prove everything, in particular Con(PA)). If we already know that ZFC is consistent (for instance by proving it in a stronger theory), we can use this fact as an explanation for the fact, that our procedure ends in the expected way.

  23. 23.

    In fact, our two RTMs from footnote 8 that check the truth/falsehood of the twin prime conjecture are performing (as a nice “byproduct” of the primary task) an “indirect Con(ZFC) RTM check”.

  24. 24.

    In fact, this is also the best explanation of the fact that no (human) mathematician has ever found an inconsistency. You cannot find something which does not exist, and the detailed story of your search does not contribute to the explanation.

  25. 25.

    See Feferman [33], Friedman [34], Maddy [62], Maddy [63], Steel [96] on the problem of justifying new axioms for set theory. I use the term “reliable” is the sense, that ZFC + A is reliable if the (new) axiom A is justified. Unfortunately, this problem exceeds vastly the scope of this study.

  26. 26.

    Trivially, ZFC + Con(ZFC) proves Con(ZFC) (and T + Con(T) proves Con(T)…), but this choice of theory explaining the outcome is ad hoc.

  27. 27.

    It is natural to think of such progressions in foundational studies. Gödel considered the reasonableness of formulating increasingly stronger axioms of infinity (which is often called “Gödel’s program”) and the research on large cardinals is one of the central parts of contemporary set theory. We can also think of reflection principles that express beliefs about the similarity between the universe and its parts. All these new axioms exceed ZFC.

  28. 28.

    In Reverse Mathematics, mathematical notions are encoded in the language of second-order arithmetic Z2, and the main focus of study is identifying the strength of axioms (often: set existence axioms) necessary to prove ordinary mathematical theorems. Simpson [92] is the main reference.

  29. 29.

    Further, we intend to analyze the logical structure of the theory: which assumptions are responsible for which predictions; what happens if we weaken/fine-tune the assumptions, what we could have done differently. We seek insights, a deeper understanding. We could call this approach ‘reverse relativity’ in analogy with ‘reverse mathematics’” Andreka et al. [1, 608]. An elaboration of these ideas in the context of why-questions within axiomatic relativity is given in Székely [99]. An important and interesting feature of this approach is that theories – not sentences – are considered to be answers to particular why-questions. So, loosely speaking, the content necessary to present an explanation is presented as a theory (which seems to suit scientific needs better than just presenting a single principle/axiom).

  30. 30.

    A topic I find particularly interesting is the choice of first-order logic as the framework (motivations were given for instance in Andréka et al. [4]) and the impact of this choice on the overall philosophical picture.

  31. 31.

    Performing an RTM check is not performing a supertask (i.e. a computation with, informally speaking, unbounded acceleration). The computation proceeds in an ordinary way, the hypercomputational effect appears due to the relationship between the observer’s and the computer’s time. We might view this effect due to an imbedding of the computer’s time into the observer’s time without preserving cofinality, which allows the outcome of the computational process within a bounded time to be ‘observed’ (cf. Stannett [95, 14]). This is not an inherent feature of the computing subject (computer/computer), in particular we do not have to resolve puzzles like “What happens just after the infinite accelerating computation terminates?”, as in Thompson-lamp-like examples. (For supertasks see Manchak and Roberts [64] or the former entry Laraudagoitia [58].

  32. 32.

    Such doubts concern, for example, the ARNN model [90, 91], which—in cases in which the coefficients are computable—is equivalent to the standard Turing model. Davis [26] claims therefore that the model does not really go beyond the Turing limit in any reasonable sense. See also Davis [25] for a forceful critique of the very idea of hypercomputation. See also Copeland [1619] for a general discussion of the notion of computation.

  33. 33.

    In a series of papers, Da Costa and Doria [20, 23, 24] provide interesting examples of sentences that are undecidable within ZFC and that seem to have a natural physical interpretation. (In Chaitin et al. [15] the results are put together and given a popular presentation.) Independence results in the language of relativity theory (involving models based on Archimedean fields) are presented in Andréka et al. [3].

  34. 34.

    It is also not clear whether there might be hypercomputation on the quantum level. Kieu [51, 52] defined such a quantum system: a retort is given in Smith [93]. The status of quantum hypercomputation is not settled. In Cubitt et al. [22] a quantum system is constructed for which the spectral gap problem is undecidable; so, if it were possible to verify it by empirical methods, it would yield an “empirical solution” to an undecidable problem.

  35. 35.

