Trading symmetry for Hilbert-space dimension in Bell-inequality violation
Abstract
In quantum information, asymmetry, i.e., the lack of symmetry, is a resource allowing one to accomplish certain tasks that are otherwise impossible. Similarly, in a Bell test using any given Bell inequality, the maximum violation achievable using quantum strategies respecting or disregarding a certain symmetry can be different. In this work, we focus on the symmetry involved in the exchange of parties and explore when we have to trade this symmetry for a lower-dimensional quantum strategy in achieving the maximal violation of given Bell inequalities. For the family of symmetric Collins-Gisin-Linden-Massar-Popescu inequalities, we provide evidence showing that there is no such trade-off. However, for several other Bell inequalities with a small number of dichotomic measurement settings, we show that symmetric quantum strategies in the minimal Hilbert space dimension can only lead to a suboptimal Bell violation. In other words, there exist symmetric Bell inequalities that can only be maximally violated by asymmetric quantum strategies of minimal dimension. In contrast, one can also find examples of asymmetric Bell inequalities that are maximally violated by symmetric correlations. The implications of these findings on the geometry of the set of quantum correlations and the possibility of performing self-testing therefrom are briefly discussed.
I Introduction
Quantum nonlocality, i.e., the quantum violation of a Bell inequality [1], manifests that quantum theory is incompatible with the notion of local causality [2]. In particular, no local-hidden-variable (LHV) theory can reproduce all quantum correlations (measurement statistics) appearing in a Bell experiment. Apart from its foundational significance, the possibility of device-independent [3, 4] (DI) quantum information processing (QIP) also arises as an important byproduct of investigating the general phenomenon of Bell nonlocality [4]. Notable examples of DIQIP protocols include quantum key distribution [5, 6, 7, 8], randomness expansion [9, 10], and various possibilities of black-box certification (see, e.g., [4, 11, 12, 13, 14, 15, 16, 17]).
Although the discussion of Bell nonlocality often centers around the lowest-dimensional qubit systems, the Bell violation of higher-dimensional (HD) quantum states has also been explored both theoretically [18, 19, 20, 21, 22, 23, 24, 25, 26, 27] and experimentally [28, 29, 30, 31]. In fact, it has long been recognized that HD quantum systems can lead to a stronger violation, and hence better resistance to (white) noise [18, 19, 20] and losses [25]. The local Hilbert space dimensions (HSDs) of the shared state, which reflect the complexity of the underlying degrees of freedom, thus serve as a resource for demonstrating Bell nonlocality.
However, for any Bell inequality, there is usually a finite-dimensional quantum strategy (QS)—consisting of the state shared by distant observers and their choice of local measurements—that can attain the maximal violation allowed in quantum theory. For example, it suffices [32] to consider qubits in maximizing the violation of any Bell inequality involving only two binary-outcome local measurements. The general problem of determining the minimal HSD required, nevertheless, is highly nontrivial, see, e.g., [33, 34, 35].
In the case when a Bell inequality comes with a certain symmetry, such as being party-permutation-invariant (PPI) [36, 37, 12, 38, 39], the task of finding its maximal quantum violation can be simplified [40] by considering only quantum correlations, and hence QSs that respect the same symmetry. In particular, when combined with a semidefinite programming (SDP) characterization of the quantum set of correlations [41], either with [42, 43, 44, 45] or without [46, 47, 48, 40] a dimension constraint, the above observation leads to a significant reduction [40, 49, 50, 51] in the number of optimization parameters. In finding an explicit QS that realizes this maximal violation, note, however, that this reduction to symmetry QSs may come at a price of an increase [40] in the required HSD.
Will it be possible to enjoy this symmetry reduction while keeping the required QS at its minimal HSD? Notice that correlations respecting any given symmetry represent a strict subset of all possible correlations. In this context, asymmetry is evidently a resource for Bell-inequality violation. Hence, if the answer to this question is negative for any given Bell inequality, there exists a trade-off between two kinds of resources—HSD and asymmetry—that one may employ to maximize its quantum violation. Here, we systematically explore and answer this question for Bell inequalities defined in several bipartite Bell scenarios, including those with binary outcomes and up to four alternative measurements, as well as those with an arbitrary number of outcomes but only binary measurement choices.
We structure the rest of this paper as follows. In Section˜II, we introduce our notations, recapitulate essential notions of Bell nonlocality [4], and provide a more formal explanation of the notion of a minimal QS. After that, in Section˜III, we introduce the various definitions related to the symmetry of PPI and remind specifically in Section˜III.2 how a QS giving rise to a symmetric correlation can always be converted into a purified, symmetric QS. Examples of Bell inequalities that can be maximally violated using a symmetric QS in the minimal dimension are then provided in Section˜IV. In contrast, examples where a trade-off exists are presented in Section˜V. In Section˜VI, we give a general discussion of asymmetric QSs giving rise to symmetric correlations. Then, we discuss in Section˜VII the implications of some of these examples on the geometry of the set of quantum correlations and self-testing. Finally, we conclude in Section˜VIII. We provide further details about the numerical methods employed in appendix˜A and other miscellaneous results in appendix˜B.
II Preliminaries
II.1 Correlations in Bell scenarios
Consider the bipartite Bell scenario , in which two parties, Alice and Bob, can both perform measurements, each resulting in outcomes. Let . We label Alice’s and Bob’s settings by and their outcomes by , respectively. The statistics of a Bell test yield the joint probability distribution (or correlation) , which manifests how well their measurement outcomes correlate. Throughout, when there is no risk of confusion, the subscripts AB will be omitted for simplicity.
