Skip to main content
Log in

The Status of Determinism in Proofs of the Impossibility of a Noncontextual Model of Quantum Theory

  • Published:
Foundations of Physics Aims and scope Submit manuscript

Abstract

In order to claim that one has experimentally tested whether a noncontextual ontological model could underlie certain measurement statistics in quantum theory, it is necessary to have a notion of noncontextuality that applies to unsharp measurements, i.e., those that can only be represented by positive operator-valued measures rather than projection-valued measures. This is because any realistic measurement necessarily has some nonvanishing amount of noise and therefore never achieves the ideal of sharpness. Assuming a generalized notion of noncontextuality that applies to arbitrary experimental procedures, it is shown that the outcome of a measurement depends deterministically on the ontic state of the system being measured if and only if the measurement is sharp. Hence for every unsharp measurement, its outcome necessarily has an indeterministic dependence on the ontic state. We defend this proposal against alternatives. In particular, we demonstrate why considerations parallel to Fine’s theorem do not challenge this conclusion.

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

Access this article

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

Instant access to the full article PDF.

Similar content being viewed by others

Notes

  1. More precisely, they rule out models that satisfy Bell’s notion of local causality and certain other assumptions, such as the freedom of settings.

  2. The fact that all noncontextual ontological models must represent sharp measurements outcome-deterministically implies in particular that the \(\psi \)-complete ontological model of quantum theory fails to be a noncontextual model. The \(\psi \)-complete ontological model of quantum theory is the one wherein the ontic states are the pure quantum states (i.e. no hidden variables) and the response function associated to an effect is simply the conditional probability defined by the Born rule [19]. This model is sometimes called the “orthodox” interpretation of quantum theory. Given the use of the Born rule, the response functions in this model are outcome-indeterministic even for sharp measurements and therefore our result implies that the model cannot be noncontextual. This inference is correct because the \(\psi \)-complete ontological model fails to be preparation noncontextual, as explained in Sect. VIII.B of Ref. [7].

  3. More generally, an ontological model may be defined such that \(\Lambda \) is a continuous space, in which case \(\mu :\mathcal {P}\times \Lambda \rightarrow \mathbb {R}_+\) is a probability density and sums over \(\lambda \) are replaced by integrals.

  4. There is a subtlety in our choice of terminology which might lead to confusion. For a set of response functions to be considered outcome-indeterministic, it is sufficient for a single response function in the set to be outcome-indeterministic. Hence, every set of response functions is either outcome-deterministic or outcome-indeterministic. On the other hand, for a given class of measurements to be considered outcome-(in)deterministic, it is necessary for every measurement in the class to be outcome-(in)deterministic. Consequently, we might have neither outcome-determinism nor outcome-indeterminism for a given class.

  5. Harrigan and Rudolph have also previously discussed ontological models that assign ontic state spaces to the measurement apparatus [21].

  6. Terms coined by John Smolin and Matt Leifer respectively.

  7. This question was posed by a referee of this article.

  8. It is tempting to think that constraints KS1–KS3 are the content of the assumption of traditional noncontextuality, but this is inaccurate. Rather, the assumption of measurement noncontextuality is a prerequisite to KS1–KS3 making any sense. It is this assumption that warrants positing a function \(v\) that depends only on the projector associated with a measurement outcome. So once one is discussing the properties of a valuation over projectors, the assumption of noncontextuality has already done its work. KS2 then follows from how one must represent coarse-graining in an ontological model (given by Eq. (10)), KS3 follows from the fact that for every measurement, the sum of probabilities of all the outcomes must be 1, and KS1 encodes the assumption of outcome determinism for sharp measurements.

  9. We do not claim that any actual proponent of ODUM would make the arguments that are made by our imaginary proponent. Nonetheless, certain parts of our dialogue are inspired by various proposals for how to define a notion of noncontextuality for unsharp measurements, as we note explicitly throughout.

  10. Grudka and Kurzynski [29] have also criticized the notion of noncontextuality used in the Cabello-Nakamura proofs. They argue that in a noncontextual model, one should only assign deterministic values to the projectors that appear in a Naimark extension of the POVM, rather than the POVM elements themselves. It then suffices to note that the projector that extends a given effect varies with the POVM in which that effect appears, and therefore that a noncontextual model does not assign a unique deterministic value to a given effect. In the language of the present article, they argue that a noncontextual and outcome-deterministic value-assignment to projectors on system+ancilla does not imply a noncontextual and outcome-deterministic value-assignment to effects on the system. This attitude is entirely consistent with the view espoused here.