    “Ohne ein Quentchen Metaphysik läßt sich, meiner Überzeugung nach, keine exacte Wissenschaft begründen.” (from Cantor’s Nachlass).

References

  1. Andréka, H., Madarász, J., & Németi, I. (2007). Logic of space-time and relativity theory. In M. Aiello, I. Pratt-Hartmann, & J. Van Benthem (Eds.), Handbook of spatial logics. Dordrecht: Springer.

    Google Scholar 

  2. Andréka, H., Németi, I., & Németi, P. (2009). General relativistic hypercomputing and foundation of mathematics. Natural Computing, 8(3), 499–516.

    Article  Google Scholar 

  3. Andréka, H., Madarász, J., & Németi, I. (2012). Decidability, undecidability, and Gödel’s incompleteness in relativity theories. Parallel Processing Letters, 22(3).

    Google Scholar 

  4. Andréka, H., Madarász, J. X., Németi, I., Németi, P., & Székely, G. (2011). Vienna Circle and logical analysis of relativity theory. In A. Máté, M. Rédei, & F. Stadler (Eds.), Der Wiener Kreis in Ungarn/The Vienna Circle in Hungary (Vol. 16). Veröffentlichungen des Instituts Wiener Kreis. Vienna: Springer.

    Google Scholar 

  5. Andréka, H., Madarász, J., Németi, I., & Székely, G. (2012). What are the numbers in which spacetime? arXiv:1204.1350v1 [gr-qc].

  6. Andréka, H., Németi, I., & Székely, G. (2012). Closed timelike curves in relativistic computation. Parallel Processing Letters (3).

    Google Scholar 

  7. Appel, K., & Haken, W. (1977). Every planar map is four colorable, part I: Discharging. Illinois Journal of Mathematics, 21, 429–490.

    Google Scholar 

  8. Appel, K., Haken, W., & Koch, J. (1977). Every planar map is four colorable, part II: Reducibility. Illinois Journal of Mathematics, 21, 491–567.

    Google Scholar 

  9. Arrigoni, T., & Friedman, S.-D. (2013). The hyperuniverse program. Bulletin of Symbolic Logic, 19(1), 77–96.

    Article  Google Scholar 

  10. Baker, A. (2005). Are there genuine mathematical explanations of physical phenomena? Mind, 114(454), 223–238.

    Article  Google Scholar 

  11. Baker, A. (2009). Mathematical explanations in science. British Journal for the Philosophy of Science, 60(3), 611–633.

    Article  Google Scholar 

  12. Baker, A., & Colyvan, A. (2011). Indexing and mathematical explantation. Philosophia Mathematica, 19, 224–232.

    Article  Google Scholar 

  13. Baron, S. (2014). Optimization and mathematical explanation: Doing the Levy Walk. Synthese, 191(2014), 459–479.

    Article  Google Scholar 

  14. Bournez, O., & Pouly, A. (2018, 14 May). A survey of analogue models of computation. arXiv:1805.05729v1 [cs.CC].

  15. Chaitin, G., Da Costa, N. C. A., & Doria, F. A. (2012). Gödel’s way. Exploits into an undecidable world. Boca Raton: CRC Press, Taylor & Francis Group.

    Google Scholar 

  16. Copeland, J. (2002). Hypercomputation. Minds and Machines, 12, 461–502.

    Article  Google Scholar 

  17. Copeland, J. (2002). Accelerating Turing machines. Minds and Machines, 12, 281–301.

    Article  Google Scholar 

  18. Copeland, J. (2004). Hypercomputation: Philosophical issues. Theoretical Computer Science, 317, 251–267.

    Article  Google Scholar 

  19. Copeland, J. (2008). The modern history of computing. In E. N. Zalta (Ed.), The Stanford encyclopedia of philosophy (Fall 2008 Edition). http://plato.stanford.edu/archives/fall2008/entries/computing-history.

  20. Da Costa, N. C. A., & Doria, F. A. (1991). Undecidability and incompleteness in classical mechanics. International Journal Theoretical Physics, 30, 1041–1073.

    Article  Google Scholar 

  21. Cotogno, P. (2003). Hypercomputation and the physical Church-Turing thesis. British Journal for the Philosophy of Science, 54, 181–223.

    Article  Google Scholar 

  22. Cubitt, T. S., Perez-Garcia, D., & Wolf, M. W. (2015). Undecidability of the spectra gap. Nature, 528, 207–211 (full version: arXiv:1502.04573v3 [quant-ph]).

