Indeterminism and Undecidability

  • Chapter
  • First Online:
Undecidability, Uncomputability, and Unpredictability

Part of the book series: The Frontiers Collection ((FRONTCOLL))

  • 882 Accesses

Abstract

The aim of this paper is to argue that the (alleged) indeterminism of quantum mechanics, claimed by adherents of the Copenhagen interpretation since Born [8], can be proved from Chaitin’s follow-up to Gödel’s (first) incompleteness theorem. In comparison, Bell’s [2] theorem as well as the so-called free will theorem-originally due to Heywood and Redhead [44]-left two loopholes for deterministic hidden variable theories, namely giving up either locality (more precisely: local contextuality, as in Bohmian mechanics) or free choice (i.e. uncorrelated measurement settings, as in ’t Hooft’s cellular automaton interpretation of quantum mechanics). The main point is that Bell and others did not exploit the full empirical content of quantum mechanics, which consists of long series of outcomes of repeated measurements (idealized as infinite binary sequences): their arguments only used the long-run relative frequencies derived from such series, and hence merely asked hidden variable theories to reproduce single-case Born probabilities defined by certain entangled bipartite states. If we idealize binary outcome strings of a fair quantum coin flip as infinite sequences, quantum mechanics predicts that these typically (i.e. almost surely) have a property called 1-randomness in logic, which is much stronger than uncomputability. This is the key to my claim, which is admittedly based on a stronger (yet compelling) notion of determinism than what is common in the literature on hidden variable theories.

Dedicated to the memory of Michael Redhead (1929–2020).

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

Access this chapter

Subscribe and save

Springer+ Basic
EUR 32.99 /Month
  • Get 10 units per month
  • Download Article/Chapter or Ebook
  • 1 Unit = 1 Article or 1 Chapter
  • Cancel anytime
Subscribe now

Buy Now

Chapter
EUR 29.95
Price includes VAT (Germany)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
EUR 67.40
Price includes VAT (Germany)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
EUR 85.59
Price includes VAT (Germany)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free ship** worldwide - see info
Hardcover Book
EUR 85.59
Price includes VAT (Germany)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free ship** worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

Similar content being viewed by others

Notes

  1. 1.

    In fact there are two incompleteness theorems in logic due to Gödel (see footnotes 2 and 5) and two theorems on quantum mechanics due to Bell [12, 89], but for reasons to follow in this essay I am mainly interested in the first ones, of both authors, except for a few side remarks.

  2. 2.

    See Franzén [35] for an excellent first introduction to Gödel’s theorems, combined with a fair and detailed critique of its abuses, including overstatements by both amateurs and experts (a similar guide to the use and abuse of Bell’s theorems remains to be written), and Smith [75] for a possible second go.

  3. 3.

    Yanofsky [91] nicely discusses both theorems in the context of the limits of science and reason.

  4. 4.

    Both reformulations are a bit anachronistic and purpose-made. See Gödel [37] and Bell [2]!

  5. 5.

    Gödel’s second incompleteness theorem shows that one example of \(\varphi \) is the (coded) statement that the consistency of the theory can be proved within the theory. This is often taken to refute Hilbert’s program, but even among experts it seems controversial if it really does so. For Hilbert’s program and its role in Gödel’s theorems see e.g. Zach [92], Tait [83], Sieg [74], and Tapp [84].

  6. 6.

    Greenstein [39] is a popular book on the history and interpretation of Bell’s work. Scholarly analyses include Redhead [68], Butterfield [14], Werner and Wolf [88], and the papers cited in footnote 49.

  7. 7.

    Some vocal researchers calim that Bell and Einstein were primarily interested in locality and realism, determinism being a secondary (or no) issue, but the historical record is ambiguous; more generally, over 10,000 papers about Bell’s theorems show that Bell can be interpreted in almost equally many ways. But this controversy is a moot point: whatever his own (or Einstein’s) intentions, Bell’s [2] theorem puts constraints on possible deterministic underpinnings of quantum mechanics, and that is how I take it.

  8. 8.

    For an overall survey of this theme see Kline [47].

  9. 9.

    This phase in the history of quantum mechanics is described by Mehra and Rechenberg [57].

  10. 10.

    See Landsman [53] for the view that randomness is a family resemblance (in that it lacks a meaning common to all its applications) with the special feature that its various uses are always defined antipodally.

  11. 11.

    Also cf. e.g. Breuer [9], Calude [16], Svozil, Calude and Stay [79], and Szangolies [80].

  12. 12.

    A subset \(E\subset {\mathbb N}\) is computably enumerable (c.e.) if it is the image of a computable function \(f:{\mathbb N}\rightarrow {\mathbb N}\), and computable if its characteristic function \(1_E\) is computable, which is true iff both E and \({\mathbb N}\backslash E\) are c.e.

  13. 13.

    This is true for physics almost as much as it is (more famously) for mathematics, since Hilbert played a major role in the mathematization of the two great theories of twentieth century physics, i.e. general relativity [22, 69] and quantum mechanics [54, 67]).

  14. 14.

    This is one of innumerable paradoxes of natural language, which leads to an incompleteness theorem once the notion of a description has been appropriately formalized in mathematics, much as Gödel’s first incompleteness theorem turns the liar’s paradox into a theorem.

  15. 15.

    See also Bub [13], Dieks [26], and forthcoming work by Chris Mitsch for balanced accounts.

  16. 16.

    The idea of contextuality was first formulated by Grete Hermann [23].

  17. 17.

    See Kochen and Specker [48]. Ironically, his followers attribute this theorem to Bell [3], although the result is just a technical sharpening of von Neumann’s result they so vehemently ridicule. For a deep philosophical analysis of the Kochen–Specker theorem, as well as of Bell’s theorems, see Redhead [68].

  18. 18.

    See Appendix C. This theorem is originally due to Heywood and Redhead [44], with follow-ups by Stairs [76], Brown and Svetlichny [11], and Clifton [18], but it was named and made famous by Conway and Kochen [21], whose main contribution was an emphasis on free will ([52], Chap. 6).

  19. 19.

    Since Alice and Bob are spacelike separated their observables commute (Einstein locality).

  20. 20.

    In quantum mechanics the left-hand side of (3.3.1) satisfies this locality condition for any state \(\psi \).

  21. 21.

    See [52], §6.5 for details, or Appendix C below for a summary.

  22. 22.

    There is a subtle difference between Bohmian mechanics as reviewed by e.g. Goldstein [38], and de Broglie’s original pilot wave theory [86]. This difference is immaterial for my discussion.

  23. 23.

    In Bohmian mechanics, the hidden state \(q\in Q\) just pertains to the particles undergoing measurement, whilst the settings a are supposed to be “freely chosen” for each measurement (and in particular are independent of q). The outcome is then fixed by a and q. In ’t Hooft’s theory, the hidden state \(x\in X\) of “the world” determines the settings as well as the outcomes. Beyond the issue raised in the main text, Bohmians (but not ’t Hooft!) therefore have an additional problem, namely the origin of the settings (which are simply left out of the theory). This weakens their case for determinism even further.

  24. 24.

    The function g incorporates all details of the experiment that may affect the outcome (like the setting, context, and quantum state) except the hidden variable \(\lambda \) (which it specifies). It has nothing to do with noncontextual value assignments on the set of quantum-mechanical observables (which do not exist).

  25. 25.

    The Bohmians are divided on the origin of their compatibility measure, referred to in this context as the quantum equilibrium distribution, cf. Dürr, Goldstein, and Zanghi [29] against Valentini [86]. The origin of \(\mu _{\psi }\) is not my concern, which is the need to randomly sample it and the justification for doing so.

  26. 26.

    In stating the second condition I have taken \(\sigma (a)=\{0,1\}\) with 50-50 Born probabilities, but this can be generalized to other spectra and probability measures. See Downey and Hirschfeldt [27], §6.12.

  27. 27.

    In other words, we examine whether Earman’s principle is satisfied, cf. footnote 37.

  28. 28.