  11. This proposal was considered by Methot [13].

  12. This notion was considered by Bacon et al. [30] and reported by Methot [13].

  13. Just as one can define a quantum Naimark extension by adjoining an ancilla to one’s system or by considering the system to be a subspace of a larger system, so too can one define an ontological Naimark extension in either way. We’ll use the ancilla construction here. It is possible that one could dispense with the assumption of separability if one used the subspace construction, but we do not seek to answer the question here.

  14. As an aside, note that the pair \(M\) and \(M'\) provide yet another way of understanding why the POVM \(\{ \tfrac{1}{2} I, \tfrac{1}{2} I\}\) must be represented by the set of response functions \(\{ \tfrac{1}{2},\tfrac{1}{2}\}\). If the set of response functions representing \(M\) is denoted \(\left\{ \xi \left( 0| \lambda , M\right) ,\xi \left( 1| \lambda ,M \right) \right\} \), and similarly for \(M'\), then from the definition of \(M^{\prime }\), we must have

    $$\begin{aligned} \xi \left( 0| \lambda ,M'\right)&=\xi \left( 1| \lambda ,M \right) ,\end{aligned}$$
    (38)
    $$\begin{aligned} \xi \left( 1| \lambda ,M'\right)&=\xi \left( 0| \lambda ,M \right) . \end{aligned}$$
    (39)

    Because \(M\) and \(M'\) are statistically indistinguishable for all preparations, by measurement noncontextuality they must be represented by the same set of response functions. Hence

    $$\begin{aligned} \xi \left( 0| \lambda ,M'\right)&=\xi \left( 0| \lambda ,M \right) ,\end{aligned}$$
    (40)
    $$\begin{aligned} \xi \left( 1| \lambda ,M'\right)&=\xi \left( 1| \lambda ,M \right) . \end{aligned}$$
    (41)

    But this implies that \(\xi \left( 0| \lambda ,M \right) =\xi \left( 1| \lambda ,M \right) \), and given that \(\xi \left( 0| \lambda ,M \right) +\xi \left( 1| \lambda ,M \right) =1\) for all \(\lambda \in \Lambda ,\) it follows that

    $$\begin{aligned} \xi \left( 0| \lambda ,M \right) =\xi \left( 1| \lambda ,M \right) =\tfrac{1}{2}\;\forall \lambda \in \Lambda . \end{aligned}$$
    (42)

References

  1. Bell, J.S.: On the problem of hidden variables in quantum mechanics. Rev. Mod. Phys. 38, 447 (1966). (Reprinted in Ref. [3], chap. 1.)

  2. Kochen, S., Specker, E.P.: The problem of hidden variables in quantum mechanics. J. Math. Mech. 17, 59 (1967)

    MathSciNet  MATH  Google Scholar 

  3. Bell, J.S.: Speakable and Unspeakable in Quantum Mechanics. Cambridge University Press, New York (1987)

    MATH  Google Scholar 

  4. Cabello, A., Garcia-Alcaine, G.: Proposed experimental tests of the Bell-Kochen-Specker theorem. Phys. Rev. Lett. 80, 1797 (1998)

    Article  ADS  Google Scholar 

  5. Simon, C., Brukner, C., Zeilinger, A.: Hidden variable theorems for real experiments. Phys. Rev. Lett. 86, 4427–4430 (2001)

    Article  ADS  Google Scholar 

  6. Larsson, J.-A.: A Kochen-Specker inequality. Europhys. Lett. 58, 799–805 (2002)

    Article  ADS  Google Scholar 

  7. Spekkens, R.W.: Contextuality for preparations, transformations, and unsharp measurements. Phys. Rev. A 71, 052108 (2005)

    Article  ADS  Google Scholar 

  8. Bell, J.S.: On the Einstein-Podolsky-Rosen paradox. Physics 1, 195 (1964). (Reprinted in Ref. [3], chap. 2.)

  9. Busch, P.: Quantum states and generalized observables: a simple proof of Gleason’s theorem. Phys. Rev. Lett. 91, 120403 (2003)

    Article  MathSciNet  ADS  Google Scholar 

  10. Cabello, A.: Kochen-Specker theorem for a single Qubit using positive operator-valued measures. Phys. Rev. Lett. 90, 190401 (2003)

    Article  MathSciNet  ADS  Google Scholar 

  11. Caves, C.M., Fuchs, C.A., Manne, K., Renes, J.M.: Gleason-type derivations of the quantum probability rule for generalized measurements. Found. Phys. 34, 193 (2004)

