Skip to main content
Log in

A Rigorous Analysis of the Clauser–Horne–Shimony–Holt Inequality Experiment When Trials Need Not be Independent

  • Published:
Foundations of Physics Aims and scope Submit manuscript

Abstract

The Clauser–Horne–Shimony–Holt (CHSH) inequality is a constraint that local hidden variable theories must obey. Quantum Mechanics predicts a violation of this inequality in certain experimental settings. Treatments of this subject frequently make simplifying assumptions about the probability spaces available to a local hidden variable theory, such as assuming the state of the system is a discrete or absolutely continuous random variable, or assuming that repeated experimental trials are independent and identically distributed. In this paper, we do two things: first, show that the CHSH inequality holds even for completely general state variables in the measure-theoretic setting, and second, demonstrate how to drop the assumption of independence of subsequent trials while still being able to perform a hypothesis test that will distinguish Quantum Mechanics from local theories. The statistical strength of such a test is computed.

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.

Fig. 1
Fig. 2

Similar content being viewed by others

Notes

  1. The reader may note that confirming these two assumptions by appealing to experimental data would require an assumption that the random variable sequences \(\{A_i\}\) and \(\{B_i\}\) are i.i.d.—exactly the sort of assumption we are trying to avoid in this paper. However, the difference is this: we observe \(\{A_i\}\) and \(\{B_i\}\), and we may come to a reasonable conclusion that we are observing an i.i.d. sequence, whereas we will never be able to conclude this about the unobserved sequence \(\{\lambda _i\}\).

References

  1. Bell, J.: On The Einstein Podolsky Rosen Paradox. Physics 1, 195–200 (1964)

    Google Scholar 

  2. Clauser, J., Horne, A., Shimony, A., Holt, R.: Proposed experiment to test local hidden-variable theories. Phys. Rev. Lett. 23, 880–884 (1969)

    Article  ADS  Google Scholar 

  3. Barrett, J., Hardy, L., Kent, A.: No signaling and quantum key distribution. Phys. Rev. Lett. 95, 010503 (2005)

    Article  ADS  Google Scholar 

  4. Acín, A., Brunner, N., Gisin, N., Massar, S., Pironio, S., Scarani, V.: Device-independent security of quantum cryptography against collective attacks. Phys. Rev. Lett. 98, 230501 (2007)

    Article  ADS  Google Scholar 

  5. Pironio, S., et al.: Random numbers certified by Bell’s theorem. Nature 464, 1021–1024 (2010)

    Article  ADS  Google Scholar 

  6. Fritz, T.: Beyond Bell’s theorem: correlation scenarios. New J. Phys. 14(10), 103001 (2012)

    Article  ADS  MathSciNet  Google Scholar 

  7. A. Brandenburger, H.J. Keisler.: Fiber products of measures and quantum foundations. URL http://pages.stern.nyu.edu/abranden/fpmqf-10-29-12.pdf. To appear in Logic and Algebraic Structures in Quantum Computing and Information. Lecture Notes in Logic, Association for Symbolic Logic, Cambridge University Press, Cambridge (2012)

  8. Barrett, J., Collins, D., Hardy, L., Kent, A., Popescu, S.: Quantum nononlocality, Bell inequalities, and the memory loophole. Phys. Rev. A 66, 042111 (2002)

    Article  ADS  Google Scholar 

  9. Gill, R.D.: Accardi Contra Bell (Cum Mundi): the impossible coupling. Math. Stat. Appl. Festschr. Constance van Eeden IMS Lect. Notes Monogr. 42, 133–154 (2003)

    Google Scholar 

  10. Hänggi, E., Renner, R., Wolf, S.: The impossibility of non-signaling privacy amplification. Theor. Comput. Sci. 486, 27–42 (2013)

    Article  MATH  Google Scholar 

  11. Barrett, J., Colbeck, R., Kent, A.: Memory attacks on device-independent quantum cryptography. Phys. Rev. Lett. 110, 010503 (2013)

