1 Introduction

Contextuality formalized in the form of violation of noncontextuality inequalities, Bell-contextuality [1, 2], is a hot topic in quantum physics (see, e.g., [3, 4] and references herein). Unfortunately, it is typically presented in the mathematical framework and its physical meaning is unclear.

We stress that, in fact, one has to distinguish two different notions of contextuality, Bohr-contextuality and Bell-contextuality. In this paper, we consider the latter, see Appendix 2 for for the former. Therefore we shall speak simply about contextuality (having in mind Bell-contextuality).

We point out that discussions on this sort contextuality are typically started with heuristically more attractive definition going back to Bell [1, 2]. It can be called explicit contextuality: if A,B,C are three quantum observables, such that A compatible with B and C, a measurement of A might give different result depending upon whether A is measured with B or with C. However, for incompatible observables B and C, this statement is not testable experimentally.Footnote 1 Therefore we proceed with Bell-contextuality that is straightforwardly coupled to experiment. (This approach to contextuality was actively driven by Adan Cabello; so it may be natural to call it Bell-Cabello contextuality.)

It is easy to show that, for quantum observables, there is no contextuality without incompatibility: for compatible observables, it is impossible to violate any noncontextuality inequality (Theorem 1, Section 2). The natural question arises:

Has contextuality without incompatibility any physical meaning?

Generally this is the very complex question. I do not know the answer to it for general noncontextuality inequalities. And I hope that this paper would stimulate foundational research in this direction. We concentrate on contextuality for four quantum observables - noncontextuality analog of the CHSH-inequality.

We proved that, for “natural quantum observables” , contextuality is reduced to incompatibilityFootnote 2 (in [9,10,11], the same conclusion was obtained for quantum nonlocality, cf. [12,13,14,15,16]).

At the same time, we shown that generally contextuality without incompatibility may have some physical content. We found a mathematical constraint extracting the contextuality component from incompatibility. However, the physical meaning of this constraint is not clear.

We also remark that there exist positive answers to the inverse question: there can be (non-quantum, quasi-classical) incompatibility without contextuality; as exposed by finite automata [21] as well as for generalized urn models [22], see [23].

In appendix, we briefly discuss another sort of contextuality that is understood more generally in accordance with the Bohr message [24] that all experimental arrangement (experimental context) has to be taken into account in the process of measurement; he pointed to “the impossibility of any sharp separation between the behavior of atomic objects and the interaction with the measuring instruments which serve to define the conditions under which the phenomena appear.” Here “phenomenon” is understood as the individual output of measurement [10, 24,25,26]. Bohr-contextuality is the basis of quantum foundations. Moreover, this is the root of the Bohr’s complementarity principle. Incompatibility is a consequence of contextuality, but the latter has to be understood as Bohr-contextuality, Appendix 1 (see also [27]).

2 Quantum Theory: Bell-contextuality vs. Bohr-incompatibility

In this paper, we consider dichotomous observables taking values ± 1.

We follow paper [4] (one of the best and clearest representations of contextuality). Consider a set of observables {X1,...,Xn}; a context C is a set of indexes such that Xi,Xj are compatible for all pairs i,jC. A contextuality structure for these observables is given by a set of contexts \(\mathcal {C}=\{ C \},\) or simply the maximal contexts. For each context C, we measure pairwise correlations for observables Xi and Xj with indexes i,jC as well as averages 〈Xi〉 of observables Xi. The n-cycle contextuality scenario is given by n observables X1,...,Xn and the set of maximal contexts

$$ \mathcal{C}_{n}=\{\{X_{1},X_{2}\}, . . . ,\{X_{n-1},X_{n}\},\{X_{n},X_{1}\}\}. $$
(1)

Statistical data associated with this set of contexts is given by the collection of averages and correlations:

$$ \{\langle X_{1}\rangle,...., \langle X_{n}\rangle; \langle X_{1} X_{2}\rangle, . . . , \langle X_{n-1}X_{n}\rangle, \langle X_{n},X_{1}\rangle\}. $$
(2)

Theorem 1 from paper [4] describes all tight noncontextuality inequalities. In particular, for n = 4 we have inequality:

$$ | \langle X_{1} X_{2} \rangle + \langle X_{2} X_{3} \rangle + \langle X_{3} X_{4} \rangle - \langle X_{4} X_{1}\rangle | \leq 2 . $$
(3)

Theorem 2 [4] demonstrates that, for n ≥ 4 (cf. Appendix 1 for n = 3), aforementioned tight noncontexuality inequalities are violated by quantum correlations. But,

what is the physical root of quantum violations?