  23. Da Costa, N. C. A., & Doria, F. A. (1994). Suppes predicates an the construction of unsolvable problems in the axiomatized sciences. In P. Humpreys (Ed.), Patric Suppes: Scientific philosopher (pp. 151–193). Kluwer Academic Publishers.

    Google Scholar 

  24. Da Costa, N. C. A., & Doria, F. A. (1996). Structures, Suppes predicates, and Boolean-valued models in physics. In P. I. Bystrov & V. N. Sadovsky (Eds.), Philosophical logic and logical philosophy (pp. 91–118). Kluwer Academic Publishers.

    Google Scholar 

  25. Davis, M. (2006). Why there is no such discipline as hypercomputation. Applied Mathematics and Computation, 178, 4–7.

    Article  Google Scholar 

  26. Davis, M. (2004). The myth of hypercomputation. In: C. Teuscher (Ed.), Alan Turing: Life and legacy of a great thinker (pp. 195–212). Berlin: Springer.

    Google Scholar 

  27. Dawson, J. W., Jr. (2006). Why do mathematicians re-prove theorems. Philosophia Mathematica, III, 14, 269–286.

    Article  Google Scholar 

  28. Deutsch, D., Ekert, A., & Lupacchini, R. (2000). Machines, logic and quantum physics. The Bulletin of Symbolic Logic, 6(3), 265–283.

    Article  Google Scholar 

  29. Earman J., Norton J.D. (2003). Forever is a day: supertasks in Pitowsky and Malament-Hogarth spacetimes. Philosophy of Science, 60, 22–42.

    Google Scholar 

  30. Easton, W. B. (1970). Powers of regular cardinals. Annals of Mathematical Logic, 1, 139–178.

    Article  Google Scholar 

  31. Ellentuck, E. (1975). Gödel’s square axioms for the continuum. Mathematische Annalen, 216, 29–33.

    Google Scholar 

  32. Etesi, G., & Németi, I. (2002). Turing computability and Malament-Hogarth spacetimes. International Journal of Theoretical Physics, 41(2), 342–370.

    Article  Google Scholar 

  33. Feferman, S. (2000). Why the programs for new axioms need to be questioned. The Bulletin of Symbolic Logic, 6, 401–413.

    Article  Google Scholar 

  34. Friedman, H. (2000). Normal mathematics will need new axioms. The Bulletin of Symbolic Logic, 6, 434–446.

    Google Scholar 

  35. Friend, M. (2015). On the epistemological significance of the Hungarian project. Synthese, 192, 2035–2051.

    Article  Google Scholar 

  36. Friend, M., & Molinini, D. (2015). Using mathematics to explain a scientific theory. Philosophia Mathematica, 24(2), 185–213.

    Article  Google Scholar 

  37. Gödel, K. (1970). Some considerations leading to the probable conclusion, that the true power of the continuum is ℵ2. In: S. Feferman (Ed.), Kurt Gödel. Collected works (Vol. 3, pp. 420–421). Oxford: Oxford University Press.

    Google Scholar 

  38. Gödel, K. (1970). A proof of Cantor’s continuum hypothesis from a highly plausible axiom about orders of growth. In: S. Feferman (Ed.), Kurt Gödel. Collected works (Vol. 3, pp. 422–423). Oxford: Oxford University Press.

    Google Scholar 

  39. Hales, T. C. (2000). Cannonballs and honeycombs. Notices of the American Mathematical Society, 47(4), 440–449.

    Google Scholar 

  40. Hales, T. C. (2005). A proof of the Kepler conjecture. Annals of Mathematics. Second Series, 162(3), 1065–1185.

    Article  Google Scholar 

  41. Hamami, Y. (2018). Mathematical inference and logical inference. The Review of Symbolic Logic, 11(4), 665–704.

    Article  Google Scholar 

  42. Hamkins, J. D. (2002). Infinite time Turing machines. Minds and Machines, 12, 521–539.

    Article  Google Scholar 

  43. Hamkins, J. D. (2012). The set-theoretic multiverse. Review of Symbolic Logic, 5(3), 416–449.

    Article  Google Scholar 

  44. Hamkins, J. D., & Lewis, A. (2000). Infinite time Turing machines. Journal of Symbolic Logic, 65, 567–604.

    Article  Google Scholar 

  45. Hauser K. (2002). Is Cantor’s continuum problem inherently vague?. Philosophia Mathematica, 10, 257–292.

    Google Scholar 

  46. Hogarth, M. L. (1992). Does General Relativity allow an observer to view an eternity in a finite time? Foundations of Physics Letters, 5, 173–181.

