From no-signalling to quantum states
Abstract
Characterising quantum correlations from physical principles is a central problem in the field of quantum information theory. Entanglement breaks bounds on correlations put by Bell’s theorem, thus challenging the notion of local causality as a physical principle. A natural relaxation is to study no-signalling as a constraint on joint probability distributions. It has been shown that when considered with respect to so-called locally quantum observables, bipartite non-signalling correlations never exceed their quantum counterparts; still, such correlations generally do not derive from quantum states. This leaves open the search for additional principles which identify quantum states within the larger set of (collections of) non-signalling joint probability distributions over locally quantum observables. Here, we suggest a natural generalisation of no-signalling in the form of no-disturbance to dilated systems. We prove that non-signalling joint probability distributions satisfying this extension correspond with bipartite quantum states up to a choice of time orientation in subsystems.
I Introduction
Given the violation of Bell inequalities AspectEtAl1981 ; ShalmEtAl2015 ; ZeilingerEtAl2015 (see also, WuShaknov1950 ) and the paramount importance of entanglement for quantum information theory Shor1994 ; RaussendorfBriegel2001 ; AcinGisinMasanes2006 ; WisemanJonesDoherty2007 ; PawlowskiBrunner2011 ; deVicente2014 ; BravyiGossetKoenig2020 ; google2020 , characterising quantum correlations constitutes an important and ongoing research objective PawlowskiEtAl2009 ; CabelloSeveriniWinter2010 ; vanDam2013 ; FritzEtAl2013 ; JanottaHinrichsen2014 ; Popescu2014 . One line of this research has focused on the fact that while quantum mechanics is hardly reconcilable with the classical notion of local causality Bell1976 , quantum correlations obey the more general principle of no-signalling: the joint probability distributions for different local measurements with respective outcome sets , marginalise to the same local distributions,
(1) |
Importantly, Eq. (1) depends on the set of local measurements, consequently one must specify the possible choices of measurements on either subsystem in order to evaluate the constraints inherent to no-signalling. For instance, the PR-box correlations (for the CHSH scenario) restrict to just two measurements on either side PopescuRohrlich1994 . A physically more interesting scenario is that in which arbitrary local measurements are allowed. Remarkably, one can show that this already bounds the correlations to be quantum in the bipartite case BarnumEtAl2010 ; AcinEtAl2010 . Nevertheless, the collections of non-signalling distributions over the set of locally quantum observables do not correspond with quantum states KlayRandallFoulis1987 ; Wallach2002 . There are more non-signalling distributions than quantum states, i.e., while the correlations are as strong as quantum ones, the underlying distributions need not derive from a quantum state.
In this paper we identify two physical principles that characterise those non-signalling bipartite correlations which correspond with quantum states. To set the stage, we review some basic facts about correlations over non-signalling correlations in Sec. II. In Sec. III we reformulate the problem from the perspective of contextuality and identify no-disturbance as the key underlying principle. This subsumes no-signalling, but makes explicit the intimate relationship with non-contextuality. Based on this, we suggest an extension of the no-disturbance principle to dilated systems in Sec. IV. Thm. 2 shows that this strengthened principle almost singles out quantum states. In Sec. V we identify the missing piece of data by introducing a notion of time orientation. Our main result, Thm. 3, proves that under a related consistency condition with respect to unitary evolution in subsystems correlations indeed derive from quantum states. Sec. VI concludes.
II No-signalling and locally quantum observables
Throughout, we denote by the algebra of linear operators on some finite-dimensional Hilbert space , by the lattice of projections on , and by the real-linear space of self-adjoint (Hermitian) operators, representing the observables of a system.
In contrast to restricted sets of observables such as those in PopescuRohrlich1994 , BarnumEtAl2010 study no-signalling for all locally quantum observables. Locally quantum here means that every local system is described in terms of an observable algebra of self-adjoint (Hermitian) operators corresponding to the respective quantum system described by the Hilbert space . However, rather than assuming the tensor product structure between Hilbert spaces, the composite system is described solely in terms of product observables, i.e., pairs where for . Using the spectral decomposition of observables for and , bipartite correlations arise from measures , (here, denotes the identity matrix). For the latter the no-signalling constraints in Eq. (1) read:
(2) |
for all mutually orthogonal sets of projections , i.e., and .111For convenience, we restrict ourselves to (measures over) projective measurements. It is straightforward to extend the discussion to positive operator-valued measures (POVM), equivalently effect spaces. This allows to include the two-dimensional case in Thm. 1 (see BarnumEtAl2010 ; FrembsDoering2019b ). For later reference, we point out that the structure of observables implies that (in addition to no-signalling) is also independent of contexts, in a sense to be made precise in Sec. III.