    To make this argument completely rigorous one would need to define what a “description” provided by a deterministic theory means logically. There is a logical characterization of deterministic theories [59], and there are some arguments to the effect that the evolution laws in deterministic theories should be computable, cf. Earman [31], Chap. 11, and Pour-El and Richards [65], passim, but this literature makes no direct reference to output strings or sequences of the kind we analyze and in any case the identification of “deterministic” with “computable” is obscure even in situations where the latter concept is well defined. For example, if we stipulate that \(h:{\mathbb N}\rightarrow \Lambda \) is computable (and likewise \(g:\Lambda \rightarrow \underline{2}\)) then the above appeal to Chaitin’s first incompleteness theorem is not even necessary, but this seems too easy. A somewhat circular solution, proposed by Scriven [70], is to simply say that T is deterministic iff the output strings or sequences it describes are not random, but this begs for a more explicit characterization. One might naively expect such a characterization to come from the arithmetical hierarchy (found in any book on computability): if, as before, we identify \(\underline{2}^{{\mathbb N}}\) with the power set \(P({\mathbb N})\) of \({\mathbb N}\), then \(S\subset {\mathbb N}\) is called arithmetical if there is a formula \(\psi (x)\) in PA (Peano Arithmetic) such that \(n\in S\) iff \({\mathbb N}\vDash \psi (n)\), that is, \(\psi (n)\) is true in the usual sense. We may then classify the arithmetical subsets through the logical form of \(\psi \), assumed in prenex normal form (i.e., all quantifiers have been moved to the left): S is in \(\Sigma _0^0=\Pi _0^0\) iff \(\psi \) has no quantifiers or only bounded quantifiers (in which case S is computable), and then recursively \(S\in \Sigma ^0_{n+1}\) iff \(\psi (x)=\exists _y\varphi (x,y)\) with \(\varphi \in \Pi _n^0\), and \(\varphi \in \Pi _{n+1}^0\) iff \(\psi (x)=\forall _y\varphi (x,y)\) with \(\varphi \in \Sigma _n^0\). Here any singly quantified expression \(\exists _y\varphi (x,y)\) may be replaced by \(\exists {y_1}\cdots \exists {y_k}\varphi (x,y_1, \ldots , y_k)\) and likewise for \(\forall _y\). By convention \(\Sigma ^0_n\subset \Sigma ^0_{n+1}\) and \(\Pi ^0_n\subset \Pi ^0_{n+1}\), and \(\Delta ^0_n:=\Sigma _n^0\cap \Pi _n^0\). Since in classical logic \(\forall _y\varphi (x,y)\) is equivalent to \(\lnot \exists _y\lnot \varphi (x,y)\), it follows that \(\Pi _n^0\) sets are the complements of \(\Sigma _n^0\) sets. One would then like to locate deterministic theories somewhere in this hierarchy, preferably above the computable \(\Delta _0^0\). The idea of a hidden variable (namely y) suggests \(\Sigma _1^0\) and closure under complementation (it would be crazy if some deterministic theory prefers ones over zeros) then leads to \(\Delta _1^0\), but this equals \(\Delta _0^0\). The next level \(\Delta _2^0\) is impossible since this already contains 1-random sets like Chaitin’s \(\Omega \). Hence more research is needed.

  29. 29.

    For Bell’s proof it is irrelevant whether or not some hidden variable is able to sample the compatibility measure, since the Bell inequalities follow from pointwise bounds, cf. Landsman [52], Eq. (6.119).

  30. 30.

    Here a state \(\omega \) is a positive normalized linear functional on B(H), as in the C*-algebraic approach to quantum mechanics (Haag [41], Landsman [52]). One may think of expectation values \(\omega (a)=\text{ Tr }\,(\rho a)\), where \(\rho \) is a density operator on H, with the special case \(\omega (a)=\langle \psi , a\psi \rangle \), where \(\psi \in H\) is a unit vector. For a proof of Theorem A.1 see Landsman [52], Sect. 4.1, Corollary 4.4.

  31. 31.

    This can even be replaced by a single measurement, see Landsman [52], Corollary A.20.

  32. 32.

    The uses of states themselves may be justified by Gleason’s theorem ([52], §§2,7, 4.4).

  33. 33.

    The Born rule for commuting operators follows from the single operator case ([52], §2.5).

  34. 34.

    If B is infinite-dimensional, for technical reasons the so-called projective tensor product should be used.

  35. 35.

    See Landsman [52], §8.4 for this approach. The details are unnecessary here.

  36. 36.

    Cf. Tychonoff’s theorem. The associated Borel structure is the one defined by the cylinder sets.

  37. 37.

    For details see Volchan [87], Terwijn [85], Diaconis and Skyrms ([25], Chap. 8), and Eagle [30] for starters, technical surveys by Zvonkin and Levin [93], Muchnik et al. [61], Downey et al. [28], Grünwald and Vitányi [40], and Dasgupta [24], and books by Calude [15], Li and Vitányi [56], Nies [63], and Downey and Hirschfeldt [27]. For history see van Lambalgen [50, 51] and Li and Vitányi [56]. For physical applications see e.g. Earman [31], Svozil [77, 78], Calude [16], Wolf [90], Bendersky et al. [4, 5], Senno [73], Baumeler et al. [1], and Tadaki [81, 82].

  38. 38.

    A Turing machine T is prefix-free if its domain D(T) consists of a prefix-free subset of \(\underline{2}^*\), i.e., if \(\sigma \in D(T)\) then \(\sigma \tau \notin D(T)\) for any \(\sigma ,\tau \in \underline{2}^*\), where \(\sigma \tau \) is the concatenation of \(\sigma \) and \(\tau \): if T halts on input \(\sigma \) then it does not halt on either any initial part or any extension of \(\sigma \). The prefix-free version is only needed to correctly define randomness of sequences in terms of randomness of their initial parts, which is necessary to satisfy Earman’s Principle: ‘While idealizations are useful and, perhaps, even essential to progress in physics, a sound principle of interpretation would seem to be that no effect can be counted as a genuine physical effect if it disappears when the idealizations are removed.’ See Earman [32], p. 191. For finite strings \(\sigma \) one may work with the plain Kolmogorov complexity \(C(\sigma )\), defined as the length (in bits) of the shortest computer program (run on some fixed universal Turing machine U) that computes \(\sigma \).

  39. 39.

    Recall that \(f(n)=O(g(n))\) iff there are constants C and N such that \(|f(n)|\le C|g(n)|\) for all \(n\ge N\).

  40. 40.

    It is easy to show that least \(2^N-2^{N-c+1}+1\) strings \(\sigma \) of length \(|\sigma |=N\) are c-Kolmogorov random.

  41. 41.

    Here “sound” means that all theorems proved by T are true; this is a stronger assumption than consistency (in fact only the arithmetic fragment of T needs to be sound). One may think of Peano Arithmetic (PA) or of Zermelo–Fraenkel set theory with the axiom of choice (ZFC). As in Gödel’s theorems, one also assumes that T is formalized as an axiomatic-deductive system in which proofs could in principle be carried out mechanically by a computer. The status of the true but unprovable sentences \(K(\sigma )>C\) in Chaitin’s theorem is similar to that of the sentence G in Gödel’s original proof of his first incompleteness theorem, which roughly speaking is an arithmetization of the statement “I cannot be proved in T”: assuming soundness and hence consistency of T, one can prove \(\mathsf {G}\) and \(K(\sigma )>C\) in the usual interpretation of the arithmetic fragment of T in the natural numbers \({\mathbb N}\). See Chaitin [20] for his own presentation and analysis of his incompleteness theorem. Raatikainen [66] also gives a detailed presentation of the theorem, including a critique of Chaitin’s ideology. Incidentally, he shows that there even exists a U with respect to which \(K(\cdot )\) is defined such that \(C=0\) in ZFC. See also Franzén [35] and Gács [36].

  42. 42.

    The following contradiction can be made more dramatic by taking n such that \(n>>|P|+|n|\).

  43. 43.

    See Calude [15], Theorem 6.38 (attributed to Chaitin) for this equivalence.

  44. 44.

    Any pattern in a sequence x would make it compressible, but one has to define the notion of a pattern very carefully in a computational setting. This was accomplished by Martin-Löf in 1966, who defined a pattern as a specific kind of probability-zero subset T of \(\underline{2}^{{\mathbb N}}\) (called a “test”) that can be computably approximated by subsets \(T_n\subset \underline{2}^{{\mathbb N}}\) of increasingly small probability \(2^{-n}\); if \(x\in T\), then x displays some pattern and it is patternless iff \(x\notin T\) for all such tests. Martin-Löf’s definition yields what usually called 1-randomness, in view of his use of so-called \(\Sigma ^0_1\) sets. See the textbooks Li and Vitányi [56], Calude [15], Nies [63], and Downey and Hirschfeldt [27] for the equivalences between Levin–Chaitin randomness (incompressibility), Martin-Löf randomness (patternlessness), and a third notion (unpredictability) that evolved from the work of von Mieses and Ville, finalized by Schnorr. The name Levin–Chaitin randomness, taken from Downey et al. [28], is justified by its independent origin in Levin [55] and Chaitin [19].

  45. 45.

    For details and proofs see Calude [15], Corollary 6.32 in §6.3 and almost all of §6.4.

  46. 46.

    To see this, use the measure-theoretic isomorphism between \((\underline{2}^{{\mathbb N}},\Sigma _K,P^{\infty })\) and \(([0,1],\Sigma _L,dx)\), where \(\Sigma _K\) is the “Kolmogorov” \(\sigma \)-algebra generated by the cylinder sets \([\sigma ]=\{x\in \underline{2}^{{\mathbb N}}\mid x_{||\sigma |}=\sigma \}\), where \(\sigma \in \underline{2}^*\), and \(\Sigma _K\) is the “Lebesgue” \(\sigma \)-algebra generated by the open subsets of [0, 1]. See also Nies [63], §1.8.