    Article  Google Scholar 

  47. Hogarth, M. L. (1993). Predicting the future in relativistic spacetimes. Studies in History and Philosophy of Science. Studies in History and Philosophy of Modern Physics, 24, 721–739.

    Article  Google Scholar 

  48. Hogarth, M. L. (1994). Non-Turing computers and non-Turing computability. PSA, 1, 126–138.

    Google Scholar 

  49. Jackson, F., & Pettit, P. (1990). Program explanations: A general perspective. Analysis, 50(2), 107–117.

    Article  Google Scholar 

  50. Jones, J. P. (1980). Undecidable diophantine equations. Bulletin of the American Mathematical Society, 3(2), 859–862.

    Article  Google Scholar 

  51. Kieu, T. (2002). Quantum hypercomputation. Minds and Machines, 12, 541–561.

    Article  Google Scholar 

  52. Kieu, T. (2003). Quantum algorithm for Hilbert’s tenth problem. International Journal of Theoretical Physics, 42(7), 1461–1478.

    Article  Google Scholar 

  53. Kreisel, G. (1974). A notion of mechanistic theory. Synthese, 29, 11–26.

    Article  Google Scholar 

  54. Kreisel, G. (1982). Review of Pour-El and Richards. Journal of Symbolic Logic, 47, 900–902.

    Article  Google Scholar 

  55. Kreisel, G. (1965). Mathematical logic. In T. L. Saaty (Ed.), Lectures on modern mathematics (Vol. 3). New York: Wiley.

    Google Scholar 

  56. Kreisel, G. (1967). Mathematical logic: What has it done for the philosophy of mathematics? In R. Schoenman (Ed.), Bertrand Russell: Philosopher of the century. London: George Allen and Unwin.

    Google Scholar 

  57. Lange, M. (2013) What makes a scientific explanation distinctively mathematical?, British Journal for the Philosophy of Science, 64(3), 485–511.

    Google Scholar 

  58. Laraudogoitia, J. P. (2013). Supertasks. In E. N. Zalta (Ed.), The Stanford Encyclopedia of Philosophy (Fall 2013 Edition). https://plato.stanford.edu/archives/fall2013/entries/spacetime-supertasks/.

  59. Levy, A., & Solovay, R. M. (1967). Measurable cardinals and the continuum hypothesis. Israel Journal of Mathematics, 5, 234–248.

    Article  Google Scholar 

  60. Lyon, A., & Colyvan, M. (2008). The explanatory power of phase spaces. Philosophia Mathematica, 16(2), 227–243.

    Article  Google Scholar 

  61. Lyon, A. (2012). Mathematical explanations of empirical facts, and mathematical realism. Australasian Journal of Philosophy, 90(3), 559–578.

    Google Scholar 

  62. Maddy. (2000). Does mathematics need new axioms? The Bulletin of Symbolic Logic, 6, 413–422.

    Google Scholar 

  63. Maddy. (2011). Defending the axioms. Oxford: Oxford University Press.

    Google Scholar 

  64. Manchak, J., & Roberts, B. W. (2016). Supertasks. In E. N. Zalta (Ed.), The stanford encyclopedia of philosophy (Winter 2016 Edition). https://plato.stanford.edu/archives/win2016/entries/spacetime-supertasks.

  65. Mancosu, P. (2001). Mathematical explanation: Problems and prospects. Topoi, 20, 97–117.

    Article  Google Scholar 

  66. Mancosu, P. (2008). Mathematical explanation: Why it matters. In P. Mancosu (Ed.), Philosophy of mathematical practice (pp. 134–150). Oxford: Oxford University Press.

    Chapter  Google Scholar 

  67. Mancosu, P. (2018). Explanation in mathematics. Stanford encyclopedia of philosophy. https://plato.stanford.edu/archives/sum2018/entries/mathematics-explanation/.

  68. Montanaro, A. (2015). Quantum algorithms: An overview. https://www.nature.com/articles/npjqi201523; (also: arXiv:1511.04206v2).

  69. Mordell, L. (1959). Reflections of a mathematician. Montreal: Canadian Mathematical Congress.

    Google Scholar 

  70. Nielsen, M. A., & Chuang, I. L. (2000). Quantum computation and quantum information. Cambridge University Press.

    Google Scholar 

  71. Németi, I., & Dávid, G. (2006). Relativistic computers and the Turing barrier. Journal of Applied Mathematics and Computation, 178(1), 118–142.