We say that a correlation is Bell-local (hereafter abbreviated as local) if it admits an LHV description [1]:
| (1) |
where is the local-hidden variable (equivalently, shared randomness), is its distribution weight, and with are local response functions. Note that —the set of local correlations—forms a convex polytope, called the Bell polytope, which consists of finitely many extreme points, each corresponding to a local deterministic strategy.
When the outcomes are binary, i.e., , we can also conveniently express a correlation through the expectation values (or correlators) , , and , defined as:
| (2) | ||||
where () is the outcome of Alice (Bob) for her (his) -th (-th) measurement, while and are obtained from via the marginalization over and , respectively. For example, for all .
In contrast with those compatible with an LHV description, a correlation is said to be quantum if it arises from locally measuring a shared quantum state , say, acting on . Explicitly, a quantum correlation associated with a quantum strategy (QS)
| (3) |
of local HSD upper bounded by may be computed from Born’s rule as:
| (4) |
Here, (respectively ) is a positive operator-valued measure (POVM) for Alice’s -th (Bob’s -th) measurement, i.e., and for all where is the identity operator acting on . Henceforth, we denote by the set of quantum correlations (for any given Bell scenario). A celebrated discovery by Bell [1] is that not all can be cast in the form of Eq.˜1, viz. some quantum lies outside of .
Mathematically, a Bell polytope can equivalently be described in terms of a minimal set of halfspaces, called Bell inequalities. In the Bell scenario, a general linear Bell inequality reads as:
| (5) |
where is a vector of real coefficients that determine the corresponding local bound . By definition, a Bell inequality, cf. Eq.˜5, is satisfied by all . Hence, the fact that a , i.e., the correlation is nonlocal, can be witnessed from its violation of a Bell inequality.
II.2 Minimal QS for maximal quantum violation
For any given Bell inequality , it is natural to wonder the extent to which it can be violated quantum-mechanically and identify the QSs that result in this maximal violation. In particular, from a resource-theoretic perspective, it is of interest to determine the minimal, i.e., the smallest local HSD capable of realizing such a quantum violation. For concreteness, we refer to any QS in that realizes the maximal quantum violation of a Bell inequality as a minimal maximizing QS, which we denote by .
For example, for Bell inequalities defined in an Bell scenario (with being any integer larger than or equal to ), it is known [32] that and we may take to be an -qubit pure state along with projection-valued measures (PVMs), i.e., projective measurements. In particular, a well-known two-qubit strategy achieving the maximal violation of the CHSH Bell inequality [52] of Eq.˜7 consists of the following state and observables:
| (8) |
where with are Pauli matrices, while and are Alice’s -th and Bob’s -th observable, which are related to their POVMs by and .
Beyond CHSH, surprisingly little is known about the of various Bell inequalities. Results from [21], [47], and [51] show that for the family of CGLMP inequalities [20] with integer , their minimal dimension . For the specific case of , a dimension bound deduced from a negativity [53] lower bound [40] further shows that this bound is tight, likewise for the cases of and , as we show in Section˜B.1.2.
On the other hand, for the inequality from [54], its could well be infinite, see [34]. Beyond these, the maximal Bell violation of various other Bell inequalities in the bipartite [33, 24, 30, 55] and multipartite [56] Bell scenarios has also been investigated. Again, each explicit QS attaining the quantum maximum provides an upper bound on the corresponding . Some other, more general upper bounds on have also been recently established in [35].
III The symmetry of party-permutation invariance
In Moroder et al. [40], it has been shown that any PPI quantum correlation can always be realized by a QS involving a PPI state and the same set of local measurements performed by each party. Consequently, when there is no restriction in HSD, the maximal quantum violation of a PPI Bell inequality is always attainable [40] using a PPI QS. In this section, we introduce several definitions pertaining to the symmetry of PPI, which is the only symmetry considered in this work.
III.1 Symmetric correlations, Bell inequalities, and strategies
Definition 1 (Symmetric correlation).
A correlation arising from a bipartite scenario is said to be symmetric if it remains invariant under the exchange of parties, , i.e., the simultaneous exchange of the parties’ settings and outcomes :
| (9) |
Definition 2 (Symmetric Bell inequality).
Consider a Bell inequality, cf. Eq.˜5, characterized by the coefficients, . We say that a Bell inequality is symmetric if the coefficients satisfy
| (10) |
Throughout, we use () to denote a symmetric correlation (Bell inequality) whenever we want to emphasize that it satisfies Eq.˜9 [Eq.˜10]. Notably, synchronous correlations [57] arising from the maximally entangled states [58] are symmetric. Synchronous correlations are characterized by the extra constraint: , which means that in the context of a nonlocal game [59], the two players must return the same answer upon receiving identical inputs.
Let us now recall from [40, Proposition 1] a particular kind of QSs that yield a symmetric correlation .
Definition 3 (Symmetric quantum strategy).
In a bipartite Bell scenario, a symmetric QS (abbreviated as SQS), is a QS where both parties perform the same measurements (i.e., ) on a shared PPI state , where
| (11) |
and
| (12) |
is the swap (unitary) operator and is an orthonormal set of basis vectors.
From Eq.˜4 and Definition˜3, one can see that an SQS must give rise to a symmetric quantum correlation :
| (13) | ||||
where we have used the unitarity of and the cyclic property of trace to arrive at the second equality, Eqs.˜11 and 12 to arrive at the third equality, and Definition˜3 to arrive at the fourth equality. Thus, an SQS yields a symmetric correlation, cf. Definition˜1.