  47. 47.

    See Calude [15], Theorem 6.63. Hence meagre subsets of [0, 1] exist with unit Lebesgue measure!

  48. 48.

    More precisely, only finitely many true statements of the form: ‘the n’th bit \(x_n\) of x equals its actual value’ (i.e. 0 or 1) are provable in T (where a proof in T may be seen as a computation, since one may algorithmically search for this proof in a list). See Calude [15], Theorem 8.7, which is stated for Chaitin’s \(\Omega \) but whose proof holds for any 1-random sequence. Indeed, as pointed out to the author by Bas Terwijn, even more generally, ZFC (etc.) can only compute finitely many digits of any immune sequence (we say that a sequence \(x\in \underline{2}^{{\mathbb N}}\) is immune if the corresponding subset \(S\subset {\mathbb N}\) (i.e. \(1_S=x\)) contains no infinite c.e. subset), and by (for example) Corollary 6.42 in Calude [15] any 1-random sequence is immune.

  49. 49.

    There exists a U for which not a single digit of \(\Omega _U\) can be known, see Calude [15], Theorem 8.11.

  50. 50.

    The original reference for Bell’s theorem is Bell [2]; see further footnote 6, and in the context of this appendix also Esfeld [34] and Sen and Valentini [72] are relevant. The free will theorem originates in Heywood and Redhead [44], followed by Stairs [76], Brown and Svetlichny [11], Clifton [18], and, as name-givers, Conway and Kochen [21]. Both theorems can and have been presented and interpreted in many different ways, of which we choose the one that is relevant for the general discussion on randomness in the main body of the paper. This appendix is taken almost verbatim from Landsman [53].

  51. 51.

    This addresses a problem Bell faced even according to some of his most ardent supporters [64, 71], namely the tension between the idea that the hidden variables (in the pertinent causal past) should on the one hand include all ontological information relevant to the experiment, but on the other hand should leave Alice and Bob free to choose any settings they like. Whatever its ultimate fate, ’t Hooft’s staunch determinism has drawn attention to issues like this, as has the free will theorem.

  52. 52.

    If her setting is a basis \((\vec {e}_1,\vec {e}_2,\vec {e}_3)\), Alice measures the quantities \((J_{\vec {e}_1}^2, J_{\vec {e}_2}^2, J_{\vec {e}_3}^2)\), where \(J_{\vec {e}_1}=\langle \vec {J},\vec {e}_i\rangle \) is the component of the angular momentum operator \(\vec {J}\) of a massive spin-1 particle in the direction \(\vec {e}_i\).

  53. 53.

    Here \(S^2=\{(x,y,z)\in {\mathbb R}^3\mid x^2+y^2+z^2=1\}\) is the 2-sphere, seen as the space of unit vectors in \({\mathbb R}^3\). Equation (3.5.36) means that the outcome of Alice’s measurement of \(J_{\vec {e}_i}^2\) is independent of the “context” \((J_{\vec {e}_1}^2, J_{\vec {e}_2}^2, J_{\vec {e}_3}^2)\); she might as well measure \(J_{\vec {e}_i}^2\) by itself. The last equation is trivial, since \((J_{-\vec {e}_i})^2=(J_{\vec {e}_i})^2\).

  54. 54.

    The assumptions imply the existence of a coloring \(C_{\lambda }: \mathcal {P}\rightarrow \{0,1\}\) of \({\mathbb R}^3\), where \({\mathcal P}\subset S^2\) consist of all unit vectors contained in all bases in S, and \(\lambda \) “goes along for a free ride”. A coloring of \({\mathbb R}^3\) is a function \(C:\mathcal {P}\rightarrow \{0,1\}\) such that for any set \(\{e_1,e_2,e_3\}\) in \(\mathcal {P}\) with \(e_ie_j=\delta _{ij} 1_3\) and \(e_1+e_2+e_3=1_3\) where \(1_3\) is the \(3\times 3\) unit matrix) there is exactly one \(e_i\) for which \(C(e_i)=1\). Indeed, one finds \(C_{\lambda }(\vec {e})=\tilde{L}(\vec {e},\lambda )\). The key to the proof of Kochen–Specker is that on a suitable choice of the set S such a coloring cannot exist.

  55. 55.

    The existence of \(\mu \) is of course predicated on X being a measure space with corresponding \(\sigma \)-algebra of measurable subsets, with respect to which all functions in (3.5.38) and below are measurable.

  56. 56.

    In Bell’s theorem quantum theory can be replaced by experimental support [43].

  57. 57.

    As in Kochen–Specker, this is because Alice and Bob measure squares of (spin-1) angular momenta.

  58. 58.

    By definition, this also implies that the pairs (AB), (AH), and (BH) are also independent.

References

  1. A. Baumeler, C.A. Bédard, G. Brassard, S. Wolf, Kolmogorov amplification from Bell correlation, in IEEE International Symposium on Information Theory (ISIT) (2017). https://doi.org/10.1109/ISIT.2017.8006790

  2. J.S. Bell, On the Einstein–Podolsky–Rosen paradox. Physics 1, 195–200 (1964)

    Article  MathSciNet  Google Scholar 

  3. J.S. Bell, On the problem of hidden variables in quantum mechanics. Rev. Mod. Phys. 38, 447–452 (1966)

    Article  ADS  MathSciNet  Google Scholar 

  4. A. Bendersky, G. de la Torre, G. Senno, S. Figueira, A. Acín, Algorithmic pseudorandomness in quantum setups. Phys. Rev. Lett. 116, 230402 (2016)

    Article  ADS  Google Scholar 

  5. A. Bendersky, G. Senno, G. de la Torre, S. Figueira, A. Acín, Nonsignaling deterministic models for nonlocal correlations have to be incomputable. Phys. Rev. Lett. 118, 130401 (2017)

    Article  ADS  MathSciNet  Google Scholar 

  6. D. Bohm, A suggested interpretation of the quantum theory in terms of ‘hidden’ variables, I, II. Phys. Rev. 85, 166–179, 180–193 (1952)

    Google Scholar 

  7. N. Bohr, Can quantum-mechanical description of physical reality be considered complete? Phys. Rev. 48, 696–702 (1935)

    Article  ADS  Google Scholar 

  8. M. Born, Zur Quantenmechanik der Stoßvorgänge. Z. für Phys. 37, 863–867 (1926)

    Article  ADS  Google Scholar 

  9. T. Breuer, Von Neumann, Gödel and quantum incompleteness, in John von Neumann and the Foundations of Quantum Physics, ed. by M. Rédei, M. Stöltzner (Kluwer, Dordrecht, 2001), pp. 75–82

    Google Scholar 

  10. L. de Broglie, La nouvelle dynamique des quanta, Electrons et Photons: Rapports et Discussions du Cinquième Conseil de Physique (Gauthier-Villars, Paris, 1928), pp. 105–132

    Google Scholar 

  11. H. Brown, G. Svetlichny, Nonlocality and Gleason’s lemma. Part I. Deterministic theories. Found. Phys. 20, 1379–1387 (1990)

    Article  ADS  MathSciNet  Google Scholar 

  12. H. Brown, C.G. Timpson, Bell on Bell’s theorem: the changing face of nonlocality (2014), ar**v:1501.03521

  13. J. Bub, Is von Neumann’s ‘no hidden variables’ proof silly? in Deep Beauty: Mathematical Innovation and the Search for Underlying Intelligibility in the Quantum World, ed. by H. Halvorson (Cambridge University Press, Cambridge, 2011), pp. 393–408

    Google Scholar 

  14. J. Butterfield, Bell’s theorem: what it takes. Br. J. Philos. Sci. 43, 41–83 (1992)

    Article  MathSciNet  Google Scholar 

  15. C.S. Calude, Information and Randomness: an Algorithmic Perspective, 2nd edn. (Springer, Berlin, 2002)

    Google Scholar 

  16. C.S. Calude, Algorithmic randomness, quantum physics, and incompleteness. Lect. Notes Comput. Sci. 3354, 1–17 (2004)

    MathSciNet  MATH  Google Scholar 

  17. E. Cator, N.P. Landsman, Constraints on determinism: Bell versus Conway–Kochen. Found. Phys. 44, 781–791 (2014)

    Article  ADS  MathSciNet  Google Scholar 

  18. R. Clifton, Getting contextual and nonlocal elements-of-reality the easy way. Am. J. Phys. 61, 443–447 (1993)

    Article  ADS  Google Scholar 

  19. G.J. Chaitin, A theory of program size formally identical to information theory. J. Assoc. Comput. Mach. 22, 329–340 (1975)