    Article  MathSciNet  MATH  ADS  Google Scholar 

  12. Aravind, P.K.: The generalized Kochen-Specker theorem. Phys. Rev. A 68, 052104 (2003)

    Article  MathSciNet  ADS  Google Scholar 

  13. Methot, A.A.: Minimal Bell-Kochen-Specker proofs with POVMs on qubits. Int. J. Quantum Inf. 5, 353 (2007)

    Article  MATH  Google Scholar 

  14. Zhang, Q., Li, H., Yang, T., Yin, J., Du, J., Pan, J.W.: Experimental test of the Kochen-Specker theorem for single qubits using positive operator-valued measures. arXiv:0412049v2 (2004)

  15. Mancinska, L., Scarpa, G., Severini, S.: New separations in zero-error channel capacity through projective Kochen-Specker sets and quantum coloring. IEEE Trans. Inform. Theory 59, 4025 (2013)

    Article  MathSciNet  Google Scholar 

  16. Fine, A.: Hidden variables, joint probability, and the Bell inequalities. Phys. Rev. Lett. 48, 291 (1982)

    Article  MathSciNet  ADS  Google Scholar 

  17. Simon, C., Brukner, C., Zeilinger, A.: Hidden-variable theorems for real experiments. Phys. Rev. Lett. 86, 4427 (2001)

    Article  ADS  Google Scholar 

  18. Peres, A.: Quantum Theory: Concepts and Methods. Kluwer Academic, Boston (1995)

    MATH  Google Scholar 

  19. Harrigan, N., Spekkens, R.W.: Einstein, incompleteness, and the epistemic view of quantum states. Found. Phys. 40, 125 (2010)

    Article  MathSciNet  MATH  ADS  Google Scholar 

  20. Spekkens, R.W.: The paradigm of kinematics and dynamics must yield to causal structure. arXiv:1209.0023 (2012)

  21. Harrigan, N., Rudolph, T.: Ontological models and the interpretation of contextuality. arXiv:0709.4266 (2007)

  22. Hardy, L.: Nonlocality for two particles without inequalities for almost all entangled states. Phys. Rev. Lett. 71, 1665 (1993)

    Article  MathSciNet  MATH  ADS  Google Scholar 

  23. Greenberger, D.M., Horne, M.A., Zeilinger, A.: Going beyond Bells theorem. In Bells theorem, quantum theory and conceptions of the universe (pp. 69–72). Springer, Netherlands (1989).

  24. Spekkens, R.W., Buzacott, D.H., Keehn, A.J., Toner, B., Pryde, G.J.: Preparation contextuality powers parity-oblivious multiplexing. Phys. Rev. Lett. 102, 010401 (2009)

    Article  ADS  Google Scholar 

  25. Howard, M., Wallman, J., Veitch, V., Emerson, J.: Contextuality supplies the magic for quantum computation. Nature 510, 351 (2014)

    ADS  Google Scholar 

  26. Bartlett, S.D., Rudolph, T., Spekkens, R.W.: Reconstruction of Gaussian quantum mechanics from Liouville mechanics with an epistemic restriction. Phys. Rev. A 86, 012103 (2012)

    Article  ADS  Google Scholar 

  27. Gross, D.: Hudsons theorem for finite-dimensional quantum systems. J. Math. Phys. 47, 122107 (2006)

    Article  MathSciNet  ADS  Google Scholar 

  28. Spekkens, R.W.: Negativity and contextuality are equivalent notions of nonclassicality. Phys. Rev. Lett. 101, 020401 (2008)

    Article  MathSciNet  ADS  Google Scholar 

  29. Grudka, A., Kurzynski, P.: Is there contextuality for a single qubit? Phys. Rev. Lett. 100, 160401 (2008)

    Article  MathSciNet  ADS  Google Scholar 

  30. Toner, B.F., Bacon, D., Ben-Or, M.: Kochen-Specker theorem for generalized measurements. Unpublished manuscript (2005).