III.2 Symmetrization and purification of a QS
While a symmetric quantum correlation need not originate from an SQS, when there is no restriction in the Hilbert space dimension, any asymmetric QS giving a symmetric correlation can always be transformed into an SQS producing the same . We now recall from [40] such a transformation. Given an arbitrary realizing a symmetric correlation, its symmetrized version
| (15) |
can be constructed by introducing ancillary projectors associated with each party (see [40, Eqs. (17,18)]):
| (16) |
where () is a label of Alice’s (Bob’s) ancillary space. By construction, both parties employ the same POVMs given by , whilst the PPI nature of can also be straightforwardly verified. To see that both and give the same symmetric correlation, let us denote by components of the correlation arising from Eq.˜15 and notice from Eqs.˜4 and III.2 that
| (17) | ||||
where the last equality follows from the assumed symmetry of . Thus, this symmetrization embeds the original acting on to an SQS acting on , doubling the original local HSD, while preserving the produced correlation . Importantly, as we see below, the symmetrization of a QS may also be achieved, in some cases, via a local unitary transformation.
On the other hand, it is also well-known that for any given realizing a correlation , one can obtain, through Naimark dilation ([61, Section 9]) and quantum state purification, a purified QS (PQS)
| (18) |
that realizes . In particular, if the state defined in is a (multipartite) PPI density operator, one can follow the method described in [62, Section 4.2] to purify it into a pure state lying in the symmetric subspace. Hence, by concatenating the symmetrization procedure of [40] and the purification procedure, one can always obtain a purified SQS (PSQS) realizing any given . Alternatively, from , we can also obtain a PSQS
| (19) |
via
| (20) | ||||
Hence, we have shown that the following Proposition holds.
Proposition 1.
In a bipartite Bell scenario, a symmetric correlation, cf. Definition˜1, can always be realized using an SQS consisting of a PPI bipartite pure state and with both parties performing the same local PVMs.
Notice that Proposition 1 can be seen as a strengthening of the observation given in [40], in that we may not only take the QS reproducing any given to be symmetric, but also purified. Evidently, Definitions˜1, 2 and 3 can be naturally generalized to an -partite Bell scenario (with ) by demanding PPI for all possible permutations of parties. In this regard, we remark that the above symmetrization procedures, and hence Proposition 1, can be generalized to give the following result.
Proposition 2.
In a multipartite Bell scenario, a PPI correlation can always be realized using a PPI QS consisting of a pure state lying on the symmetric subspace and with all parties performing the same local PVMs.
IV Examples where SQS in the minimal dimension can be maximizing
We now give some examples of symmetric Bell inequalities where we observe no trade-off between symmetry and dimension. In other words, for these inequalities, one can indeed find a that is also an SQS, which we shall denote by
IV.1 The CHSH Bell inequality
Our first example involves the CHSH Bell inequality of Eq.˜6, which is—modulo the freedom in relabeling [54] of settings and outcomes— known [63] to be the only nontrivial facet-defining Bell inequalities in the simplest bipartite Bell scenario. The maximal quantum violation of CHSH can be attained using the asymmetric QS defined in Eq.˜8, giving the symmetric Tsirelson correlation of Eq.˜14 and the so-called Tsirelson bound [60] of . However, as mentioned in [11, Section IV.A] (see also [64]), this quantum value of can also be obtained using an SQS.
To this end, it suffices 111Note that the same transformation is obtained by first rotating by about the -axis, then by about the -axis on the Bloch sphere. to apply on Bob’s qubits a rotation about the -axis of the Bloch sphere, where and . The resulting shared state, which is indeed symmetric, can be written as
| (21a) | ||||
| where and . Moreover, after the same unitary transformation, Alice’s and Bob’s observables become | ||||
| (21b) | ||||
which are clearly symmetric. As we see next in Section˜IV.2, the SQS of Eq.˜21 can be seen as the instance of a more general family of SQSs.
IV.2 The CGLMP Bell inequalities
The CGLMP inequality [20] (see also [65]), denoted by , constitutes a facet-defining inequality of the Bell polytope [66] in the Bell scenario.222For , the outcome-lifted [67] CHSH is another facet-defining Bell inequality in this scenario, but we know from the results of [68] that it cannot exhibit a trade-off. Specifically, for , it may be rewritten as the CHSH Bell inequality of Eq.˜7 after the relabeling . Although in its original form [20] is not PPI, it can be recast into a symmetric form (cf. Definition˜2) by, e.g., relabeling all of Alice’s outcomes from to .
Alternatively, as demonstrated in [69, Appendix B.1.1], by applying the no-signaling conditions [70, 71], normalization constraints, and appropriate relabeling of the outcomes, the CGLMP inequality can be converted to the inequality [54], which is manifestly symmetric:
| (22) | ||||
where and refer, respectively, to the set of such that and . Moreover, for any given , the Bell value of and that of are related [69] by
| (23) |
The best-known Bell violation of the CGLMP inequality (or equivalently ) can be achieved by locally measuring a partially entangled two-qudit state [21] in the judiciously chosen “Fourier-transformed” bases [20] (implementable via a multiport beam splitter in a photonic setup, see, e.g., [18, 19]). Numerical optimizations from [21, 47] confirm this QS to be optimal for dimensions . In particular, for , a sum-of-squares decomposition in the symmetric form provides an analytic bound on the maximal quantum violation [51].