    Article  MathSciNet  Google Scholar 

  20. G.J. Chaitin, Information, Randomness and Incompleteness: Papers on Algorithmic Information Theory (World Scientific, Singapore, 1987)

    Google Scholar 

  21. J.H. Conway, S. Kochen, The strong free will theorem. Not. Am. Math. Soc. 56, 226–232 (2009)

    MathSciNet  MATH  Google Scholar 

  22. L. Corry, David Hilbert and the Axiomatization of Physics (1898–1918): from Grundlagen der Geometrie to Grundlagen der Physik (Kluwer, Dordrecht, 2004)

    Google Scholar 

  23. E. Crull, G. Bacciagaluppi, Grete Hermann-Between Physics and Philosophy (Springer, Berlin, 2016)

    Google Scholar 

  24. A. Dasgupta, Mathematical foundations of randomness, in Handbook of the Philosophy of Science. Volume 7: Philosophy of Statistics, ed. by P.S. Bandyopadhyay, M.R. Forster (North-Holland/Elsevier, Amsterdam, 2011), pp. 641–710

    Google Scholar 

  25. P. Diaconis, B. Skyrms, Ten Great Ideas About Chance (Princeton University Press, Princeton, 2018)

    Google Scholar 

  26. D. Dieks, Von Neumann’s impossibility proof: mathematics in the service of rhetorics. Stud. Hist. Philos. Mod. Phys. 60, 136–148 (2016)

    Article  MathSciNet  Google Scholar 

  27. R. Downey, D.R. Hirschfeldt, Algorithmic Randomness and Complexity (Springer, Berlin, 2010)

    Google Scholar 

  28. R. Downey, D.R. Hirschfeldt, A. Nies, S.A. Terwijn, Calibrating randomness. Bull. Symb. Log. 12, 411–491 (2006)

    Article  MathSciNet  Google Scholar 

  29. D. Dürr, S. Goldstein, N. Zanghi, Quantum equilibrium and the origin of absolute uncertainty. J. Stat. Phys. 67, 843–907 (1992)

    Article  ADS  MathSciNet  Google Scholar 

  30. A. Eagle, Chance versus randomness, The Stanford Encyclopedia of Philosophy (2019), https://plato.stanford.edu/archives/spr2019/entries/chance-randomness/

  31. J. Earman, A Primer on Determinism (D. Reidel, Dordrecht, 1986)

    Book  Google Scholar 

  32. J. Earman, Curie’s principle and spontaneous symmetry breaking. Int. Stud. Philos. Sci. 18, 173–198 (2004)

    Article  MathSciNet  Google Scholar 

  33. A. Einstein, B. Podolsky, N. Rosen, Can quantum-mechanical description of physical reality be considered complete? Phys. Rev. 47, 777–780 (1935)

    Article  ADS  Google Scholar 

  34. M. Esfeld, Bell’s theorem and the issue of determinism and indeterminism. Found. Phys. 45, 471–482 (2015)

    Article  ADS  MathSciNet  Google Scholar 

  35. T. Franzén, Gödel’s Theorem: an Incomplete Guide to Its Use and Abuse (AK Peters, Natick, 2005)

    Google Scholar 

  36. P. Gács, Review of “algorithmic information theory” by Gregory J. Chaitin. J. Symb. Log. 54, 624–627 (1989)

    Google Scholar 

  37. K. Gödel, Über formal unentscheidbare Sätze der Principia Mathematica und verwandter Systeme I. Monatshefte für Mathematik und Physik 38, 173–198 (1931)

    Article  MathSciNet  Google Scholar 

  38. S. Goldstein, Bohmian mechanics, The Stanford Encyclopedia of Philosophy (2017), https://plato.stanford.edu/archives/sum2017/entries/qm-bohm/

  39. G. Greenstein, Quantum Strangeness: Wrestling with Bell’s Theorem and the Ultimate Nature of Reality (The MIT Press, Cambridge, 2019)

    Google Scholar 

  40. P.D. Grünwald, P.M.B. Vitányi, Algorithmic information theory, inHandbook of the Philosophy of Information, ed. by P. Adriaans, J. van Benthem (Elsevier, Amsterdam, 2008), pp. 281–320

    Google Scholar 

  41. R. Haag, Local Quantum Physics: Fields, Particles, Algebras (Springer, Berlin, 1992)

    Google Scholar 

  42. I. Hacking, The Taming of Chance (Cambridge University Press, Cambridge, 1990)

    Google Scholar 

  43. B. Hensen et al., Experimental loophole-free violation of a Bell inequality using entangled electron spins separated by 1.3 km. Nature 526, 682–686 (2015)

    Article  ADS  Google Scholar 

  44. P. Heywood, M. Redhead, Nonlocality and the Kochen–Specker paradox. Found. Phys. 13, 481–499 (1983)

    Article  ADS  MathSciNet  Google Scholar 

  45. D. Hilbert, Mathematical problems. Lecture delivered before the international congress of mathematicians at Paris in 1900. Bull. Am. Math. Soc. 8, 437–479. Translated from Göttinger Nachrichten 1900, 253–297 (1902)

    Google Scholar 

  46. G. ’t Hooft, The Cellular Automaton Interpretation of Quantum Mechanics (Springer Open, 2016), https://www.springer.com/gp/book/9783319412849

  47. M. Kline, Mathematics: The Loss of Certainty (Oxford University Press, Oxford, 1980)

    Google Scholar 

  48. S. Kochen, E. Specker, The problem of hidden variables in quantum mechanics. J. Math. Mech. 17, 59–87 (1967)

    MathSciNet  MATH  Google Scholar 

  49. A.N. Kolmogorov, Grundbegriffe de Wahrscheinlichkeitsrechnung (Springer, Berlin, 1933)

    Google Scholar 

  50. M. van Lambalgen, Random sequences. Ph.D. thesis, University of Amsterdam (1987), https://www.academia.edu/23899015/RANDOM_SEQUENCES

  51. M. van Lambalgen, Randomness and foundations of probability: Von Mises’ axiomatisation of random sequences, Statistics, Probability and Game Theory: Papers in Honour of David Blackwell. IMS Lecture Notes–Monograph Series, vol. 30 (1996), pp. 347–367

    Google Scholar 

  52. K. Landsman, Foundations of Quantum Theory: from Classical Concepts to Operator Algebras (Springer Open, 2017), https://www.springer.com/gp/book/9783319517766

  53. K. Landsman, Randomness? What randomness? Found. Phys. 50, 61–104 (2020)

    Article  ADS  MathSciNet  Google Scholar 

  54. K. Landsman, Quantum theory and functional analysis, in Oxford Handbook of the History of Interpretations and Foundations of Quantum Mechanics, ed. by O. Freire, to appear (Oxford University Press, Oxford, 2021)

    Google Scholar 

  55. L.A. Levin, On the notion of a random sequence. Sov. Math.-Dokl. 14, 1413–1416 (1973)

    MATH  Google Scholar 

  56. M. Li, P.M.B. Vitányi, An Introduction to Kolmogorov Complexity and Its Applications, 3rd edn. (Springer, Berlin, 2008)

    Google Scholar 

  57. J. Mehra, H. Rechenberg, The Historical Development of Quantum Theory. Vol. 6: the Completion of Quantum Mechanics 1926–1941. Part 1 (Springer, Berlin, 2000)

    Google Scholar 

  58. R. von Mises, Grundlagen der Wahrscheinlichkeitsrechnung. Math. Z. 5, 52–99 (1919)

    Article  MathSciNet  Google Scholar 

  59. R. Montague, Deterministic theories, in Formal Philosophy: Selected Papers of Richard Montague, ed. by R.H. Thomason (Yale University Press, London, 1974), pp. 303–360

    Google Scholar 

  60. J.R. Moschovakis, Iterated definability, lawless sequences and Brouwer’s continuum, in Gödel’s Disjunction, ed. by L. Horsten, P. Welch (Oxford University Press, Oxford, 2016), pp. 92–107

    Google Scholar 

  61. A.A. Muchnik, A.L. Semenov, V.A. Uspensky, Mathematical metaphysics of randomness. Theor. Comput. Sci. 207, 263–317 (1998)

    Article  MathSciNet  Google Scholar 

  62. J. von Neumann, Mathematische Grundlagen der Quantenmechanik (Springer). English translation: Mathematical Foundations of Quantum Mechanics (Princeton University Press, Princeton, 1955) (1932)