Applying Gleason’s theorem Gleason1975 to the respective subsystems one shows that extends to a linear functional , equivalently, that there exists a linear operator such that for every . Since is positive, is positive on all product observables with and . We call such normalised, i.e., , linear functionals positive on pure tensor (POPT) states.
Theorem 1.
KlayRandallFoulis1987 ; BarnumEtAl2010 Let , be Hilbert spaces with finite. There is a one-to-one correspondence between measures , which satisfy the no-signalling constraints in Eq. (2), and POPT states .
Crucially, a POPT state is generally not positive. Hence, defines a quantum state if and only if is positive. Despite this discrepancy, BarnumEtAl2010 ; AcinEtAl2010 show that the correlations arising from POPT states do not exceed those arising from quantum states in the bipartite case.
The existence of unphysical POPT states gives rise to the problem of finding a sharper classification that rules out such states. This paper offers a solution to this problem. To this end, we first argue that rather than no-signalling, the natural principle underlying Eq. (2) is no-disturbance. We then extend the scope of the no-disturbance principle to dilations of local systems, and show that up to a consistency condition with respect to unitary evolution in subsystems this establishes a one-to-one correspondence between non-signalling joint probability distributions and bipartite quantum states.
We remark that distributions over product observables have been studied before, e.g. in KlayRandallFoulis1987 , out of the attempt to define a tensor product intrinsic to quantum logic.222For a recent contribution to this problem, see AbramskyBarbosa2021 (and references therein). We do not consider this problem here, but note that our result has the potential to give new impulse to this research programme. Another perspective on this problem is given by Wallach in Wallach2002 , who considers a generalisation of Gleason’s theorem to composite systems. Since it distracts from the physical significance of the present discussion, we study this problem in more generality elsewhere FrembsDoering2022b .
III Non-contextuality and no-disturbance
A crucial feature of locally quantum non-signalling joint probability distributions is that they satisfy a more general constraint than no-signalling, called no-disturbance. We introduce this notion in the following sections, before extending the no-disturbance principle to dilations in Sec. IV. Our approach naturally embeds within the analysis of the principal role of contextuality in quantum theory FreDoe19a . For a recent review of the wider subject, see Masse2021 .
III.1 Non-contextuality and marginalisation constraints
Contextuality is a key principle in foundations, separating quantum from classical physics Specker1960 ; KochenSpecker1967 ; IshamButterfieldI ; DoeringIsham2011 ; AbramskyBrandenburger2011 ; CabelloSeveriniWinter2010 . Moreover, it has been shown to be a resource for quantum computation in various architectures Howard2014 ; Raussendorf2016 ; BravyiGossetKoenig2020 .
At the core of contextuality lies the notion of simultaneous measurability, which equips the set of observables with a reflexive, symmetric, but generally non-transitive relation Specker1960 . We call any subset of simultaneously measurable observables a context. The set of all contexts carries an intrinsic order relation arising from coarse-graining on outcomes of observables. The resulting partially ordered set is called the partial order of contexts. For the close connection between contextuality in this sense and the non-existence of valuation functions in the sense of Kochen and Specker KochenSpecker1967 , we refer to FreDoe19a .
In quantum theory, two observables are simultaneously measurable if and only if they commute. Consequently, contexts are given by commutative subalgebras , which are ordered by inclusion into the corresponding partial order of contexts denoted by . In this setup a quantum state becomes a collection of probability distributions , one for every context. Moreover, coarse-graining constrains these across contexts: let , be probability distributions in contexts such that , and denote by the inclusion relation between their respective projections, then is obtained from by marginalisation,
(3) |
While individually inconspicuous, marginalisation constraints impose a strong condition in conjunction with non-contextuality: probabilities of events corresponding to an observable are independent of other observables , , i.e., they are independent of context:
(4) |
This non-contextuality assumption on probability distributions is at the heart of Gleason’s theorem Gleason1975 ; for a reformulation of the latter in this language see Doering2008 ; FreDoe19a ; FrembsDoering2022b .