Unfortunately, the formal mathematical calculations [4] used to show violation of noncontextuality inequalities for quantum observables do not clarify physics behind these violations.

Let us turn to the quantum physics, i.e., X1,...,Xn are not arbitrary observables, but quantum physical ones. In the quantum formalism, they are represented by Hermitian operators \(\hat {X}_{1}, . . . , \hat {X}_{n}.\) Denote the orthogonal projectors onto the corresponding eigenspaces by the symbols \(\hat E_{j \alpha }, \alpha = \pm 1.\)

Suppose now that these observables are compatible with each other, i.e., any two observables Xi,Xj can be jointly measurable, so in the operator formalism, \([\hat {X}_{i},\hat {X}_{j}]=0.\) The quantum theory has one amazing feature that is not so widely emphasized:

Pairwise joint measurability implies k-wise joint measurability for any kn.

If all pairs can be jointly measured, then even any family of observables \(\{ X_{i_{1}},..., X_{i_{k}} \}\) can be jointly measured as well. In principle, there is no reason for this. This is the specialty of quantum theory.

The joint probability distribution (JPD) of compatible observables is defined by the following formula [28]:

$$ P_{i_{1}...i_{k}} (\alpha_{i_{1}},...,\alpha_{i_{k}}) = \text{Tr} \rho \hat E_{i_{1} \alpha_{i_{1}}} {\cdots} \hat E_{i_{k} \alpha_{i_{k}}}. $$
(4)

In particular, by setting k = n we obtain JPD of all observables,

$$ P_{1...n} (\alpha_{1},..., \alpha_{n})=\text{Tr} \rho \hat E_{1 \alpha_{1}} {\cdots} \hat E_{n \alpha_{n}}. $$
(5)

We remark that the probability distributions given by (4) can be obtained from the latter JPD as the marginal probability distributions:

$$ P_{i_{1}...i_{k}} (\alpha_{i_{1}},...,\alpha_{i_{k}})= \underset{\alpha_{j}, j\not= i_{1}...i_{k}}{\sum} P_{1...n} (\alpha_{1},..., \alpha_{n}). $$
(6)

This formula implies as well that the marginals of JPD \(P_{i_{1}...i_{k}}\) of the rank k generate JPDs of the rank k − 1. In particular, we have the consistency rules for JPDs of ranks 2 and 1,

$$ P_{i} (\alpha_{i})= \underset{\alpha_{j}}{\sum} P_{ij} (\alpha_{i},\alpha_{j}) $$
(7)

(in quantum physics, this condition is known as no signaling), and ranks 3 and 2 consistency:

$$ P_{ij} (\alpha_{i}, \alpha_{j})= \underset{\alpha_{k}}{\sum} P_{ijk} (\alpha_{i},\alpha_{j}, \alpha_{k}) $$
(8)

We have the classical probability framework; the Kolmogorov probability model with the probability measure PP1...n. In this classical probabilistic framework we can prove any noncontextuality inequality (any Bell-type inequality, cf. [29,30,31,32,33,34,35,36], [26, 27]). It is impossible to violate them for compatible quantum observables. We can formulate this result as a simple mathematical statement:

Theorem 1

For quantum observables X1,...,Xn, (Bell-)contextuality implies incompatibility of at least two of them.

Thus, there is no Bell-contextuality without incompatibility. Does the latter contain something more than incompatibility?

Finally, we remark that noncontextuality inequalities started to be used in applications outside of physics, e.g., in psychology, cognitive science, and decision making [17,18,19,20]. If one does not assume that observables are represented by Hermitian operators in Hilbert space, then “no-go” Theorem 1 loses its value.