    Google Scholar 

  63. A. Nies, Computability and Randomness (Oxford University Press, Oxford, 2009)

    Google Scholar 

  64. T. Norsen, Local causality and completeness: Bell vs Jarrett. Found. Phys. 39, 273–294 (2009)

    Article  ADS  MathSciNet  Google Scholar 

  65. M.B. Pour-El, J.I. Richards, Computability in Analysis and Physics (Cambridge University Press, Cambridge, 2016)

    Google Scholar 

  66. P. Raatikainen, On interpreting Chaitin’s incompleteness theorem. J. Philos. Log. 27, 569–586 (1998)

    Article  MathSciNet  Google Scholar 

  67. M. Rédei, M. Stöltzner (eds.), John von Neumann and the Foundations of Quantum Physics (Kluwer, Dordrecht, 2001)

    Google Scholar 

  68. M. Redhead, Incompleteness, Nonlocality, and Realism (Oxford University Press, Oxford, 1989)

    Google Scholar 

  69. J. Renn (ed.), The Genesis of General Relativity, vol. 4 (Springer, Berlin, 2007)

    Google Scholar 

  70. M. Scriven, The present status of determinism in physics. J. Philos. 54, 727–741 (1957)

    Article  Google Scholar 

  71. M.P. Seevinck, J. Uffink, Not throwing out the baby with the bathwater: Bell’s condition of local causality mathematically ‘sharp and clean’. Explan. Predict. Confirmation 2, 425–450 (2011)

    Article  Google Scholar 

  72. I. Sen, A. Valentini, Superdeterministic hidden-variables models I: nonequilibrium and signalling (2020), ar**v:2003.11989, II: conspiracy, ar**v:2003.12195

  73. G. Senno, A computer-theoretic outlook on foundations of quantum information. Ph.D. thesis, Universidad de Buenos Aires (2017)

    Google Scholar 

  74. W. Sieg, Hilbert’s Programs and Beyond (Oxford University Press, Oxford, 2013)

    Google Scholar 

  75. P. Smith, An Introduction to Gödel’s Theorems (Cambridge University Press, Cambridge, 2013)

    Google Scholar 

  76. A. Stairs, Quantum logic, realism, and value definiteness. Philos. Sci. 50, 578–602 (1983)

    Article  MathSciNet  Google Scholar 

  77. K. Svozil, Randomness and Undecidability in Physics (World Scientific, Singapore, 1993)

    Google Scholar 

  78. K. Svozil, Physical (A)Causality: Determinism, Randomness and Uncaused Events (Springer Open, 2018), https://www.springer.com/gp/book/9783319708140

  79. K. Svozil, C.S. Calude, M.A. Stay, From Heisenberg to Gödel via Chaitin. Int. J. Theor. Phys. 44, 1053–1065 (2005)

    Article  Google Scholar 

  80. J. Szangolies, Epistemic horizons and the foundations of quantum mechanics (2018), ar**v:1805.10668

  81. K. Tadaki, A refinement of quantum mechanics by algorithmic randomness (2018), ar**v:1804.10174

  82. K. Tadaki, A Statistical Mechanical Interpretation of Algorithmic Information Theory (Springer, Berlin, 2019)

    Google Scholar 

  83. W. Tait, The Provenance of Pure Reason: Essays in the Philosophy of Mathematics and Its History (Oxford University Press, Oxford, 2005)

    Google Scholar 

  84. C. Tapp, An den Grenzen des Endlichen: Das Hilbertprogramm (Springer, Berlin, 2013)

    Google Scholar 

  85. S.A. Terwijn, The mathematical foundations of randomness, in The Challenge of Chance: a Multidisciplinary Approach from Science and the Humanities, ed. by K. Landsman, E. van Wolde (Springer, Berlin, 2016), pp. 49–66, https://www.springer.com/gp/book/9783319262987

  86. A. Valentini, Foundations of statistical mechanics and the status of the Born rule in de Broglie-Bohm pilot-wave theory (2019), ar**v:1906.10761

  87. S.B. Volchan, What is a random sequence? Am. Math. Mon. 109, 46–63 (2002)

    Article  MathSciNet  Google Scholar 

  88. R.F. Werner, M.M. Wolf, Bell inequalities and entanglement. Quantum Inf. Comput. 1, 1–25 (2001)

    MathSciNet  MATH  Google Scholar 

  89. H.M. Wiseman, The two Bell’s theorems of John Bell. J. Phys. A 47, 424001 (2014)

    Article  ADS  MathSciNet  Google Scholar 

  90. S. Wolf, Nonlocality without counterfactual reasoning. Phys. Rev. A 92, 052102 (2015)

    Article  ADS  Google Scholar 

  91. N.S. Yanofsky, The Outer Limits of Reason (The MIT Press, Cambridge, 2013)

    Google Scholar 

  92. R. Zach, Hilbert’s finitism: historical, philosophical, and metamathematical perspectives, Ph.D. thesis, University of California at Berkeley (2001)

    Google Scholar 

  93. A.K. Zvonkin, L.A. Levin, The complexity of finite objects and the development of the concepts of information and randomness by means of the theory of algorithms. Russ. Math. Surv. 25, 83–124 (1970)

    Article  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Klaas Landsman .

Editor information

Editors and Affiliations

Appendices

Appendix A: The Born Rule

The Born measure is a probability measure \(\mu _a\) on the spectrum \(\sigma (a)\) of a (bounded) self-adjoint operator a on some Hilbert space H, defined as follows by any state \(\omega \) on B(H)Footnote 30:

Theorem A.1

Let H be a Hilbert space, let \(a^*=a\in B(H)\), and let \(\omega \) be a state on B(H). There exists a unique probability measure \(\mu _a\) on the spectrum \(\sigma (a)\) of a such that

$$\begin{aligned} \omega (f(a))=\int _{\sigma (a)} d\mu _a(\lambda )\, f(\lambda ), \,\, \mathrm { for\, all }\,\, f \in C(\sigma (a)). \end{aligned}$$
(3.5.1)

The Born measure is a mathematical construction; what is its relationship to experiment? This relationship must be the source of the (alleged) randomness of quantum mechanics, for the Schrödinger equation is deterministic. We start by postulating, as usual, that \(\mu _a(\Delta )\) is the (single case) “probability” that measurement of the observable a in the state \(\omega \) (which jointly give rise to the pertinent Born measure \(\mu _a\)) gives a result \(\lambda \in \Delta \subset \sigma (a)\). Here I identify single-case “probabilities” with numbers (consistent with the probability calculus) provided by theory, upon which long-run frequencies provide empirical evidence for the theory in question, but do not define probabilities. The Born measure is a case in point: these probabilities are theoretically given, but have to be empirically verified by long runs of independent experiments. In other words, by the results reviewed below such experiments provide numbers whose role it is to test the Born rule as a hypothesis. This is justified by the following sampling theorem (strong law of large numbers): for any (measurable) subset \(\Delta \subset \sigma (a)\) and any sequence \((x_n)\in \sigma (a)^{{\mathbb N}}\) we have \(\mu _a^{\infty }\)-almost surely:

$$\begin{aligned} \lim _{N\rightarrow \infty } \frac{1}{N}(1_{\Delta }(x_1)+\cdots + 1_{\Delta }(x_N))= \mu _a(\Delta ). \end{aligned}$$
(3.5.2)

Proof of Theorem 3.4.1. Let \(a=a^*\in B(H)\), where H is a Hilbert space and B(H) is the algebra of all bounded operators on H, and let \(\sigma (a)\) be the spectrum of a. For simplicity (and since this is enough for our applications, where \(H={\mathbb C}^2\)) I assume \(\dim (H)<\infty \), so that \(\sigma (a)\) simply consists of the eigenvalues \(\lambda _i\) of a (which may be degenerate). Let us first consider a finite number N of identical measurements of a (a “run”). The first option in the theorem corresponds to a simultaneous measurement of the commuting operators

$$\begin{aligned} a_1&=a\otimes 1_H\otimes \cdots \otimes 1_H; \end{aligned}$$
(3.5.3)
$$\begin{aligned}&\cdots \nonumber \\ a_N&=1_H\otimes \cdots \otimes 1_H\otimes a, \end{aligned}$$
(3.5.4)

all defined on the N-fold tensor product \(H^N\equiv H^{\otimes N}\) of H with itself.Footnote 31 To put this in a broader perspective, consider any set \((a_1, \ldots , a_N)\equiv \underline{a}\) of commuting operators on any Hilbert space K (of which (3.5.3)–(3.5.4) is obviously a special case with \(K=H^N\)). These operators have a joint spectrum \(\sigma (\underline{a})\), whose elements are the joint eigenvalues \(\underline{\lambda }=(\lambda _1, \ldots , \lambda _N)\), defined by the property that there exists a nonzero joint eigenvector \(\psi \in K\) such that \(a_i\psi =\lambda _i\psi \) for all \(i=1, \ldots , N\); clearly,

$$\begin{aligned} \sigma (\underline{a})=\{\underline{\lambda }\in \sigma (a_1)\times \cdots \times \sigma (a_n)\mid e_{\underline{\lambda }}\equiv e^{(1)}_{\lambda _1}\cdots e^{(n)}_{\lambda _n} \ne 0\} \subseteq \sigma (a_1)\times \cdots \times \sigma (a_N), \end{aligned}$$
(3.5.5)

where \(e^{(i)}_{\lambda _i}\) is the spectral projection of \(a_i\) on the eigenspace for the eigenvalue \(\lambda _i\in \sigma (a_i)\). Von Neumann’s Born rule for the probability of finding \(\underline{\lambda }\in \sigma (\underline{a})\) then simply reads

$$\begin{aligned} p_{\underline{a}}(\underline{\lambda })=\omega (e_{\underline{\lambda }}), \end{aligned}$$
(3.5.6)

where \(\omega \) is the state on B(K) with respect to which the Born probability is defined.Footnote 32 If \(\dim (K)<\infty \), as I assume, we always have \(\omega (a)=\text{ Tr }\,(\rho a)\) for some density operator \(\rho \), and for a general Hilbert space K this is the case iff the state \(\omega \) is normal on B(K). For (normal) pure states we have \(\rho =|\psi \rangle \langle \psi |\) for some unit vector \(\psi \in K\), in which case

$$\begin{aligned} p_{\underline{a}}(\underline{\lambda })=\langle \psi ,e_{\underline{\lambda }}\psi \rangle . \end{aligned}$$
(3.5.7)