III.2 No-signalling from no-disturbance
There is a natural notion of composition for (partial orders of) contexts: their canonical product, denoted , is the Cartesian product with elements for , and order relations such that for all , ,
(5) |
Restricted to product contexts, Eq. (4) says that for all and
(6) |
We call the collection of constraints over all contexts in Eq. (6) the no-disturbance principle.333The term appears in RamanathanEtAl2012 , but has been used under different names before, e.g. in CabelloSeveriniWinter2010 (see also, DoeringIsham2011 ; FreDoe19a ). Note that no-disturbance reduces to no-signalling in Eq. (1) when restricted to the trivial contexts 444The respective contexts correspond to the trivial events of observing that there is a local system. on the respective local subsystem,
(7) |
Observe that general non-signalling joint probability distributions over locally quantum observables correspond with collections of probability distributions satisfying Eq. (7). The latter are more general than the non-signalling measures in Eq. (2), which satisfy the more restrictive no-disturbance condition in Eq. (6). In this sense, no-disturbance is already implicit in BarnumEtAl2010 . With the latter, we re-emphasise that is a meagre part of the full quantum structure of the composite Hilbert space . In particular, the product system is not described by the tensor product of the individual Hilbert spaces.
Having established the crucial role of non-contextuality and no-disturbance, in the next section we extend this principle to dilations on local subsystems.
IV No-disturbance for dilations
A key idea of this work is to impose the no-disturbance condition in Eq. (6) not only between locally quantum observables, but to extend its scope to dilations of (at least) one of the two subsystems. We give some motivational background for this step first.
Recall that by Thm. 1 a measure defines a POPT state with corresponding linear operator . Similarly, for every we define a positive linear operator from for all . It follows from Thm. 1 that extends to a positive linear map . On the other hand, for any single we may assume to arise via coarse-graining from some on a larger system (to be thought of as the system together with an (environmental) ancillary system), by tracing out the extra degrees of freedom:
(8) |
is a purification of , where the unitary is chosen to separate the input from the ancillary system. We want to arrange the purifications for all in an economical way. To start with, note that since for every , defines a non-normalised () positive operator-valued measure (POVM) on for every set of mutually orthogonal projections in , i.e., for every context . We may thus write
(9) |
for all and , mimicking the purification in Eq. (8). Eq. (9) is called a dilation of the POVM to the projection-valued measure (PVM) . More generally, by Naimark’s theorem Naimark1943 ; Stinespring1955 every POVM admits a dilation to a PVM such that , where is a linear map.555Here, we concentrate the dependence on contexts in the mapping , by choosing sufficiently large Naimark1943 ; Stinespring1955 ; Choi1975 and by absorbing any context dependence on into for every .
Importantly, the dilations generally depend on contexts . In contrast, recall that no-disturbance in Eq. (6) encodes non-contextuality constraints between product observables. By comparison, this suggests to extend the no-disturbance principle also to product observables with respect to the dilations .
Definition 1.
We say that the measure satisfies no-disturbance for dilations if for some Hilbert space , linear map , and projection-valued measures , such that
(10) |
satisfies the no-disturbance principle in Eq. (6) for all product contexts in .
From a physical point of view, we interpret the (conditional) probability distributions in contexts as states of incomplete information. We have seen that this interpretation arises naturally from the explicit dilations in Eq. (9), which express the state as arising from coarse-graining of ancillary degrees of freedom. This view is further corroborated if one allows not only projective measurements on the respective subsystems and , but arbitrary positive operator-valued measures, e.g. as discussed in BarnumEtAl2010 .
Put the other way around, if did not satisfy no-disturbance for dilations in Def. 1—while potentially being non-disturbing (non-signalling) when restricted to the system —it would fail to be non-disturbing (non-signalling) for every choice of dilations to a larger system. Such measures would at the very least prompt a substantial revision of the concept of mixed states in terms of states of information in quantum theory. As such, the extension to dilated systems in Def. 1 is a natural one, and arguably conservative compared with other related approaches, e.g. PawlowskiEtAl2009 ; vanDam2013 ; FritzEtAl2013 ; Popescu2014 .