3 Is Contextuality Reduced to Incompatibility?

In [9], I analyzed in details the CHSH-inequality; the CHSH-correlation has the form:

$$ {\Gamma} =\langle A_{1} B_{1} \rangle + \langle A_{1} B_{2} \rangle + \langle A_{2} B_{1} \rangle- \langle A_{2} B_{2} \rangle, $$
(9)

where observables Ai are compatible with observables Bj,i,j = 1,2. In [9], the tensor product structure of the state space was not explored and quantum observables were represented by Hermitian operators \(\hat A_{i}, \hat B_{j}\) acting an arbitrary Hilbert space. In this framework the CHSH-inequality can be treated as the noncontextuality inequality for four observables; by setting in (9) A2 = X1,B1 = X2,A1 = X3,B2 = X4, we obtain the correlation:

$$ {\Gamma} = \langle X_{1} X_{2} \rangle + \langle X_{2} X_{3} \rangle + \langle X_{3} X_{4} \rangle - \langle X_{4} X_{1} \rangle, $$
(10)

since we work with quantum observables, we proceed under the compatibility assumption

$$ [\hat{X}_{1}, \hat{X}_{2}]=0, [\hat{X}_{3}, \hat{X}_{2}]=0, [\hat{X}_{3}, \hat{X}_{4}]=0, [\hat{X}_{1}, \hat{X}_{4}]=0. $$
(11)

Now set

$$ \hat M_{13}= i [\hat{X}_{1}, \hat{X}_{3}] \text{and} \hat M_{34}= i [\hat{X}_{2}, \hat{X}_{4}]. $$
(12)

These are Hermitian operators, so they represent some quantum observables M13 and M34. We remark that these observables are compatible:

$$ [\hat M_{13}, \hat M_{34}] =0. $$
(13)

The following theorem is the noncontextuality reinterpretation of the main result of paper [9]:

Theorem 2

Condition

$$ \hat M_{13} \circ \hat M_{34} \not= 0. $$
(14)

is necessary and sufficient for violation of the noncontextuality inequality (3) for some quantum state.

Proof’s scheme. Consider the operator

$$ \hat {\Gamma} = \hat{X}_{1} \hat{X}_{2} + \hat{X}_{2} \hat{X}_{3} + \hat{X}_{3} \hat{X}_{4} - \hat{X}_{4} \hat{X}_{1}. $$
(15)

Then we have

$$ \hat {\Gamma}^{2} = 4 + [\hat{X}_{1}, \hat{X}_{3}] [\hat{X}_{2}, \hat{X}_{4}] = 4 + \hat M_{13} \hat M_{34}. $$
(16)

Then it is easy to show that \(\| \hat {\Gamma }^{2} \| >4,\) if and only if condition (14) holds. Finally, we note that

$$ \underset{\| \psi\|=1}{\sup} | \langle \psi| \hat {\Gamma}| \psi \rangle = \| \hat {\Gamma} \| = \sqrt{\| \hat {\Gamma}^{2} \|}. $$
(17)

We remark that condition (14) is trivially satisfied for incompatible observables, if the state space and observables have the tensor product structure: H = H13H24 and

$$ \hat{X}_{i}=\hat{\textbf{X}}_{i} \otimes I, \hat{X}_{j}=I\otimes \hat{\textbf{X}}_{j}, $$
(18)

where

$$ \hat{\textbf{X}}_{i} : H_{13} \to H_{13}, i=1,3, \hat{\textbf{X}}_{j}: H_{24} \to H_{44}, j=2,4. $$
(19)

Here condition (14) is reduced to incompatibility condition:

$$ [\hat{\textbf{X}}_{i}, :\hat{\textbf{X}}_{j}] \not= 0, i=1,3; j=2,4. $$
(20)

In particular, for compound systems, contextuality (“nonlocality”) is exactly incompatibility. The same is valid for any tensor decomposition of the state space of a single quantum system with observables of the type (18). In the tensor product case, contextuality without incompatibility leads to the notion with the empty content.

But, it may happen that Xi-observables, i = 1,3, and Xj-observables, j = 2,4, are not connected via the tensor product structure. In this case, the interpretation of constraint (14) is nontrivial. What is its physical meaning? I have no idea.

Of course, the main problem is that it is not clear at all how to measure the observables of the commutator-type.

4 Conclusion

In quantum physics, there is no contextuality without incompatibility. This is well known, but not so highly emphasized feature of quantum observables.

For fourth quantum observables, these two notions coincide under validity of constraint (14). If it is violated, then, for such observables, there is still a hope that quantum contextuality without incompatibility has some nontrivial physical meaning. (What?) The problem of nontrivial physical meaning of “pure contextuality”, i.e., one distilled from incompatibility, for n > 4 observables (as well as n = 3, see Appendix 2) is open.

Finding the right physical interpretation for contextuality beyond incompatibility is important for demystification of quantum physics (cf. with discussion of Svozil [23] on “quantum focus pocus”).