  31. Kunjwal, R., Ghosh, S.: A minimal state-dependent proof of measurement contextuality for a qubit. arXiv:1305.7009 (2013)

  32. Liang, Y.C., Spekkens, R.W., Wiseman, H.M.: Specker’s parable of the overprotective seer: a road to contextuality, nonlocality and complementarity. Phys. Rep. 506, 1 (2011)

    Article  MathSciNet  ADS  Google Scholar 

  33. Spekkens, R.W., Kunjwal, R.: In preparation.

Download references

Acknowledgments

The author would like to thank John Sipe and Howard Wiseman for their insistence that the topic of outcome-determinism for unsharp measurements deserved a better treatment. Thanks also to Ernesto Galvão, Ben Toner, Howard Wiseman, and Ravi Kunjwal for discussions, and to an anonymous referee for suggesting a simplification of the proof of Theorem 1. Research at Perimeter Institute is supported in part by the Government of Canada through NSERC and by the Province of Ontario through MRI.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Robert W. Spekkens.

Appendices

Appendix 1: Ontological Extension: Modelling an Unsharp Measurement Using an Outcome-deterministic Response Function

The intuition at play in premiss 2 of the argument from Sect. 3 is that one can always imagine any subjective uncertainty in the outcome of a measurement on some system \(s\) as being due to uncertainty about the ontic state of some other system that, together with \(s,\) determines the outcome, for instance, hidden variables in the apparatus. We will see that this is indeed the case. The resulting representation of the measurement will be called an ontological extension. This is the analogue, within the ontological model, of the Naimark extension of a measurement within operational quantum theory.Footnote 13

Suppose that for an operational theory \(\left( \mathcal {P},\mathcal {M} ,p\right) \) on system \(s,\) we have found a (possibly contextual) ontological model \(\left( \Lambda _{s},\mu ,\xi \right) \) on \(s,\) such that the operational statistics are reproduced as

$$\begin{aligned} p(k|M,P)=\sum _{\lambda _{s}\in \Lambda _{s}}\mu (\lambda _{s}|P)\xi (k|\lambda _{s},M). \end{aligned}$$
(24)

We suppose that the ontological model is outcome-indeterministic, so that \(0<\xi (k|\lambda _{s},M)<1\) for some \(\lambda _{s}.\)

We can define a new model with outcome-deterministic response functions as follows. Introduce an ancilla system \(a\) with ontic state space equal to the unit interval (hence continuous), that is, \(\Lambda _{a}=[0,1]\). The ontic state space of the composite is then \(\Lambda _{sa}\equiv \Lambda _{s}\times \Lambda _{a}.\) Now define

$$\begin{aligned} \mu (\lambda _{s},\lambda _{a})\equiv \mu (\lambda _{s}|P)\mu _{0}(\lambda _{a}), \end{aligned}$$
(25)

where \(\mu _{0}(\lambda _{a})\) is the uniform distribution over \(\left[ 0,1\right] \). Also, define

$$\begin{aligned} \omega _{0}(\lambda _{s})&= 0, \end{aligned}$$
(26)
$$\begin{aligned} \omega _{k}(\lambda _{s})&= \sum _{j=1}^{k} \xi (j|\lambda _{s},M), \end{aligned}$$
(27)

and define outcome-deterministic response functions on \(\Lambda _{s} \times \Lambda _{a}\) by

$$\begin{aligned} \xi \left( k| \lambda _{s},\lambda _{a}\right) \equiv \left\{ \begin{array}{l} 1\quad \text { if } \ \omega _{k-1}(\lambda _{s})\le \lambda _{a}\le \omega _{k} (\lambda _{s})\\ 0 \quad \mathrm{otherwise} \\ \end{array}\right. \end{aligned}$$
(28)

The new model is empirically adequate because

$$\begin{aligned}&\sum _{\lambda _{s}\in \Lambda _{s}}\int _{\Lambda _{a}}\hbox {d}\lambda _{a}\; \mu (\lambda _{s},\lambda _{a})\xi (k| \lambda _{s},\lambda _{a})\end{aligned}$$
(29)
$$\begin{aligned}&\quad =\sum _{\lambda _{s}\in \Lambda _{s}}\mu (\lambda _{s}|P) \int _{\omega _{k-1}(\lambda _{s})}^{\omega _{k}(\lambda _{s})}\hbox {d}\lambda _{a} \; \mu _{0}\left( \lambda _{a}\right) \end{aligned}$$
(30)
$$\begin{aligned}&\quad =\sum _{\lambda _{s}\in \Lambda _{s}}\mu (\lambda _{s}|P) \left[ \omega _{k}(\lambda _{s})-\omega _{k-1}(\lambda _{s})\right] \end{aligned}$$
(31)
$$\begin{aligned}&\quad =\sum _{\lambda _{s}\in \Lambda _{s}}\mu (\lambda _{s}|P)\xi (k|\lambda _{s},M)\end{aligned}$$
(32)
$$\begin{aligned}&\quad =p\left( k|P,M\right) . \end{aligned}$$
(33)

A similar trick for devising an ontological model that is outcome-deterministic was described by Bell [1] and its relevance for eliminating determinism as an assumption in the proof of Bell’s theorem was emphasized by Fine [16]. It follows that one can always extend an ontological model such that it represents measurements outcome-deterministically if one so wishes. This justifies premiss \(2'\) of the second argument in Sect. 3.