Note also that Def. 1 does not change the fact that the composite system is described by means of the Cartesian product of contexts in Eq. (5), not the tensor product.
In order to state the implications of Def. 1, we remind the reader of some basic facts about Jordan algebras. Recall that the set of self-adjoint (Hermitian) matrices is closed under the anti-commutator for all . This defines the (special) Jordan algebra .666A Jordan algebra is called special if it is isomorphic to the subalgebra of the self-adjoint part of an associative algebra. Otherwise, it is called exceptional. Moreover, we obtain a Jordan -algebra by extending the operation to the complexified algebra . This expresses the fact that is the self-adjoint part of . Finally, a Jordan -homomorphism is a linear map such that and for all .
We have the following key result, which extracts Jordan structure from Def. 1.
Theorem 2.
Proof.
Since satisfies no-disturbance for dilations in Def. 1, there exists a Hilbert space , a linear map , and projection-valued measures such that for all , and . Moreover, the extension in Eq. (10), where , , and , satisfies the no-disturbance principle in Eq. (6). The latter are equivalent to the non-contextuality constraints in Eq. (4), restricted to product contexts . Consequently, does not depend on contexts, and therefore defines a map . It follows that is an orthomorphism, i.e., (i) , (ii) , (iii) and (iv) all hold whenever , . This follows since by definition is a projection-valued measure in every context . In particular, note that conditions (i)-(iii) hold whenever , , which implies for some . Finally, by a result due to Bunce and Wright (Cor. 1, BunceWright1993 ) every orthomorphism lifts to a Jordan -homomorphism as desired. ∎
Of course, every quantum state has a purification which yields a dilation of the form in Thm. 2. Our argument works in the reverse direction: by requiring that measures have a non-contextual extension, i.e., that they satisfy the no-disturbance principle for at least one choice of dilations , is of the form in Thm. 2. Next, we show that, in this case, already defines a quantum state up to a choice of time orientation in local subsystems.
V Unitary evolution and time-oriented states
Extending no-disturbance to dilations via Def. 1 is not quite sufficient to ensure that a POPT state becomes positive and thus a quantum state—but it almost is. The difference is best expressed in terms of the Choi-Jamiołkowski isomorphism Jamiolkowski1972 . In particular, by Choi’s theorem the linear operator is positive if and only if a related map is completely positive Choi1975 . Moreover, by Stinespring’s theorem a map is completely positive if and only if it corresponds with an algebra homomorphism (a representation) on a larger system Stinespring1955 , i.e., is completely positive if and only if there exists a Hilbert space , a linear map , and an algebra homomorphism such that . By Thm. 2, if satisfies no-disturbance for dilations, it is of this form with the only difference that is a merely Jordan -homomorphism.
The Jordan algebra does not completely determine the algebra , since it lacks the antisymmetric part or commutator in the associative (and noncommutative for ) product for all . In particular, the Jordan -homomorphism is an algebra homomorphism if and only if it also preserves commutators, i.e., for all . This property has a distinctive physical meaning in terms the unitary evolution in local subsystems.
To see this, we describe the dynamics of a system represented by the observable algebra in terms of one-parameter groups of automorphisms . Note that possesses a canonical action on itself by conjugation , for all .777The map ‘selects’ unitary (as opposed to anti-unitary Wigner_GruppentheorieUndQM ; Bargmann1964 ) symmetries on . Inherent in this is an identification of self-adjoint elements in and generators of symmetries, called a dynamical correspondence. In other words, ‘selects’ a canonical dynamical correspondence on : is the exponential of the (canonical) dynamical correspondence on . For more details, see AlfsenShultz1998 ; Doering2014 . The latter expresses the unitary evolution in the system, thus promoting the parameter to a time parameter with playing the role of a Hamiltonian. Note however that without fixing a preferred Hamiltonian the value of has a priori no objective physical meaning, it is intrinsically relational. Nevertheless, the sign of turns out to be independent of (the choice of Hamiltonian) AlfsenShultz1998 .
By differentiation, , where is the commutator in the ambient algebra . It follows that preserves commutators if and only if it preserves the canonical unitary evolution of the subsystem algebras. We call the canonical time orientation on AlfsenShultz1998 ; Doering2014 ; Frembs2022a ; Frembs2022b , and (cf. Prop. 15 in AlfsenShultz1998a ) the reverse time orientation, which corresponds with the opposite order of composition in , equivalently the opposite sign for the commutator in the respective algebras,
where we set for all . Of course, .