The Born rule (3.5.6) is similar to the single-operator case ([52], §4.1)Footnote 33: the continuous functional calculus gives a Gelfand isomorphism of commutative C*-algebras

$$\begin{aligned} C^*(\underline{a},1_K)\cong C(\sigma (\underline{a})), \end{aligned}$$
(3.5.8)

under which the restriction of the state \(\omega \), originally defined on B(K), to its commutative C*-subalgebra \(C^*(\underline{a})\) defines a probability measure \(\mu _{\underline{a}}\) on the joint spectrum \(\sigma (\underline{a})\) via the Riesz isomorphism. This is the Born measure, whose probabilities are given by (3.5.6). For the case (3.5.3)–(3.5.4) we have equality in (3.5.5); since in that case \(\sigma (a_i)=\sigma (a)\), we obtain

$$\begin{aligned} \sigma (\underline{a})=\sigma (a)^N, \end{aligned}$$
(3.5.9)

and therefore, for all \(\lambda _i\in \sigma (a)\) and states \(\omega \) on \(B(H^N)\), the Born rule (3.5.6) becomes

$$\begin{aligned} p_{\underline{a}}(\lambda _1, \ldots , \lambda _N)=\omega (e_{\lambda _1}\otimes \cdots \otimes e_{\lambda _N}). \end{aligned}$$
(3.5.10)

Now take a state \(\omega _1\) on B(H). Reflecting the idea that \(\omega \) is the state on \(B(H^N)\) in which N independent measurements of \(a\in B(H)\) in the state \(\omega _1\) are carried out, choose

$$\begin{aligned} \omega =\omega _1^N, \end{aligned}$$
(3.5.11)

the state on \(B(H^N)\) defined by linear extension of its action on elementary tensors:

$$\begin{aligned} \omega _1^N(b_1\otimes \cdots \otimes b_n)=\omega _1(b_1) \cdots \omega _N(b_N). \end{aligned}$$
(3.5.12)

It follows that

$$\begin{aligned} \omega ^N(e_{\lambda _1}\otimes \cdots \otimes e_{\lambda _N})=\omega _1(e_{\lambda _1})\cdots \omega _1(e_{\lambda _N})=p_a(\lambda _1)\cdots p_a(\lambda _N), \end{aligned}$$
(3.5.13)

so that the joint probability of the outcome \((\lambda _1, \ldots , \lambda _N)\in \sigma (\underline{a})\) is simply

$$\begin{aligned} p_{\vec {a}}(\lambda _1, \ldots , \lambda _N)=p_a(\lambda _1)\cdots p_a(\lambda _N). \end{aligned}$$
(3.5.14)

Since these are precisely the probabilities for option 2 (i.e. the Bernoulli process), i.e.,

$$\begin{aligned} \mu _{\underline{a}}=\mu _a^N, \end{aligned}$$
(3.5.15)

this proves the claim for \(N<\infty \). To describe the limit \(N\rightarrow \infty \), let B be any C*-algebra with unit \(1_B\); below I take \(B=B(H)\), \(B=C^*(a,1_H)\), or \(B=C(\sigma (a))\). We now take

$$\begin{aligned} A_N=B^{\otimes N}, \end{aligned}$$
(3.5.16)

the N-fold tensor product of B with itself.Footnote 34 The special cases above may be rewritten as

$$\begin{aligned} B(H)^{\otimes N}&\cong B(H^N);\end{aligned}$$
(3.5.17)
$$\begin{aligned} C^*(a,1_H)^{\otimes N}&\cong C^*(a_1, \ldots , a_N,1_{H^N}); \end{aligned}$$
(3.5.18)
$$\begin{aligned} C(\sigma (a))^{\otimes N}&\cong C(\sigma (a) \times \cdots \times \sigma (a)), \end{aligned}$$
(3.5.19)

with N copies of H and \(\sigma (a)\), respectively, and in (3.5.18) the \(a_i\) are given by (3.5.3)–(3.5.4). We may then wonder if these algebras have a limit as \(N\rightarrow \infty \). They do, but it is not unique and depends on the choice of observables, that is, of the infinite sequences \(\mathbf {a}=(a_1,a_2, \ldots )\), with \(a_N\in A_N\), that are supposed to have a limit. One possibility is to take sequences \(\mathbf {a}\) for which there exists \(M\in {\mathbb N}\) and \(a_M\in A_M\) such that for each \(N\ge M\),

$$\begin{aligned} a_N=a_M \otimes 1_B\cdots \otimes 1_B, \end{aligned}$$
(3.5.20)

with \(N-M\) copies of \(1_B\). On that choice, one obtains the infinite tensor product \(B^{\otimes \infty }\), see Landsman [52], §C.14. The limit of (3.5.17) in this sense is \(B(H^{\otimes \infty })\), where \(H^{\otimes \infty }\) is von Neumann’s ‘complete’ infinite tensor product of Hilbert spaces,Footnote 35 in which \(C^*(a,1_H)^{\otimes \infty }\) is the C*-algebra generated by \((a_1,a_2, \ldots )\) and the unit on \(H^{\otimes \infty }\). The limit of (3.5.19) is

$$\begin{aligned} C(\sigma (a))^{\otimes \infty }\cong C(\sigma (a)^{{\mathbb N}}), \end{aligned}$$
(3.5.21)

where \(\sigma (a)^{{\mathbb N}}\), which we previously saw as a measure space (as a special case of \(X^{{\mathbb N}}\) for general compact Hausdorff spaces X), is now seen as a topological space with the product topology, in which it is compact.Footnote 36 As in the finite case, we have an isomorphism

$$\begin{aligned} C^*(a,1_H)^{\otimes \infty } \cong C(\sigma (a))^{\otimes \infty }, \end{aligned}$$
(3.5.22)

and hence, on the given identifications, we obtain an isomorphism of C*-algebras

$$\begin{aligned} C^*(a_1, a_2, \ldots , 1_{H^{\otimes \infty }})\cong C(\sigma (a)^{{\mathbb N}}). \end{aligned}$$
(3.5.23)

It follows from the definition of the infinite tensor products used here that each state \(\omega _1\) on B defines a state \(\omega _1^{\infty }\) on \(B^{\otimes \infty }\). Take \(B=B(H)\) and restrict \(\omega _1^{\infty }\), which a priori is a state on \(B(H^{\otimes \infty })\), to its commutative C*-subalgebra \(C^*(a_1, a_2, \ldots , 1_{H^{\otimes \infty }})\). The isomorphism (3.5.23) then gives a probability measure \(\mu _{\underline{a}}\) on the compact space \(\sigma (a)^{{\mathbb N}}\), where the label \(\underline{a}\) now refers to the infinite set of commuting operators \((a_1, a_2, \ldots )\) on \(H^{\otimes \infty }\). To compute this measure, I use (3.5.1) and the fact that by construction functions of the type

$$\begin{aligned} f(\lambda _1, \lambda _2, \ldots )=f^{(N)}(\lambda _1, \ldots , \lambda _N), \end{aligned}$$
(3.5.24)

where \(N<\infty \) and \(f^{(N)}\in C(\sigma (a)^N)\), are dense in \(C(\sigma (a)^{{\mathbb N}})\) (with respect to the appropriate supremum-norm), and that in turn finite linear combinations of factorized functions \(f^{(N)}(\lambda _1, \ldots , \lambda _N)=f_1(\lambda _1)\cdots f_N(\lambda _N)\) are dense in \(C(\sigma (a)^{N})\). It follows from this that

$$\begin{aligned} \mu _{\underline{a}}=\mu _a^{\infty }. \end{aligned}$$
(3.5.25)

Since this generalizes (3.5.15) to \(N=\infty \), the proof of Theorem 3.4.1 is finished. \(\Box \)

Appendix B: 1-Randomness

In what follows, the notion of 1-randomness, originally defined by Martin-Löf in the setting of constructive measure theory, will be explained through an equivalent definition in terms of Kolmogorov complexity.Footnote 37 We assume basic familiarity with the notion of a computable function \(f:{\mathbb N}\rightarrow {\mathbb N}\), which may be defined through recursion theory or Turing machines.