    Article  Google Scholar 

  72. Németi, I. (1987). On logic, relativity, and the limitations of human knowledge. Iowa State University, Department of Mathematics, graduate course during the academic year 1987/88.

    Google Scholar 

  73. Németi, P., & Székely, G. (2012). Existence of faster than light signals implies hypercomputation already in special relativity. In S. B. Cooper, A. Dawar, & B. Löwe (Eds.), How the World Computes. CiE 2012 (Vol. 7318). Lecture Notes in Computer Science. Berlin, Heidelberg: Springer.

    Google Scholar 

  74. Piccinini, G. (2011). The physical Church-Turing thesis: Modest or bold? British Journal for the Philosophy of Science, 62, 733–769.

    Article  Google Scholar 

  75. Piccinini, G. (2017). Computation in physical systems. https://plato.stanford.edu/archives/sum2017/entries/computation-physicalsystems/.

  76. Pitowsky, I. (1990). The physical Church thesis and physical computational complexity. Iyyun, 39, 81–99.

    Google Scholar 

  77. Pour-El, M. B., & Richards, J. I. (1979). A computable ordinary differential equation which possesses no computable solution. Annals of Mathematical Logic, 17, 61–90.

    Article  Google Scholar 

  78. Pour-El, M. B., & Richards, J. I. (1981). The wave equation with computable initial data such that its unique solution is not computable. Advances in Mathematics, 39, 215–239.

    Article  Google Scholar 

  79. Pour-El, M. B., & Richards, J. I. (1989). Computability in analysis and physics. Berlin: Springer.

    Book  Google Scholar 

  80. Pour-El, M., & Zhong, N. (1997). The wave equation with computable initial data whose unique solution is nowhere computable. Mathematical Logic Quarterly, 43(4), 499–509.

    Article  Google Scholar 

  81. Quine, W. v. O. (1953). Two dogmas of empiricism. In From a logical point of view (pp. 20–46). Cambridge: Harvard University Press.

    Google Scholar 

  82. Rav, Y. (1999). Why do we prove theorems? Philosophia Mathematica, 7(1999), 5–41.

    Article  Google Scholar 

  83. Reutlinger, A., & Saatsi, J. (2018). Explanation beyond causation. Philosophical perspectives on non-causal explanation. Oxford: Oxford University Press.

    Book  Google Scholar 

  84. Rota, G.-C. (1997). The phenomenology of mathematical proof. Synthese, 111, 183–196.

    Article  Google Scholar 

  85. Scarpellini, B. (2003). Comments on: ‘Two undecidable problems of analysis’. Minds and Machines, 13, 79–85.

    Article  Google Scholar 

  86. Scarpellini, B. (1963). Zwei Unentscheitbare Probleme der Analysis. Zeitschrift für mathematische Logik und Grundlagen der Mathematik, 9, 265–289 (English, revised version: Minds and Machines, 2003, 13, 49–77).

    Google Scholar 

  87. Shagrir, O., & Pitowsky, I. (2003). Physical hypercomputation and the Church-Turing thesis. Minds and Machines, 13, 87–101.

    Article  Google Scholar 

  88. Shor, P. (1997). Polynomial-time algorithms for prime factorization and discrete logarithms on a quantum computer. SIAM Journal on Computing, 26, 1484–1509.

    Article  Google Scholar 

  89. Shor, P. (1994). Algorithms for quantum computation: Discrete logarithms and factoring. In: Proceedings 35th Annual Symposium on Foundations of Computer Science (pp. 124–134).

    Google Scholar 

  90. Siegelmann, H. T. (1998). Neural networks and analog computation: Beyond the Turing limit. Boston, MA: Birkhauser.

    Google Scholar 

  91. Siegelmann, H. T. (2003). Neural and super-Turing computing. Minds and Machines, 13, 103–114.

    Article  Google Scholar 

  92. Simpson, S. (2009). Subsystems of second order arithmetic. Cambridge: Cambridge University Press.

    Book  Google Scholar 

  93. Smith, W. D. (2006). Three counterexamples refuting Kieu’s plan for ‘‘quantum adiabatic hypercomputation’’; and some uncomputable quantum mechanical tasks. Applied Mathematics and Computation, 178, 184–193.