Definition 2.
Two remarks on Def. 2 are in order. First, note that Eq. (11) requires consistency with respect to the commutators on and . However, since arises from by restriction, the time orientation on extends to a unique time orientation on . Second, we have used the reverse time orientation on the first system. This choice of relative time orientation is intrinsic to Choi’s theorem Choi1975 (see also Frembs2022b ), which allows us to deduce positivity of from complete positivity of in Thm. 2.
Theorem 3.
Let , be Hilbert spaces with finite, and let be time-oriented as by Def. 2 (and thus satisfy no-disturbance for dilations). Then uniquely extends to a quantum state .
Proof.
Since satisfies no-disturbance for dilations in Def. 1, the dilations uniquely extend to a Jordan -homomorphism by Thm. 2. In particular, preserves anti-commutators. Moreover, since is time-oriented (see Def. 2), also preserves commutators between the algebras and , hence, is an algebra homomorphism. By Stinespring’s theorem, this implies that the linear map is completely positive. Finally, we define a linear operator by the relation for all . Note that is the inverse of the Choi-Jamiołkowski isomorphism Jamiolkowski1972 ; Choi1975 , defined on the computational basis in by
(12) |
where denotes the Hermitian adjoint (see also Frembs2022b ). Since is completely positive and since reverses time orientations by Eq. (11), it follows from Choi’s theorem Choi1975 that is positive. Hence, defines a quantum state. By construction, is the unique linear extension of , i.e., .
We finish this section with a few remarks. We defined the canonical time orientation on with respect to unitary symmetries. In turn, anti-unitary symmetries correspond with unitary symmetries on the algebra (and vice versa). Recalling that every anti-unitary operator is the product of a unitary and the time reversal operator further corroborates the notion of time-orientedness in Def. 2.
Moreover, note that Eq. (11) is invariant under changing time orientations, equivalently under time reversal in both subsystems (cf. Frembs2022b ). On the other hand, it follows from Thm. 3 that applying time reversal to one subsystem only will generally map outside the set of bipartite quantum states. This is reminiscent of the positive partial transpose (PPT) criterion for bipartite entanglement due to Peres Peres1996 ; Horodeckisz1996 . In fact, following this analogy the criterion can be given a sharp physical interpretation in terms of time orientations Frembs2022a .
Finally, note that Thm. 3 represents a generalisation of Gleason’s theorem to composite systems. We extend this perspective to the general setting of von Neumann algebras in FrembsDoering2022b .
VI Conclusion and Outlook
We studied the physical principle of no-disturbance on joint probability distributions over product observables, which reduces to no-signalling as shown in Eq. (7). As such we identified no-disturbance as the key principle in BarnumEtAl2010 , which shows that correlations over product observables satisfying no-disturbance cannot exceed bipartite quantum correlations. However, no-disturbance is not sufficient to restrict to quantum states. Our main result, Thm. 3, resolves this issue: by extending the scope of no-disturbance to dilations in Def. 1, and by enforcing a consistency condition with respect to unitary evolution in local subsystems in Def. 2, we show that the resulting joint probability distributions correspond with bipartite quantum states unambiguously.
Our research naturally relates to other approaches, which seek to single out quantum correlations from more general non-signalling distributions, by imposing additional principles, e.g. PawlowskiEtAl2009 ; vanDam2013 ; FritzEtAl2013 ; Popescu2014 . Apart from the extension of the no-disturbance principle to dilated systems, which fits nicely with the usual interpretation of mixed states as states of incomplete information as argued above, our only additional assumption already has a clear physical meaning in terms of the arrow of time Frembs2022a .
Acknowledgements.
This work is supported through a studentship in the Centre for Doctoral Training on Controlled Quantum Dynamics at Imperial College funded by the EPSRC, by grant number FQXi-RFP-1807 from the Foundational Questions Institute and Fetzer Franklin Fund, a donor advised fund of Silicon Valley Community Foundation, and ARC Future Fellowship FT180100317.