    Article  ADS  Google Scholar 

  12. van Dam, W., Gill, R.D., Grunwald, P.D.: The statistical strength of nonlocality proofs. IEEE Trans. Inf. Theory 51, 2812–2835 (2005)

    Article  MATH  Google Scholar 

  13. Zhang, Y., Glancy, S., Knill, E.: Asymptotically optimal data analysis for rejecting local realism. Phys. Rev. A 84, 062118 (2011)

    Article  ADS  Google Scholar 

  14. Chung, K.L.: A Course in Probability Theory. A Course in Probability Theory, 2nd edn. Academic Press, San Diego (1974)

    Google Scholar 

  15. Brandenburger, A., Yanofsky, N.: A classification of hidden-variable properties. J. Phys. A 41, 425302 (2008)

    Article  ADS  MathSciNet  Google Scholar 

  16. Hoeffding, W.: Probability inequalities for sums of bounded random variables. J. Am. Stat. Assoc. 58, 13–30 (1963)

    Article  MATH  MathSciNet  Google Scholar 

  17. Azuma, K.: Weighted sums of certain dependent random variables. Tohoku Math. J. 19(3), 357–367 (1967)

    Article  MATH  MathSciNet  Google Scholar 

  18. Zhang, Y., Glancy, S., Knill, E.: Efficient quantification of experimental evidence against local realism. Phys. Rev. A 88, 052119 (2013).

  19. Gisin, N.: Non-realism: deep Thought or a Soft Option? Found. Phys. 42, 80–85 (2012)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  20. Weihs, G., Jennewein, T., Simon, C., Weinfurter, H., Zeilinger, A.: Violation of Bell’s inequality under strict Einstein locality conditions. Phys. Rev. Lett. 81, 5039–5043 (1998)

  21. Pearle, P.M.: Hidden-variable example based upon data rejection. Phys. Rev. D 2, 1418–1425 (1970)

  22. Clauser, J., Horne, M.: Experimental consequences of objective local theories. Phys. Rev. Lett. 10(2), 526–535 (1974)

    ADS  Google Scholar 

  23. Mermin, N.D., Garg, A.: Detector inefficiencies in the Einstein–Podolsky–Rosen experiment. Phys. Rev. D 35(12), 3831–3835 (1987)

    Article  ADS  Google Scholar 

  24. P. Bierhorst, A mathematical foundation for locality. Ph.D. thesis, Tulane University (2014)

Download references

Acknowledgments

The author would like to thank Michael Mislove and Keye Martin for their support and guidance, as well as Gustavo Didier and Lev Kaplan for their helpful comments and suggestions. This work was partially supported by grant FA9550-13-1-0135 from the US Air Force Office of Scientific Research and Grant N00014-10-1-0329 P00004 from the US Office of Naval Research.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Peter Bierhorst.

Appendix

Appendix

In this paper, we worked in the most general measure-theoretic setting. In addition to requiring more work, the general setting can make it harder to gain an intuitive grasp of the probabilistic assumptions in the model. In contrast, the system of Brandenburger and Yanofsky [15] involves a simplifying assumption that the state of the system, \(\lambda \), is a discrete random variable with finitely many outputs. With this assumption, notions such as “locality” and “\(\lambda \)-independence” are easier to formulate and easier to understand. Though the finite-\(\lambda \) assumption restricts the type of theories one can model, this is not as egregious as it might seem: in some hidden-variable situations, any possible correlation scenario can be modeled by a finite-output \(\lambda \), as discussed in [6].

In this appendix, we investigate what happens to the system of Sect. 2 if we make the additional assumption that \(\lambda \) is a discrete random variable with finitely many outputs. This will allow us to directly compare our system to the system of [15], as well as to illustrate and clarify the nature of our particular choices of assumptions.