However, the optimal measurements considered in [20] are not PPI, even after we incorporate the outcome-relabeling needed to cast in a PPI form. In the following, we present a family of SQSs that match the best known (and hence ) Bell-inequality violation. For this purpose, we specify the POVMs (with ) as:
| (24a) | |||
| where is the computational basis, and the unitary operator is defined via333To see that the operator define as such is indeed unitary, see Section B.1.1. | |||
| (24b) | |||
Accordingly, the quantum state to be measured is obtained by solving the eigenvector corresponding to the largest eigenvalue of the symmetric Bell operator [72].
As an explicit example, the optimal QS for consists of measuring
| (25) |
where and is a normalization factor. Modulo a local change of basis, is exactly the optimal state for found in [21] (see also [11]). More generally, the optimal quantum state obtained via the Bell operator defined by the PPI measurement strategy of Eq.˜24 is easily seen to be PPI, too. Furthermore, for to , our computation shows that—to six or more significant digits—the SQS formed by these states and measurements indeed gives an (and hence ) value that coincides with the upper bound obtained from the SDP hierarchy of Navascués-Pironio-Acín (NPA). See Section˜B.1.2 for details.
V Examples where SQS in the minimal dimension must be suboptimal
Next, we present some examples of symmetric Bell inequalities where a trade-off between symmetry and dimension has been found.
V.1 Inequalities in the Bell scenario
The complete set of facet-defining Bell inequalities for this Bell scenario was first determined by Froissart [73], then independently rediscovered by Pitowsky and Svozil [74], Sliwa [75], as well as Collins and Gisin [54]. In this case, apart from the (input-lifted [67]) CHSH inequality of Eq.˜6, there is also the so-called inequality, commonly known in an asymmetric form [54, Eq. (19)]. However, via an appropriate relabeling, we can also rewrite this inequality in a PPI form: see [75] and [76] for a symmetric form of this inequality written, respectively, in terms of correlators and probabilities, cf. Eq.˜5. Since for this inequality could well be infinite [34], and one can always follow the procedure outlined in Fig.˜1 to obtain an SQS that is still infinite dimensional, there is little hope to exhibit a trade-off between symmetry and dimension for this Bell inequality.
At first glance, the symmetric -like Bell inequality of [77, Eq. (27)] may seem like an example exhibiting a trade-off, since it was shown therein that its maximal violation can be attained using a family of PPI correlations that arise from asymmetric two-qubit strategies. However, a closer inspection reveals that these strategies—as with Eq.˜8—can be transformed into a symmetric form via a local unitary, see Section˜B.2 for details.
In contrast, consider for the following family of symmetric Bell inequalities in correlator form:
| (26) |
where the local bound is easily verified to be saturated by the symmetric local deterministic strategy
| (27) |
For concreteness, consider now the special case of . Quantum mechanically, it can be verified that an asymmetric two-qubit strategy (see Section˜B.3 for details), which involves a degenerate measurement for , leads to the value , correspondingly. Moreover, this quantum violation arising from the resulting asymmetric correlation agrees with the quantum upper bound on obtained from the NPA hierarchy to a precision better than . Hence, for this inequality. However, if we, instead, only consider SQS facilitated by qubits and PVMs,444Note that it suffices [59, 78] to consider PVMs in maximizing the Bell violation of all two-outcome Bell inequalities. then our computation—using the method described in Section˜A.1.2—shows that such strategies cannot even violate the Bell inequality of Section˜V.1 at all. In other words, there is a trade-off between symmetry and dimension in getting the maximal Bell violation of .
More generally, as we illustrate in Fig.˜2, even though an asymmetric qubit strategy involving a degenerate measurement for can lead to the maximal quantum violation of (up to numerical precision) for all , the maximal quantum violation attainable by a symmetric qubit QS is always suboptimal. In fact, for , these latter strategies can, at best, give the local bound of .
V.2 Inequalities in the Bell scenario
| Quantum bound | SQS bound | Quantum bound | SQS bound | ||||||
|---|---|---|---|---|---|---|---|---|---|
| 0.4142 | 0.2500 | 0.3846 | LB | 0.4878 | 0.4843 | 0.4843 | 0.4843 | ||
| 0.4349 | 0.2500 | 0.3466 | UB | 0.4878 | 0.4878 | ||||
| 0.4349 | 0.3913 | 0.4067 | LB | 0.4972 | 0.4514 | 0.4832 | 0.4836 | ||
| 0.6722 | 0.5682 | 0.5682 | UB | 0.4972 | 0.4972 | ||||
| 0.7559 | 0.7500 | 0.7500 | LB | 0.6927 | 0.6671 | 0.6671 | 0.6722 | ||
| 0.8484 | 0.8195 | 0.8195 | UB | 0.6927 | 0.6927 | ||||
As we increase the number of settings to four, it is known [79, 55, 80] that the Bell polytope is completely characterized by 175 facet-defining classes of Bell inequalities. Among them, 55 are known [36] to admit a symmetric representation. Several of these, namely, the trivial positivity facet, the CHSH inequality of Eq.˜6, and the inequality, are liftings [67] of facet-defining Bell inequalities from simpler Bell scenarios.
For the remaining 52 symmetric Bell inequaliites with four settings, except for (whose remains unknown [34]) and 8 others (see Table˜1 as well as Tables˜5 and 6 of Section˜B.4), all the rest can be maximally violated using qubit strategies, i.e., having . Moreover, our computation results show that for 42 out of these 52 inequalities, there is no trade-off, i.e., there exists a . For ease of reference, we provide this list of 42 inequalities and their maximal quantum violation in Table˜5 of Section˜B.4.