Appendix 2: The Multiplicity of Naimark Extensions of a POVM and its Significance for Ontological Models

We here show explicitly that a given POVM can have two distinct Naimark extensions, that is, that there are multiple distinct choices of projective measurement on \(sa\) that yield the same POVM on \(s.\) Specifically, we construct two Naimark extensions of the fair coin flip POVM \(\{ \tfrac{1}{2} I, \tfrac{1}{2} I\}\) (discussed in Sect. 6).

The first Naimark extension is as follows. We implement a preparation of the ancilla corresponding to the mixed state

$$\begin{aligned} \rho _a=\tfrac{1}{3}\Big \vert 1\rangle \langle 1\Big \vert +\tfrac{1}{3}\Big \vert 2\rangle \langle 2\Big \vert +\tfrac{1}{3}\Big \vert 3\rangle \langle 3\Big \vert . \end{aligned}$$

The measurement on \(sa\) is the binary-outcome projector-valued measure \(\{ \Pi _{sa}, I-\Pi _{sa} \}\), where

$$\begin{aligned} \Pi _{sa}&\equiv \Pi _{s,1}\otimes \left| 1\right\rangle \left\langle 1\right| +\Pi _{s,2}\otimes \left| 2\right\rangle \left\langle 2\right| +\Pi _{s,3}\otimes \left| 3\right\rangle \left\langle 3\right| . \end{aligned}$$
(34)

where \(\Pi _{s,i}\equiv \left| \psi _i \right\rangle \left\langle \psi _i\right| \) and \(\{ \left| \psi _i \right\rangle : i \in \{1,2,3\}\}\) are a triple pure states on the system \(s\) with pairwise overlaps \(|\langle \psi _i | \psi _j \rangle |^2 = \tfrac{1}{4}\) for \(i\ne j\), that is, in the Bloch representation, they are separated by 120 \(^{\circ }\) on an equatorial plane of the Bloch sphere. To see that the effective measurement on the system is the POVM \(\{\frac{1}{2} I,\frac{1}{2} I\}\), it suffices to note that

$$\begin{aligned}&\mathrm {Tr}_a\left( \Pi _{sa} \rho _a \right) =\tfrac{1}{3}\Pi _{s,1}+\tfrac{1}{3}\Pi _{s,2}+\tfrac{1}{3}\Pi _{s,3} =\tfrac{1}{2}I. \end{aligned}$$

The second Naimark extension is as follows. We implement a preparation of the ancilla corresponding to the mixed state

$$\begin{aligned} \sigma _a=\tfrac{1}{2}\left| 1\right\rangle \left\langle 1\right| +\tfrac{1}{2}\left| 2\right\rangle \left\langle 2\right| , \end{aligned}$$
(35)

and implement a measurement on the composite of system+ancilla corresponding to the three-outcome projector-valued measure

$$\begin{aligned}&\{I\otimes \left| 1\right\rangle \left\langle 1\right| ,I\otimes \left| 2\right\rangle \left\langle 2\right| ,I\otimes \left| 3\right\rangle \left\langle 3\right| \}\end{aligned}$$
(36)
$$\begin{aligned}&\equiv \{\Pi _{sa,1},\Pi _{sa,2},\Pi _{sa,3}\} \end{aligned}$$
(37)

(although technically this is a measurement on the composite, it is obviously only nontrivial on the ancilla). Again, it is straightforward to verify that the effective measurement on the system is the POVM \(\{\tfrac{1}{2}I,\tfrac{1}{2}I\}\).