Before restricting ourselves to the finite-\(\lambda \) setting, we can show that, working in the system of Sect. 2, we can derive the following alternate version of (4):

Experimental Assumption 3*:

$$\begin{aligned} (A, B) \perp \!\!\!\perp \lambda . \end{aligned}$$
(45)

The above alternative version of (4) is more similar to the “\(\lambda \)-independence” assumption as formulated in [15]. (45) is also a stronger assumption than (4): one can verify that (45) directly implies (4), but the converse does not hold. However, if we also assume (2) and (5), we can derive (45) from (4).

Proposition 5

The condition (4), in conjunction with (2) and (5), implies (45).

Proof

We show that \(P\big ( (A, B) = (a, b) | \lambda \big ) = P\big ( (A, B) = (a, b)\big )\); the proof for the three other cases \((a, b')\), \((a', b)\), and \((a', b')\) is the same. We have

$$\begin{aligned} P\big ( (A, B) \!=\! (a, b) | \lambda \big ) \!=\! \sum _{i=\pm 1}\sum _{j=\pm 1}P\big (\{A\!=\!a\}\cap \{B=b\} \cap (\{D_1 = i\} \cap \{D_2 \!=\! j\})\big |\lambda \big ), \end{aligned}$$

by Lemma 2. Applying (5) to the above expression yields

$$\begin{aligned} \sum _{i=\pm 1}\sum _{j=\pm 1} P\big (\{D_1=i\}\cap \{A=a\}\big |\lambda \big )\cdot P\big (\{D_2=j\}\cap \{B=b\}\big |\lambda \big ). \end{aligned}$$

Factoring the expression and applying Lemma 2, we obtain

$$\begin{aligned}&P\big (+_2\cap \{B=b\}\big |\lambda \big )\bigg [ P\big (+_1\cap \{A=a\}\big |\lambda \big ) + P\big (-_1\cap \{A=a\}\big |\lambda \big )\bigg ] \\&\quad + P\big (-_2\cap \{B=b\}\big |\lambda \big )\bigg [ P\big (+_1\cap \{A=a\}\big |\lambda \big ) + P\big (-_1\cap \{A=a\} \big |\lambda \big )\bigg ] \\&\qquad = P\big (+_2\cap \{B=b\}\big |\lambda \big )\cdot P(A=a|\lambda ) +P\big (-_2\cap \{B=b\}\big |\lambda \big )\cdot P(A=a|\lambda ) \\&\qquad =P(A = a|\lambda )P(B = b|\lambda ). \end{aligned}$$

Now, by applying (4) and then (2), we have

$$\begin{aligned} P(A = a|\lambda )P(B = b|\lambda ) = P(A = a)P(B = b) = P\big ( (A, B) = (a, b) \big ). \end{aligned}$$

\(\square \)

Proposition 5 rules out the possibility of \(\lambda \) having some dependence on the joint distribution of \(A\) and \(B\).

Moving forward, we now make the assumption that the random variable \(\lambda \), introduced in Sect. 2, is of the form

$$\begin{aligned} \lambda : \Omega \rightarrow \Lambda = \{l_1, \ldots , l_n\}, \end{aligned}$$
(46)

so \(\Lambda \) is now taken to be a finite set containing \(n\) elements. (The nature of its constituent elements \(l_i\) is not characterized, or important.) We also assume that for all \(i\), \(P(\lambda = l_i)> 0\); any zero-probability event has no observable effect on the behavior of the model, so we remove such events from consideration. Henceforth we will use \(l_i\) to refer to the event \(\{\lambda = l_i\} = \lambda ^{-1}(l_i)\subseteq \Omega \), unless doing so could create ambiguity.