    Article  Google Scholar 

  94. Solovay, R. M. (1995). Introductory note to *1970a, *1970b, *1970c. In S. Feferman (Ed.), Kurt Gödel. Collected works (Vol. 3, pp. 405–420). Oxford: Oxford University Press.

    Google Scholar 

  95. Stannett, M. (2006). The case for hypercomputation. Applied Mathematics and Computation, 178, 8–24.

    Article  Google Scholar 

  96. Steel, J. R. (2000). Mathematics needs new axioms. The Bulletin of Symbolic Logic, 6, 422–433.

    Google Scholar 

  97. Syropoulos, A. (2008). Hypercomputation. Computing beyond the Church-Turing barrier. Springer: New York.

    Google Scholar 

  98. Szabó, L. E. (2017). Meaning, truth and physics. In G. Hofer-Szabó & L. Wroński (Eds.), Making it formally explicit (pp. 165–177). European Studies in Philosophy of Science 6. Springer International Publishing.

    Google Scholar 

  99. Székely, G. (2011). On why-questions in physics. In A. Máté, M. Rédei, & F. Stadler (Eds.), Der Wiener Kreis in Ungarn/The Vienna Circle in Hungary (Vol. 16, pp. 181–189). Veröffentlichungen des Instituts Wiener Kreis. Vienna: Springer.

    Google Scholar 

  100. Székely, G. (2012). What properties of numbers are needed to model accelerated observers in relativity? In J.-Y. Béziau, D. Krause, & J. R. Becker Arenhart (Eds.), Conceptual clarifications tributes to Patrick Suppes (1922–2014) (pp. 161–174). College Publications (also arXiv:1210.0101v1 [math.LO]).

  101. Tymoczko, T. (1979). The four-color problem and its philosophical significance. The Journal of Philosophy, 76(2), 57–83.

    Article  Google Scholar 

  102. Weihrauch, K., & Zhong, N. (2002). Is wave propagation computable or can wave computers beat the Turing machine? Proceedings of the London Mathematical Society, 85(2), 312–332.

    Article  Google Scholar 

  103. Welch, P. (2008). The Extent of Computation in Malament-Hogarth Spacetimes. British Journal for the Philosophy of Science, 59, 659–674 (arXiv:gr-qc/0609035v1).

  104. Woodin. (1999). The axiom of determinacy, forcing axioms and the nonstationary ideal. Berlin, New York, de Gruyter.

    Google Scholar 

  105. Woodin. (2001). The continuum hypothesis. Parts I and II. Notices of the AMS, 48(6–7), 567–576, 681–690.

    Google Scholar 

  106. Wójtowicz, K. (2009). Theory of quantum computation and philosophy of mathematics (I). Logic and Logical Philosophy, 18(3–4), 313–332.

    Google Scholar 

  107. Wójtowicz, K. (2014). The physical version of Church’s thesis and mathematical knowledge. In A. Olszewski, B. Brożek, & P. Urbańczyk (Eds.), Church’s thesis: Logic, mind nature (pp. 417–431). Kraków: Copernicus Center Press.

    Google Scholar 

  108. Wójtowicz, K. (2019). Theory of quantum computation and philosophy of mathematics (II). Logic and logical philosophy, 29(1), 173–193.

    Google Scholar 

  109. Wüthrich, C. (2015). A quantum-information-theoretic complement to a general-relativistic implementation of a beyond-Turing computer. Synthese, 192, 1989–2008.

    Article  Google Scholar 

Download references

Acknowledgements

I would like to express my gratitude to the Editors (especially to Gergely Székely) for many helpful comments and bibliographic hints during the preparation of the paper. I am also indebted to Istvan Németi and Hajnal Andréka for their friendly support and feedback. Finally I would like to thank the anonymous referee for insightful and stimulating comments. The preparation of this paper was supported by an National Science Center (NCN) grant: 2016/21/B/HS1/01955.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Krzysztof Wójtowicz .

Editor information

Editors and Affiliations

Rights and permissions

Reprints and permissions

Copyright information

© 2021 Springer Nature Switzerland AG

About this chapter

Check for updates. Verify currency and authenticity via CrossMark

Cite this chapter

Wójtowicz, K. (2021). The Significance of Relativistic Computation for the Philosophy of Mathematics. In: Madarász, J., Székely, G. (eds) Hajnal Andréka and István Németi on Unity of Science. Outstanding Contributions to Logic, vol 19. Springer, Cham. https://doi.org/10.1007/978-3-030-64187-0_7

Download citation

Publish with us

Policies and ethics