References
- (1) S. Abramsky and R. S. Barbosa, The logic of contextuality, in 29th EACSL Annual Conference on Computer Science Logic (CSL 2021), C. Baier and J. Goubault-Larrecq, eds., vol. 183 of Leibniz International Proceedings in Informatics (LIPIcs), Dagstuhl, Germany, 2021, Schloss Dagstuhl–Leibniz-Zentrum für Informatik, pp. 5:1–5:18.
- (2) S. Abramsky and A. Brandenburger, The sheaf-theoretic structure of non-locality and contextuality, New J. Phys., 13 (2011), p. 113036.
- (3) A. Acín et al., Unified framework for correlations in terms of local quantum observables, Phys. Rev. Lett., 104 (2010), p. 140404.
- (4) A. Acín, N. Gisin, and L. Masanes, From Bell’s theorem to secure quantum key distribution, Phys. Rev. Lett., 97 (2006), p. 120405.
- (5) E. M. Alfsen and F. W. Shultz, On orientation and dynamics in operator algebras. Part I, Communications in Mathematical Physics, 194 (1998), pp. 87–108.
- (6) E. M. Alfsen and F. W. Shultz, Orientation in operator algebras, Proc. Natl. Acad. Sci. U.S.A., 95 (1998), pp. 6596–6601.
- (7) F. Arute et al., Quantum supremacy using a programmable superconducting processor, Nature (London), 574 (2019).
- (8) A. Aspect et al., Experimental tests of realistic local theories via Bell’s theorem, Phys. Rev. Lett., 47 (1981), pp. 460–463.
- (9) V. Bargmann, Note on Wigner’s theorem on symmetry operations, Journal of Mathematical Physics, 5 (1964), pp. 862–868.
- (10) H. Barnum, S. Beigi, S. Boixo, M. B. Elliott, and S. Wehner, Local quantum measurement and no-signaling imply quantum correlations, Phys. Rev. Lett., 104 (2010).
- (11) J. S. Bell, The theory of local beables, in Speakable and unspeakable in quantum mechanics, J. S. Bell, ed., Cambridge University Press, 1987 1976, pp. 52–62.
- (12) S. Bravyi, D. Gosset, R. König, and M. Tomamichel, Quantum advantage with noisy shallow circuits, Nature Physics, 16 (2020), pp. 1040–1045.
- (13) A. Cabello, S. Severini, and A. Winter, (Non-)contextuality of physical theories as an axiom, 2010.
- (14) M. D. Choi, Completely positive linear maps on complex matrices, Linear Algebra Its Appl., 10 (1975), pp. 285 – 290.
- (15) J. I. de Vicente, On nonlocality as a resource theory and nonlocality measures, Journal of Physics A: Mathematical and Theoretical, 47 (2014), p. 424017.
- (16) A. Döring, Quantum states and measures on the spectral presheaf, Adv. Sci. Lett., 2 (2008).
- (17) A. Döring, Two new complete invariants of von Neumann algebras, ArXiv e-prints, (2014).
- (18) A. Döring and M. Frembs, Contextuality and the fundamental theorems of quantum mechanics, ArXiv e-prints, (2019).
- (19) A. Döring and C. Isham, “What is a Thing?”: Topos Theory in the Foundations of Physics, Springer Berlin Heidelberg, Berlin, Heidelberg, 2011, pp. 753–937.
- (20) M. Frembs, Entanglement and the arrow of time. (forthcoming), (2022).
- (21) , Variations on the Choi-Jamiołkowski isomorphism. (forthcoming), (2022).
- (22) M. Frembs and A. Döring, No-signalling, contextuality and the arrow of time, (2019).
- (23) M. Frembs and A. Döring, Gleason’s theorem for composite systems. (forthcoming), (2022).
- (24) T. Fritz et al., Local orthogonality as a multipartite principle for quantum correlations, Nature Communications, 4 (2013).
- (25) M. Giustina et al., Significant-loophole-free test of Bell’s theorem with entangled photons, Phys. Rev. Lett., 115 (2015), p. 250401.
- (26) A. Gleason, Measures on the closed subspaces of a Hilbert space, J. Math. Mech., 6 (1957), pp. 885–893.
- (27) M. Horodecki, P. Horodecki, and R. Horodecki, Separability of mixed states: necessary and sufficient conditions, Phys. Lett. A, 223 (1996), pp. 1–8.