After discounting , we are left with 9 symmetric inequalities, where we observe a trade-off between dimension and symmetry; see Table˜1 for details. It is worth noting that for all these 9 inequalities, as with , the we have found always involves projectors of unequal ranks between Alice and Bob, i.e., for some , thereby making these QS evidently asymmetric. In fact, a closer inspection reveals that the resulting correlation is also asymmetric.
Also worth noting is that for the inequality, we further observe a gap between the maximal quantum value attainable using qubit SQS and symmetric correlations that may also arise from asymmetric strategies. Indeed, for these two classes of strategies, the maximal quantum values attainable are and , respectively. More explicitly, the latter value can be achieved by adopting the following asymmetric QS consisting of the shared state
| (28a) | ||||||
| with and the mirror-symmetric observables | ||||||
| (28b) | ||||||
| where ∗ denotes complex conjugation and the measurement (Bloch) vectors , | ||||||
| (28c) | ||||||
Equivalently, may be obtained from by the substitution . For graphical representation of and , see Fig.˜3.
In the next section, we provide a proof showing that QSs of the kind given in Eq.˜28 not only produce a symmetric correlation, but also cannot be symmetrized via a local unitary transformation.
VI Asymmetric strategies giving symmetric correlations
VI.1 Symmetric correlations from mirror-symmetric strategies
To see that the correlation resulting from Eq.˜28 is indeed PPI, we remark that a more general class of mirror-symmetric strategies must also produce symmetric correlations. Indeed, Eq.˜28a is easily seen to be a special case of the following family of two-qubit pure states:
| (29) |
where is the singlet state and is some real symmetric state (i.e., a real, linear combination of the remaining three Bell states and ).
Moreover, if we locally measure Eq.˜29 for a pair of measurement directions mirror-symmetric with respect to the plane of the Bloch sphere, i.e.,
| (30) |
it follows that the resulting correlation must be symmetric.
Theorem 3.
To see that this is the case, we start by establishing the following Lemma.
Lemma 4.
If a pair of measurement bases are mirror-symmetric with respect to the plane, i.e., Eq.˜30 holds, then their POVM elements satisfy
| (31) |
Proof.
For dichotomic measurements, the POVM elements can be expressed as
| (32) |
where is the vector of Pauli matrices, and . Since , we have
| (33) |
Therefore, . ∎
The proof of Theorem˜3 may now be completed as follows.
Proof.
Using Lemma 4, the joint probability distribution arising from measuring the state with the mirror-symmetric measurements reads as:
| (34) | ||||
Since is real-valued, we have
| (35) | ||||
where , i.e., the complex conjugation of . With the observation that , where is the swap operator, we can further reduce to
| (36) | ||||
where Lemma 4 is again utilized to arrive at the third equality. Hence, arising from this strategy satisfies the symmetry condition defined in Definition˜1. ∎
We now show that no local operations can symmetrize the measurement bases of the two players provided that their measurement directions are each not coplanar on the Bloch sphere.
Proposition 5.
Let with be non-coplanar, i.e, , and let satisfy Eq.˜30, where is a reflection. Then there do not exist local unitaries such that the resulting measurement directions satisfy for all .
Proof.
Suppose the contrary that there exists such that
| (37) | ||||||
with for all .
Let . Since for all , we have . So we can focus on proving the existence of this single unitary . It is well-known that for each there exists a rotation such that . Hence, there exists some such that . By assumption, we also have . Altogether, we have , using the fact that for any orthogonal matrix . Define . Since the set are not coplanar, they span . And since for all , the multiplicity of the eigenvalue of must be , whenever the cardinality of the set satisfies . In other words, (in some basis). However, this is impossible since , where the last equality follows from the fact that is a reflection. ∎
In addition, as a special case of the following Proposition, we see that the only pure states that can generate symmetric correlations with mirror-symmetric qubit measurements are exactly those of the form defined in Eq.˜29.
Proposition 6.
Let be a QS satisfying and that produces a symmetric correlation. Suppose further that the collection of over all possible spans the entire space that they act on. Then, apart from a global phase factor, the state must be of the form
| (38) |
Here, is a real-valued symmetric (Sym) state and is a purely imaginary antisymmetric (ASym) state, i.e.,
| (39) |
where is the swap operator defined in Eq.˜12.
Proof.
By assumption, produces a symmetric correlation when measured with POVMs satisfying Eq.˜31. For convenience, we write , , and . To prove the Lemma, we employ the three properties:
-
(i)
the correlation is symmetric, i.e., ,
-
(ii)
the correlation is real-valued, i.e., , and
-
(iii)
the POVMs satisfy Eq.˜31.
Using (ii) and (iii), we arrive at the following:
| (40) |
while with (i), (iii), and the defining property of the swap operator, we arrive at
| (41) |
Subtracting the terms at the end of the last two lines yields
| (42) |
where . Now, since span the entire space that they act on, also span the entire composite Hilbert space. In particular, we can consider an appropriate linear combination of these operators, and hence of Eq.˜42, to arrive at . However, the left-hand side of the last equation is simply the Hilbert-Schmidt norm of the operator . This means that must be identically zero, and hence
| (43) |
for some .