A string is a finite succession of bits (i.e. zeros and ones). The length of a string \(\sigma \) is denoted by \(|\sigma |\). The set of all strings of length N is denoted by \(\underline{2}^N\), where \(\underline{2}=\{0,1\}\), and

$$\begin{aligned} \underline{2}^*=\bigcup _{N\in {\mathbb N}}\underline{2}^N \end{aligned}$$
(3.5.26)

denotes the set of all strings. The Kolmogorov complexity \(K(\sigma )\) of \(\sigma \in \underline{2}^*\) is defined, roughly speaking, as the length of the shortest computer program that prints \(\sigma \) and then halts. We then say, again roughly, that \(\sigma \) is Kolmogorov random if this shortest program contains all of \(\sigma \) in its code, i.e. if the shortest computable description of \(\sigma \) is \(\sigma \) itself.

To make this precise,Footnote 38 fix some universal prefix-free Turing machine U, seen as performing a computation on input \(\tau \) (in its prefix-free domain) with output \(U(\tau )\), and define

$$\begin{aligned} K(\sigma )=\min _{\tau \in \underline{2}^*} \{|\tau |: U(\tau )=\sigma \}. \end{aligned}$$
(3.5.27)

The function \(K:\underline{2}^*\rightarrow {\mathbb N}\) is uncomputable, but that doesn’t mean it is ill-defined. The choice of U affects \(K(\sigma )\) up to a \(\sigma \)-independent constant, and to take this dependency into account we state certain results in terms of the “big-O” notation familiar from Analysis.Footnote 39 For example, if \(\sigma \) is easily computable, like the first \(|\sigma |\) binary digits of \(\pi \), then

$$\begin{aligned} K(\sigma )=O(\log |\sigma |), \end{aligned}$$
(3.5.28)

with the logarithm in base 2 (as only the length of \(\sigma \) counts). However, a random \(\sigma \) has

$$\begin{aligned} K(\sigma )=|\sigma |+ O(\log |\sigma |). \end{aligned}$$
(3.5.29)

We say that \(\sigma \) is c-Kolmogorov random, for some \(\sigma \)-independent constant \(c\in {\mathbb N}\), if

$$\begin{aligned} K(\sigma )\ge |\sigma |-c. \end{aligned}$$
(3.5.30)

Simple counting arguments show that as \(|\sigma |=N\) gets large, the overwhelming majority of strings in \(\underline{2}^N\) (and hence in \(\underline{2}^*\)) is c-random.Footnote 40 The following theorem, which might be called Chaitin’s first incompleteness theorem, therefore shows that randomness is elusiveFootnote 41:

Theorem B.1

For any sound mathematical theory T containing enough arithmetic there is a constant \(C\in {\mathbb N}\) such that T cannot prove any sentence of the form \(K(\sigma )>C\) (although infinitely many such sentences are true), and as such T can only prove (Kolmogorov) randomness of finitely many strings (although infinitely many strings are in fact random).

The proof is quite complicated in its details but it is based on the existence of a computably enumerable (c.e.) list \(\mathsf {T}=(\tau _1,\tau _2,\ldots )\) of the theorems of T, and on the fact that after Gödelian encoding by numbers, theorems of any given grammatical form can be computably searched for in this list and will eventually be found. In particular, there exists a program P (running on the universal prefix-free Turing machine U used to define \(K(\cdot )\)) such that P(n) halts iff there exists a string \(\sigma \) for which \(K(\sigma )>n\) is a theorem of T. If there is such a theorem the output is \(P(n)=\sigma \), where \(\sigma \) appears in the first such theorem of the kind (according to the list \(\mathsf {T}\)). By definition of \(K(\cdot )\), this means that

$$\begin{aligned} K(\sigma )\le |P|+|n|. \end{aligned}$$
(3.5.31)

Now suppose that no C as in the above statement of the theorem exists. Then there is \(n\in {\mathbb N}\) large enough that \(n>|P|+|n|\) and there is a string \(\sigma \in \underline{2}^*\) such that T proves \(K(\sigma )>n\). Since T is sound this is actually true,Footnote 42 which gives a contradiction between

$$\begin{aligned} K(\sigma )>n>|P|+|n|;&K(\sigma )\le |P|+|n|. \end{aligned}$$
(3.5.32)

Note that this proof shows that a proof in T of \(K(\sigma )>n\) (if true) would also identify \(\sigma \).

As an idealization of a long (binary) string, a (binary) sequence \(x=x_1x_2\cdots \) is an infinite succession of bits, i.e. \(x\in \underline{2}^{{\mathbb N}}\), with finite truncations \(x_{|N}=x_1\cdots x_N\in \underline{2}^N\) for each \(N\in {\mathbb N}\). We then call x Levin–Chaitin random if each truncation of x is c-Kolmogorov random for some c, that is, if there exists \(c\in {\mathbb N}\) such that

$$\begin{aligned} K(x_{|N})\ge N - c \end{aligned}$$

for each \(N\in {\mathbb N}\). Equivalently,Footnote 43 a sequence x is Levin–Chaitin random if eventually \(K(x_{|N})>>N\), in that

$$\begin{aligned} \lim _{N\rightarrow \infty } (K(x_{|N})-N)=\infty . \end{aligned}$$
(3.5.33)

Apart from having the same intuitive pull as Kolmogorov randomness (of strings), this definition gains from the fact that it is equivalent to two other appealing notions of randomness, namely patternlessness and unpredictability, both also defined computationally. In view of these equivalences we simply call a Levin–Chaitin random sequence 1-random.Footnote 44

A sequence \(x\in \underline{k}^{{\mathbb N}}\) is Borel normal in base k if each string \(\sigma \) has frequency \(k^{-|\sigma |}\) in x. Any hope of defining randomness as Borel normality in base 10 is blocked by Champernowne’s number \(0123456789101112131 \ldots \), which is Borel normal but clearly not random in any reasonable sense (this is also true in base 2). The decimal expansion of \(\pi \) is also conjectured to be Borel normal in base 10 (with huge numerical support), although \(\pi \) clearly is not random either. However, Borel normality seems a desirable property of truly random numbers on any good definition, and so we are fortunate to have:

Proposition B.2

A 1-random sequence is Borel normal (in base 2, but in fact in any base) and hence (“monkey typewriter theorem”) contains any finite string infinitely often.Footnote 45

Another desirable property comes from the following theorem due to Martin-Löf, in which P is the 50-50 probability on \(\{0,1\}\) and \(P^{\infty }\) is the induced probability measure on \(\underline{2}^{{\mathbb N}}\):

Theorem B.3

With respect to \(P^{\infty }\) almost every outcome sequence \(x\in \underline{2}^{{\mathbb N}}\) is 1-random.

This implies that the 1-random sequences form an uncountable subset of \(\underline{2}^{{\mathbb N}}\),Footnote 46 although topologically this subset is meagre (i.e. Baire first category).Footnote 47 Chaitin’s incompleteness theorem for (finite) strings has the following counterpart for (infinite) sequences:

Theorem B.4

If \(x\in \underline{2}^{{\mathbb N}}\) is 1-random, then ZFC (or any sufficiently comprehensive mathematical theory T meant in Theorem B.1) can compute only finite many digits of x.Footnote 48

This clearly excludes defining a 1-random number by somehow listing its digits, but some can be described by a formula. One example is Chaitin’s \(\Omega \), or more precisely \(\Omega _U\),Footnote 49 which is the halting probability of some fixed universal prefix-free Turing machine U, given by

$$\begin{aligned} \Omega _U:=\sum _{\tau \in \underline{2}^*\mid U(\tau )\downarrow }2^{-|\tau |}. \end{aligned}$$
(3.5.34)

Appendix C: Bell’s Theorem and Free Will Theorem

In support of the analysis of hidden variable theories in the main text, this appendix reviews Bell’s [2] theorem and the free will theorem, streamlining earlier expositions ([17, 52], Chap. 6) and leaving out proofs and other adornments.Footnote 50 In the specific context of ’t Hooft’s theory (where the measurement settings are determined by the hidden state) and Bohmian mechanics (where they are not, as in the original formulation of Bell’s theorem and in most hidden variable theories) an advantage of my approach is that both free (uncorrelated) und correlated settings fall within its scope; the former are distinguished from the latter by an independence assumption.Footnote 51

As a warm-up I start with a version of the Kochen–Specker theorem, whose logical form is very similar to Bell’s [2] theorem and the free will theorem, as follows:

Theorem C.1

Determinism, qm, non-contextuality, and free choice are contradictory.