The pair of projector-valued measures appearing in the two Naimark extensions, \(\left\{ \Pi _{sa},I-\Pi _{sa}\right\} \) and \(\{\Pi _{sa,1},\Pi _{sa,2},\Pi _{sa,3}\}\), are clearly distinct. This implies that there are quantum states on \(sa\) that yield different statistics for the two. For instance, the state \(\tfrac{1}{2}I\otimes \left| 3\right\rangle \left\langle 3\right| \) yields 50/50 statistics for \(\left\{ \Pi _{sa},I-\Pi _{sa}\right\} \) but always yields the third outcome for \(\{\Pi _{sa,1},\Pi _{sa,2},\Pi _{sa,3}\}\). This implies that although these projector-valued measures are each represented by a set of outcome-deterministic response functions in an ontological model that is noncontextual in the sense of definition 1 (by virtue of Theorem 1), nonetheless the two sets of response functions must be different to account for the differing statistics.

Appendix 3: An Explicit Example of Ontological Extension

We have seen in appendix 2 that because a quantum measurement can be Naimark-extended in many ways, and because those extensions may be statistically distinguishable, it follows that the sets of response functions that represent these Naimark extensions must also be inequivalent. This section provides a second way of understanding the fact that the cost of modelling statistically-indistinguishable measurements outcome-deterministically is that on the extended system, their representations are no longer equivalent. We consider the trick described in appendix 1 for replacing a response function on the system with an outcome-deterministic response function on an extension of the system. Specifically, we show by example that a pair of measurements that are represented by the same set of response functions on the system may need to be represented by different sets of response functions on its extension.

Let \(M\) be a measurement associated with the fair coin flip POVM \(\{ \tfrac{1}{2} I, \tfrac{1}{2} I\}\) that was discussed in Sect. 6. Let the second measurement \(M^{\prime }\) be defined as follows.

The procedure \(M'\): implement \(M\), and upon obtaining outcome \(b\in \left\{ 0,1\right\} ,\) output \(b\oplus 1\), that is, the bit-flip of \(b\).

Equivalently, we can say simply that \(M'\) is a measurement of \(M\) with the outcomes permuted. Clearly \(M^{\prime }\) is also represented by the fair coin flip POVM. Now consider how to represent \(M\) and \(M^{\prime }\) within a noncontextual ontological model. As argued in Sect. 6, the set of response functions must be simply \(\{ \tfrac{1}{2},\tfrac{1}{2}\}\). Footnote 14

By the construction in appendix 1 , we can model both \(M\) and \(M'\) on a larger system, a composite of system \(s\) and an ancilla \(a\) with ontic state space \(\Lambda _s \times \Lambda _a\), in such a way that the response functions are outcome-deterministic. Specifically, if we apply this construction to \(M\), we obtain

$$\begin{aligned} \xi \left( 0| \lambda _{s},\lambda _{a},M \right)&\equiv \left\{ \begin{array} [c]{l} 1\quad \text { if }0\le \lambda _{a}\le 1/2\\ 0 \quad \mathrm{otherwise} \end{array} ,\right. \end{aligned}$$
(43)
$$\begin{aligned} \xi \left( 1| \lambda _{s},\lambda _{a},M\right)&\equiv \left\{ \begin{array} [c]{l} 1\quad \text { if }1/2\le \lambda _{a}\le 1\\ 0\quad otherwise \end{array} \ .\right. \end{aligned}$$
(44)

Meanwhile, if we remember the definition of \(M^{\prime },\) it is clear that it must be represented by a set of response functions that is simply the permutation of those for \(M\),

$$\begin{aligned} \xi \left( 0| \lambda _{s},\lambda _{a},M'\right)&= \xi \left( 1|\lambda _{s},\lambda _{a},M\right) ,\end{aligned}$$
(45)
$$\begin{aligned} \xi \left( 1| \lambda _{s},\lambda _{a},M'\right)&= \xi \left( 0| \lambda _{s},\lambda _{a},M\right) . \end{aligned}$$
(46)

Now we see that although we’ve managed to have both \(M\) and \(M^{\prime }\) represented by sets of response functions on \(\Lambda _s \times \Lambda _a\) that are outcome-deterministic, the response functions are not equivalent,

$$\begin{aligned} \xi \left( 0| \lambda _{s},\lambda _{a},M'\right)&\ne \xi \left( 0|\lambda _{s},\lambda _{a},M\right) ,\end{aligned}$$
(47)
$$\begin{aligned} \xi \left( 1| \lambda _{s},\lambda _{a},M'\right)&\ne \xi \left( 1| \lambda _{s},\lambda _{a},M\right) . \end{aligned}$$
(48)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Spekkens, R.W. The Status of Determinism in Proofs of the Impossibility of a Noncontextual Model of Quantum Theory. Found Phys 44, 1125–1155 (2014). https://doi.org/10.1007/s10701-014-9833-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10701-014-9833-x

Keywords

Navigation