Without loss of generality, we can express as a linear combination of Sym and ASym components, cf. Eqs.˜38 and 39 where while both and may be complex. Substituting Eq.˜38 into Eq.˜43 gives
| (44) |
where we have used the complex conjugation of Eq.˜39. Let us define and . Then, it follows from Eq.˜44 that
| (45) |
indicating that is a real-valued symmetric vector, while is a purely imaginary anti-symmetric vector. Hence, we can write , which completes the proof. This is equivalent to the form defined in Eq.˜29 up to some global phase. ∎
VI.2 More general construction of QS giving symmetric correlations
The discussion in Section˜VI.1 makes evident that symmetric correlations may arise beyond SQSs. In fact, we may consider both SQSs and mirror-symmetric strategies discussed in Section˜VI.1 as special cases of a more general construction. To this end, let us note from Eq.˜4 and Definition˜1 that for a symmetric correlation , we have
| (46) |
By definition, we may also write the last term of Eq.˜46 as
| (47) |
Consider now a linear positive map that leaves the trace in Eq.˜46 invariant, i.e.,
| (48) |
for some . After substituting Eqs.˜47 and 48 into Eq.˜46, we see that for the latter to hold, it suffices that the following conditions hold
| (49a) | |||
| (49b) | |||
For example, if we take to be the identity map, then Eq.˜49 implies that:
| (50a) | |||
| (50b) | |||
By summing both sides of Eq.˜50b, e.g., over and using the normalization condition of a POVM, we see that Eq.˜50 is exactly the condition of an SQS, cf. Definition˜3
On the other hand, if we take to the complex conjugation (or equivalently, transposition) operation, then Eq.˜49 becomes
| (51a) | |||
| (51b) | |||
which is easily verified to hold for mirror-symmetric strategies discussed in Section˜VI.1, see Eqs.˜31 and 43.
For an explicit example, if we take mirror-symmetric qubit measurements, then for any two-qubit state, we have
| (52) |
where , and is the correlation matrix. To satisfy Eq.˜51, we require and
| (53) |
This gives the most general form of a two-qubit state that produces a symmetric correlation with mirror-symmetric measurements.
VII Asymmetry and the geometry of the quantum set of correlations
VII.1 Maximal quantum violation of a symmetric Bell inequality by an asymmetric correlation
As illustrated above, there exist symmetric Bell inequalities that can be maximally violated by an asymmetric correlation. It turns out that this has a nontrivial implication on the geometry of the quantum set of correlations, as we summarize in the following proposition.
Proposition 7.
If an asymmetric quantum correlation maximally violates a symmetric Bell inequality, then the set of quantum maximizers of this Bell inequality forms at least a one-dimensional flat region (boundary) in the space of correlations.
Proof.
Let be a symmetric Bell inequality defined (for an -partite Bell scenario) via , cf. Eq.˜10, and be a quantum maximizer of that is asymmetric. The asymmetric nature of implies that there must exist at least one element from the symmetric group such that upon its action on , we obtain where is the permutation operator effecting the permutation . Since is a maximizer of and is symmetric, we see that
| (54) | ||||
where the third equality follows from the fact that a simultaneous permutation on the Bell coefficients and the correlation leaves their inner product unchanged. Hence, must also be a maximizer of . Combining this observation with the linearity of in , we see that any convex combination of and , i.e.,
| (55) |
is also a maximizer of . Since is convex and is linear in , only those correlations lying on the boundary of can be a maximizer of in . In other words, the set of correlations defined by Eq.˜55 forms a one-dimensional flat region of the quantum boundary, thus concluding the proof. ∎
An important consequence of Proposition 7 is the following Corollary.
Corollary 8.
A symmetric Bell inequality admitting an asymmetric maximizer cannot be used to self-test any reference quantum strategy based on the observed maximal violation of alone.
Proof.
Self-testing [81] of a reference strategy based on an observed Bell value refers to the possibility of deducing from this observation that the underlying strategy, modulo irrelevant local degrees of freedom, must be . A necessary condition for this possibility is that there is only a unique quantum correlation leading to this Bell value. By assumption, the symmetric inequality admits an asymmetric maximizer, thus it follows from Proposition 7 that its maximizer is not unique, thus rendering it impossible to perform self-testing from the maximal violation of alone. ∎
Notice, however, that the above Corollary does not preclude the possibility of self-testing the underlying state (but not the measurements) from the observed Bell value, like the examples discussed in [68, 82, 83]. Also noteworthy is that flat regions of the quantum boundary have been extensively discussed in [77]. However, most of the examples presented therein concern flat regions that are described by a trivial Bell inequality, i.e., one that is not violated by quantum theory. In contrast, the flat region corresponding to of Section˜V.1, as with that for the inequalities listed in Table˜1, does not intersect with the Bell polytope.
VII.2 Maximal quantum violation of an asymmetric Bell inequality by a symmetric correlation
While a symmetric Bell inequality can be maximally violated by a symmetric correlation, we generally cannot hope that the same conclusion holds for an asymmetric Bell inequality. To see this, consider, e.g., the CHSH inequality obtained from Eq.˜7 by the relabeling :
| (56) |
If a correlation violates this inequality, one would have
| (57) |
If the is also PPI, then the correlation remains invariant under the transformation for all , which means that it must also satisfy, from Eq.˜57, after a simultaneous relabeling for and :
| (58) |
However, the conjunction of Eqs.˜57 and 58 contradicts the fact [84] that any correlation can violate at most one version of the CHSH Bell inequality. In other words, no symmetric correlation can violate the asymmetric CHSH Bell inequality of Eq.˜57, let alone maximally.