With the assumption that \(\lambda \) is finite, the expression (4) is now equivalent to

$$\begin{aligned}&\forall i\in \{1, \ldots , n\}, \mathbf{a} \in \{a, a'\}, \,\, \mathrm{and } \,\, \mathbf{b} \in \{b, b'\},\\&P(A=\mathbf{a}\cap l_i) = P(A=\mathbf{a})P(l_i) \quad \mathrm{and }\\&P(B=\mathbf{b}\cap l_i) = P(B=\mathbf{b})P(l_i). \end{aligned}$$

The interpretation of (5) is simplified as well. This is because for any event \(E\), \(P(E|\lambda )\) will now just be a simple function that is equal to \(P(E|l_i)\) on each set \(l_i\). Now (5) is equivalent to the condition

$$\begin{aligned}&\forall i\in \{1,\ldots ,n\}, \mathbf{a} \in \{a, a'\}, \mathbf{b} \in \{b, b'\}, j_a \in \{+_1, -_1\}, \,\, \mathrm{and} \,\, k_b\in \{+_2, -_2\},\nonumber \\&P(j_a, \mathbf{a}, k_b, \mathbf{b}|l_i) = P(j_a, \mathbf{a}|l_i)P(k_b, \mathbf{b}|l_i). \end{aligned}$$
(47)

Note that the conditionals in (47) are events, not random variables, resulting in a simpler construction when compared to expressions like (6). So now, all of the assumptions (2)–(5) can be expressed in terms of elementary probabilistic statements concerning a finite collection of events.

We now show that the axiomatization of Sect. 2 is essentially equivalent to the axiomatization of [15], when applied to the relevant experimental setup (i.e., an experiment with two detectors, two detector settings, and two outcomes). To do this, note that if we assume (2), (3), and replace (4) with the stronger (45), then the following statement,

$$\begin{aligned} \forall i, \mathbf{a}, \mathbf{b}, j_a, k_b,\quad {P(j_a, \mathbf{a}, k_b, \mathbf{b},l_i)\over P(l_i)} = {P(j_a, \mathbf{a},l_i)P(k_b, \mathbf{b},l_i)\over P(l_i)^2} \end{aligned}$$
(48)

is equivalent to

$$\begin{aligned} \forall i, \mathbf{a}, \mathbf{b}, j_a, k_b, \quad {P(j_a, \mathbf{a}, k_b, \mathbf{b},l_i)\over P(\mathbf{a}, \mathbf{b}, l_i)} = {P(j_a, \mathbf{a},l_i)P(k_b, \mathbf{b},l_i)\over P(\mathbf{a}, l_i)P(\mathbf{b}, l_i)}. \end{aligned}$$
(49)

Demonstrating the above biconditional is a straightforward exercise. Note that (48) is equivalent to (47). Now, recall that by Proposition 5, we have the following logical relationship between assumptions,

$$\begin{aligned} (2), (3), (4), (5) \quad \Leftrightarrow \quad (2), (3), (45), (5), \end{aligned}$$

and in the finite-\(\lambda \) setting, we have (5) \(\Leftrightarrow \) (47), so we can say that

$$\begin{aligned} (2), (3), (45), (5) \quad \Leftrightarrow \quad (2), (3), (45), (49). \end{aligned}$$

The collection of assumptions on the right side of the above equivalence is closely related to the framework of [15] as it would apply to the 2-detector, 2-setting, 2-outcome scenario. (45) is equivalent to Definition 2.4 (“\(\lambda \)-independence”) in [15], and (49) is equivalent to Definition 2.10 (“locality”). So in a finite-\(\lambda \) setting, our framework—i.e., the set of conditions (2)–(5)—is equivalent to the Brandenburger/Yanofsky framework applied to a 2-dector, 2-setting, 2-outcome scenario where measurement choices are independent from each other (the condition (2)) and none of the measurement settings have trivial probabilities (the condition (3)).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Bierhorst, P. A Rigorous Analysis of the Clauser–Horne–Shimony–Holt Inequality Experiment When Trials Need Not be Independent. Found Phys 44, 736–761 (2014). https://doi.org/10.1007/s10701-014-9811-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10701-014-9811-3

Keywords

Mathematics Subject Classification

Navigation