Of course, this unusual formulation hinges on the precise meaning of these terms.

  • determinism is the conjunction of the following two assumptions.

    1. There is a state space X with associated functions \(A: X\rightarrow S\) and \(L:X\rightarrow O\), where S is the set of all possible measurement settings Alice can choose from, namely a suitable finite set of orthonormal bases of \({\mathbb R}^3\) (11 well-chosen bases will do to arrive at a contradiction),Footnote 52 and O is some set of possible measurement outcomes. Thus some \(x\in X\) determines both Alice’s setting \(a=A(x)\) and her outcome \(\alpha =L(x)\).

    2. There exists some set \(\Lambda \) and an additional function \(H:X\rightarrow \Lambda \) such that

    $$\begin{aligned} L=L(A,H), \end{aligned}$$
    (3.5.35)

    in the sense that for each \(x\in X\) one has \(L(x)=\hat{L}(A(x),H(x))\) for a certain function \(\hat{L}:S \times \Lambda \rightarrow O\). This self-explanatory assumption just states that each measurement outcome \(L(x)=\hat{L}(a,\lambda )\) is determined by the measurement setting \(a=A(x)\) and the “hidden” variable or state \(\lambda =H(x)\) of the particle undergoing measurement.

  • qm fixes \(O=\{(0,1,1), (1,0,1), (1,1,0)\}\), which is a non-probabilistic fact of quantum mechanics with overwhelming (though indirect) experimental support.

  • non-contextuality stipulates that the function \(\hat{L}\) just introduced take the form

    $$\begin{aligned} \hat{L}((\vec {e}_1,\vec {e}_2,\vec {e}_3),\lambda )=(\tilde{L}(\vec {e}_1,\lambda ), \tilde{L}(\vec {e}_2,\lambda ), \tilde{L}(\vec {e}_3,\lambda )), \end{aligned}$$
    (3.5.36)

    for a single function \(\tilde{L}:S^2\times \Lambda \rightarrow \{0,1\}\) that also satisfies \(\tilde{L}(-\vec {e},\lambda )=\tilde{L}(\vec {e},\lambda )\).Footnote 53

  • free choice finally states that the following function is surjective:

    $$\begin{aligned} A\times H:X\rightarrow S\times \Lambda ;&x\mapsto (A(x),H(x)). \end{aligned}$$
    (3.5.37)

    In other words, for each \((a,\lambda )\in S\times \Lambda \) there is an \(x\in X\) for which \(A(x)=a\) and \(H(x)=\lambda \). This makes A and H “independent” (or: makes a and \(\lambda \) free variables).

See Landsman [52], §6.2 for a proof of the Kochen–Specker theorem in this language.Footnote 54

Bell’s [2] theorem and the free will theorem both take a similar generic form, namely:

Theorem C.2

Determinism, qm, local contextuality, and free choice, are contradictory.

Once again, I have to explain what these terms exactly mean in the given context.

  • determinism is a straightforward adaptation of the above meaning to the bipartite “Alice and Bob” setting. Thus we have a state space X with associated functions

    $$\begin{aligned} A: X\rightarrow S;&B: X\rightarrow S;&L:X\rightarrow O&R: X\rightarrow O, \end{aligned}$$
    (3.5.38)

    where S, the set of all possible measurement settings Alice and Bob can each choose from, differs a bit between the two theorems: for the free will theorem it is the same as for the Kochen–Specker theorem above, as is the set O of possible measurement outcomes, whereas for Bell’s theorem (in which Alice and Bob each measure a 2-level system), S is some finite set of angles (three is enough), and \(O=\{0,1\}\).

    • In the free will case, these functions and the state \(x\in X\) determine both the settings \(a=A(x)\) and \(b=B(x)\) of a measurement and its outcomes \(\alpha =L(x)\) and \(\beta =R(x)\) for Alice on the Left and for Bob on the Right, respectively.

    • All of this is also true in the Bell case, but since his theorem relies on impossible measurement statistics (as opposed to impossible individual outcomes), one in addition assumes a probability measure \(\mu \) on X.Footnote 55

    Furthermore, there exists some set \(\Lambda \) and some function \(H:X\rightarrow \Lambda \) such that

    $$\begin{aligned} L=L(A,B,H);&R=R(A,B,H), \end{aligned}$$
    (3.5.39)

    in the sense that for each \(x\in X\) one has functional relationships

    $$\begin{aligned} L(x)=\hat{L}(A(x),B(x),H(x));&R(x)=\hat{R}(A(x),B(x),H(x)), \end{aligned}$$
    (3.5.40)

    for certain functions \(\hat{L}:S \times S\times \Lambda \rightarrow O\) and \(\hat{R}:S \times S\times \Lambda \rightarrow O\).

  • qm reflects elementary quantum mechanics of correlated 2-level and 3-level quantum systems for the Bell and the free will cases, respectively, as followsFootnote 56:

    • In the free will theorem, \(O=\{(0,1,1), (1,0,1), (1,1,0)\}\) is the same as for the Kochen–Specker theorem. In addition perfect correlation obtains: if \(a=(\vec {e}_1,\vec {e}_2,\vec {e}_3)\) is Alice’s orthonormal basis and \(b=(\vec {f}_1,\vec {f}_2,\vec {f}_3)\) is Bob’s, one has

      $$\begin{aligned} \vec {e}_i=\vec {f}_j\, \Rightarrow \, \hat{L}_i(a,b,z)=\hat{R}_j(a,b,z), \end{aligned}$$
      (3.5.41)

      where \(\hat{L}_i, \hat{R}_j: S \times S\times \Lambda \rightarrow \{0,1\}\) are the components of \(\hat{L}\) and \(\hat{R}\), respectively. Finally,Footnote 57 if \((a',b'\)) differs from (ab) by changing the sign of any basis vector,

      $$\begin{aligned} \hat{L}(a',b',\lambda )=\hat{L}(a,b,\lambda );&\hat{R}(a',b',\lambda )=\hat{R}(a,b,\lambda ). \end{aligned}$$
      (3.5.42)
    • In Bell’s theorem, \(O=\{0,1\}\), and the statistics for the experiment is reproduced as conditional joint probabilities given by the measure \(\mu \) through

      $$\begin{aligned} P(L\ne R|A=a,B=b)=\sin ^2(a-b). \end{aligned}$$
      (3.5.43)
  • local contextuality, which replaces and weakens non-contextuality, means that

    $$\begin{aligned} L(A,B,H)=L(A,H);&R(A,B,H)=G(B,H). \end{aligned}$$
    (3.5.44)

    In words: Alice’s outcome given \(\lambda \) does not depend on Bob’s setting, and vice versa.

  • free choice is an independence assumption that looks differently for both theorems:

    • In the free will theorem it means that each \((a,b,\lambda )\in S\times S\times \Lambda \) is possible in that there is an \(x\in X\) for which \(A(x)=a\), \(B(x)=b\), and \(H(x)=\lambda \).

    • In Bell’s theorem, (ABH) are probabilistically independent relative to \(\mu \).Footnote 58

This concludes the joint statement of the free will theorem and Bell’s [2] theorem in the form we need for the main text. The former is proved by reduction to the Kochen–Specker theorem, whilst the latter follows by reduction to the usual version of Bell’s theorem via the free choice assumption; see Landsman [52], Chap. 6 for details.

For our purposes these theorems are equivalent, despite subtle differences in their assumptions. Bell’s theorem is much more robust in that it does not rely on perfect correlations (which are hard to realize experimentally), and in addition it requires almost no input from quantum theory. On the other hand, Bell’s theorem uses probability theory in a highly nontrivial way: like the hidden variable theories it is supposed to exclude it relies on the possibility of fair sampling of the probability measure \(\mu \). The factorization condition defining probabilistic independence passes this requirement of fair sampling on to both the hidden variable and the settings, which brings us back to the main text.

Different parties may now be identified by the assumption they drop: Copenhagen quantum mechanics rejects determinism, Valentini [86] rejects the Born rule and hence qm, Bohmians rejects local contextuality, and finally ’t Hooft rejects free choice. However, as we argue in the main text, even the latter two camps do not really have a deterministic theory underneath quantum mechanics because of their need to randomly sample the probability measure they must use to recover the predictions of quantum mechanics.

Rights and permissions

Reprints and permissions

Copyright information

© 2021 The Author(s), under exclusive license to Springer Nature Switzerland AG

About this chapter

Check for updates. Verify currency and authenticity via CrossMark

Cite this chapter

Landsman, K. (2021). Indeterminism and Undecidability. In: Aguirre, A., Merali, Z., Sloan, D. (eds) Undecidability, Uncomputability, and Unpredictability. The Frontiers Collection. Springer, Cham. https://doi.org/10.1007/978-3-030-70354-7_3

Download citation

Publish with us

Policies and ethics

Navigation