In contrast, consider the following two-parameter family of Bell inequalities:
| (59a) | ||||
| for | ||||
| (59b) | ||||
and the pair could otherwise take any value within the interior of the octagon spanned by the 8 vertices:
| (60) |
with . For completeness, we provide the local bound of this two-parameter family of inequalities in Eq.˜84 of Section˜B.6. It is easy to verify that the Bell inequalities of Eq.˜59 cannot be written in a PPI form, even after any relabeling of the settings and outcomes. At the same time, it has been shown in [85] that they are all maximally violated by the Tsirelson correlation of Eq.˜14, which is PPI. In other words, the family of asymmetric Bell inequalities given by Eq.˜59 are all maximally violated by the same symmetric correlation .
VIII Conclusion
Due to the arbitrariness in the classical labeling (of settings, outcomes, and parties) as well as the degeneracy arising from the normalization and no-signaling constraints, a Bell inequality may be rewritten in various forms. However, even after incorporating these degrees of freedom, not all Bell inequalities can be cast in a symmetric (party-permutation-invariant, PPI) form. In this work, we have focused on (facet-defining) Bell inequalities that indeed admit a symmetric representation.
In particular, we have explored the possibility of finding, for symmetric Bell inequalities, a symmetric quantum strategy (SQS) of minimal dimension that maximizes their Bell violation. Our results show that, for integer , there exists an SQS in dimension maximizing the Bell violation of the family of symmetric inequalities (known [69] to be equivalent to the CGLMP Bell inequalities). Moreover, for , no quantum strategy of a smaller Hilbert dimension can attain the same maximal value.
However, for a family of symmetric Bell inequalities in the Bell scenario and 9 of the nontrivial, symmetric, genuine four-setting facet-defining Bell inequalities applicable to the Bell scenario, we observe a gap between the maximal quantum violation achievable and that arising from an SQS of minimal dimension. In other words, there exists a trade-off between symmetry and dimension for achieving the maximal quantum violation of these inequalities—either we opt for an SQS that is higher-dimensional, or we go for an asymmetric quantum strategy (QS) that is of minimal dimension, but not both. Note, however, that we do not know if one can find an example of such a trade-off in a simpler Bell scenario, a problem that may deserve further investigation.
Interestingly, since the optimal, minimal-dimension quantum strategy for all these symmetric inequalities gives rise to an asymmetric correlation, it follows from Proposition 7 that the quantum maximizers of each of these Bell inequalities must define a flat region of the quantum boundary. As explained in Section˜VII, this feature has negative implications on the possibility of performing self-testing using the observed quantum violation of these inequalities.
Also worth noting is that among the inequalities showing a trade-off, the maximal (qubit) violations of under different constraints satisfy:
| (61) |
Hence, with the assumption that the employed QS is a qubit strategy, one may certify that the underlying strategy is asymmetrical, even if the observed correlation is symmetric, i.e.,
| (62) |
serves as a semi-device-independent [86] witness for an asymmetric qubit QS. In fact, the trade-off is so strong that even if we allow arbitrary symmetric correlations attainable with two-qubit QSs, there remains a gap with the maximal quantum violation of . In contrast, we present in Section˜B.5 an example of a -setting Bell inequality that again shows a gap between the maximal two-qubit SQS violation and that achievable with two-qubit symmetric correlations, but where the latter already achieves the quantum maximum.
| Bell Inequality | Facet | (SQS, , ) | |||
| CHSH [52]: Eq.˜7 | ✓ | ✓ | 2 | ✓ | Eq.˜8:(✗, ✓, ✓) |
| CHSH [52]: Eq.˜7 | ✓ | ✓ | 2 | ✓ | Eq.˜21:(✓,✓, ✓) |
| [54]: Section˜IV.2 | ✓ | ✓ | 3 | ✓ | Eq.˜24:✓, ✓, ✓) |
| [77]: Section˜V.1 | ✗ | ✓ | 2 | ✓ | Eq.˜73:(✓, ✓, ✓) |
| : Section˜V.1 | ✗ | ✓ | 2 | ✗ | Eq.˜74:(✗, ✗, ✓) |
| [33]: Section˜B.4 | ✓ | ✓ | 2 | ✗ | Eq.˜28:(✗, ✓, ✗) |
| [85]: Eq.˜59 | ✗ | ✗ | 2 | ✓ | Eq.˜21:(✓,✓, ✓) |
Of course, by considering a nontrivial, asymmetric Bell inequality, one may also hope to drop the dimension assumption and go for a fully device-independent witness for asymmetry. Indeed, for all facet-defining Bell inequalities in the and Bell scenarios with and , we observe that there always exists an asymmetric representation that cannot be violated by symmetric quantum correlations. It will certainly be interesting to see if this feature holds true in general, specifically, for all Bell scenarios. For a summary of some other results we have obtained, see Table˜2.
Note also that the realizability of symmetric distributions by a symmetric model has also been explored in the context of network nonlocality. In fact, it has recently been shown [87] that there exist symmetric distributions in the triangle network (under exchange of parties) that cannot be created via a symmetric model assuming only no-signaling, regardless of the dimension or nature of the system. Moreover, such intriguing distributions can be created via a local asymmetric model.
Finally, note also that while we have focused on the symmetry of (full) permutation invariance, there are clearly interesting alternatives to consider. For instance, one may consider Bell inequalities that exhibit translationally (also called cyclic-permutation) invariance [56, 88, 89]. Could one find a similar kind of trade-off between dimension and symmetry? We leave the exploration of this and other interesting questions for future work.
Acknowledgements.