- (28) M. Howard, J. Wallman, V. Veitch, and J. Emerson, Contextuality supplies the ‘magic’ for quantum computation, Nature, 510 (2014), pp. 351–355.
- (29) C. J. Isham and J. Butterfield, A topos perspective on the Kochen-Specker theorem: I. Quantum states as generalized valuations, Int. J. Theor. Phys., 37 (1998), pp. 2669–2733.
- (30) L. J. Bunce and J. D. Maitland Wright, On Dye’s theorem for Jordan operator algebras, Expo. Math, 11 (1993), pp. 91–95.
- (31) P. Janotta and H. Hinrichsen, Generalized probability theories: what determines the structure of quantum theory?, J. Phys. A, 47 (2014), p. 323001.
- (32) R. V. Kadison, Isometries of operator algebras, Ann. Math., 54 (1951), pp. 325–338.
- (33) M. Kläy, C. Randall, and D. Foulis, Tensor products and probability weights, Int. J. Theor. Phys., 26 (1987), pp. 199–219.
- (34) S. Kochen and E. P. Specker, The problem of hidden variables in quantum mechanics, J. Math. Mech., 17 (1967), pp. 59–87.
- (35) A. J. kowski, Linear transformations which preserve trace and positive semidefiniteness of operators, Rep. Math. Phys., 3 (1972), pp. 275 – 278.
- (36) G. Masse, From the problem of future contingents to Peres-Mermin square experiments: An introductory review to contextuality, 2021.
- (37) M. A. Naimark, On a representation of additive operator set functions., C. R. (Dokl.) Acad. Sci. URSS, n. Ser., 41 (1943), pp. 359–361.
- (38) M. Pawłowski and N. Brunner, Semi-device-independent security of one-way quantum key distribution, Phys. Rev. A, 84 (2011), p. 010302.
- (39) M. Pawłowski et al., Information causality as a physical principle, Nature (London), 461 (2009), pp. 1101–1104.
- (40) A. Peres, Separability criterion for density matrices, Phys. Rev. Lett., 77 (1996), pp. 1413–1415.
- (41) S. Popescu, Nonlocality beyond quantum mechanics, Nature Phys., 10 (2014), pp. 264–270.
- (42) S. Popescu and D. Rohrlich, Quantum nonlocality as an axiom, Found. Phys., 24 (1994), pp. 379–385.
- (43) R. Ramanathan, A. Soeda, P. Kurzyński, and D. Kaszlikowski, Generalized monogamy of contextual inequalities from the no-disturbance principle, Phys. Rev. Lett., 109 (2012), p. 050404.
- (44) R. Raussendorf, Cohomological framework for contextual quantum computations, (2016).
- (45) R. Raussendorf and H. J. Briegel, A one-way quantum computer, Phys. Rev. Lett., 86 (2001), pp. 5188–5191.
- (46) L. K. Shalm et al., Strong loophole-free test of local realism, Phys. Rev. Lett., 115 (2015), p. 250402.
- (47) P. W. Shor, Algorithms for quantum computation: discrete logarithms and factoring, in Proceedings 35th Annual Symposium on Foundations of Computer Science, 1994, pp. 124–134.
- (48) E. Specker, Die Logik nicht gleichzeitig entscheidbarer Aussagen, Dialectica, 14 (1960), pp. 239–246.
- (49) W. F. Stinespring, Positive functions on -algebras, Proc. Am. Math. Soc., 6 (1955), pp. 211–216.
- (50) W. van Dam, Implausible consequences of superstrong nonlocality, Natural Computing, 12 (2013), pp. 9–12.
- (51) N. R. Wallach, An unentangled Gleason’s theorem, vol. 305, Amer. Math. Soc., 2002, p. 291–298.
- (52) E. P. Wigner, Gruppentheorie und ihre Anwendung auf die Quantenmechanik der Atomspektren, Friedrich Vieweg und Sohn, Braunschweig, 1931.
- (53) H. M. Wiseman, S. J. Jones, and A. C. Doherty, Steering, entanglement, nonlocality, and the Einstein-Podolsky-Rosen paradox, Phys. Rev. Lett., 98 (2007), p. 140402.
- (54) C. S. Wu and I. Shaknov, The angular correlation of scattered annihilation radiation, Phys. Rev., 77 (1950), pp. 136–136.