We thank István Márton, Károly F. Pál, Gilles Pütz, William Slofstra, and Elie Wolfe for helpful discussions. HYH thanks Bo-An Tsai for sharing his code, which led to some of the results presented here. KSC thanks the hospitality of the Institut Néel, where part of this work was completed. This work was supported by the National Science and Technology Council, Taiwan (Grants No. 109-2112-M-006-010-MY3, 112-2628-M-006-007-MY4, 113-2917-I-006-023, 113-2918-I-006-001), the Foxconn Research Institute, Taipei, Taiwan, in part by the Higher Education Sprout Project, Ministry of Education, to the Headquarters of University Advancement at the National Cheng Kung University (NCKU), and in part by the Perimeter Institute for Theoretical Physics. Research at Perimeter Institute is supported by the Government of Canada through the Department of Innovation, Science, and Economic Development, and by the Province of Ontario through the Ministry of Colleges and Universities. T. V. acknowledges the support of the European Union (CHIST-ERA MoDIC), the National Research, Development and Innovation Office NKFIH (Grants No. 2023-1.2.1-ERA_NET-2023-00009 and No. K145927) and the ‘Frontline’ Research Excellence Program of the NKFIH (No. KKP133827).Appendix A Bounding quantum values of Bell inequality allowed by symmetric correlations and symmetric strategies
Let be the set of operators acting on . Given a bipartite Bell inequality specified by the coefficients , our goal is to obtain its maximal value over all possible symmetric quantum strategies (SQSs) of local Hilbert space dimension upper bounded by , i.e.,
| such that | ||||
| (63) | ||||
where defined in Eq.˜12 is the swap operator.
Since an SQS necessarily gives a symmetric correlation, cf. Eq.˜13, this problem may be relaxed, e.g., by optimizing over symmetric correlations attainable by QSs of the same HSD, i.e.,
| such that | ||||
| (64) | ||||
In other words, the optimum value of appendix˜A is always larger than or equal to that of appendix˜A.
A.1 Upper bounding the quantum violation of a Bell inequality
A.1.1 Without dimensional constraints
When there is no dimension constraint, i.e., if , there is no distinction between the Bell value achievable by SQSs and symmetric correlations, see Proposition 2. Then, an upper bound on the maximal quantum violation by SQS can be obtained by solving any SDP hierarchy that outer approximates the quantum set , such as that due to Navascués-Pironio-Acín (NPA) [46, 47] and Moroder et al. [40], but now with the additional constraints that all the moments are invariant under the action of the symmetry group . For example, in the case of a bipartite Bell scenario, cf. Definition˜1, we require that all moments remain invariant under the exchange , which means that all moments of the kind can be taken to be the same as . To this end, one may first take the moment matrix at any given level of these hierarchies and perform the twirling,
| (65) |
where is the unitary operator that effects the appropriate row-permutation of corresponding to the action of . Clearly, consists of fewer independent moments compared with . A symmetric upper bound is then obtained by maximizing the Bell value subject to instead of . For further information, see the detailed discussion given in [49, 51].
A.1.2 With dimensional constraints
When a local dimension bound is present, a powerful analog of the NPA hierarchy was proposed in [44, 45]. In particular, the software package QDimSum discussed in [50] has been developed to implement precisely this and other related finite-dimensional optimizations. To this end, notice that whenever a symmetric Bell inequality is provided to QDimSum as an objective function, Eq.˜65 is implemented with respect to the symmetric group that leaves the Bell inequality invariant. Hence, whenever we use QDimSum to maximize the value of a PPI Bell inequality, we naturally obtain an upper bound on appendix˜A, i.e., the maximal value allowed by symmetric correlations arising from dimension-bounded (not necessarily symmetric) QSs.
To obtain a (tighter) upper bound applicable to SQS, i.e., appendix˜A, we modify the sampling procedure described in [44, 45] for generating random moment and localizing matrices. Specifically, instead of sampling normalized random pure states from , we randomly sample symmetric (antisymmetric) pure states from by considering random, but normalized linear combination of vectors from the symmetric (antisymmetric) subspace of . Moreover, instead of sampling random POVMs for Alice and Bob independently, we randomly sample Alice’s POVM and make Bob’s POVMs identical to those of Alice.
A.2 Lower bounding the quantum violation of a Bell inequality
A.2.1 Over symmetric quantum strategies
To obtain a lower bound for the maximization problem of appendix˜A, we may:
-
1.
Without loss of generality, take to be a pure state and choose to be an arbitrary, normalized linear combination of basis states in the (anti)symmetric subspace.
-
2.
Set each to be a projector
(66) such that the sum of their rank for all , where may be taken as an orthonormal set of column vectors forming a unitary matrix .
-
3.
Optimize over the the parameters defining and the unitary operators .
Notice that for step (1), a general symmetric and antisymmetric state can be expressed, respectively, as a linear combination of symmetric and antisymmetric basis states. Moreover, for step (2), the QLib package [90] provides a convenient parametrization of any element of SU() via real parameters. All these parameters can then be optimized over, e.g., using the function in MATLAB to obtain a local maximum of appendix˜A. Since these are high-dimensional optimizations over real parameters, with being the number of measurement settings, it is usually necessary to perform multiple optimizations to get a good lower bound.
A.2.2 Over symmetric quantum correlations
For the maximization problem of appendix˜A, which provides an upper bound to that of appendix˜A, we may adapt the (see-saw) algorithm described in [78, 33, 24] to iteratively optimize over
-
1.
Alice’s POVM for given state and Bob’s POVM ;
-
2.
Bob’s POVM for given state and Alice’s POVM ;
-
3.
for given Alice’s POVM and Bob’s POVM .
Note that each of these optimization problems is an SDP and remains so even if we include the symmetry requirement of Eq.˜9, cf. the first constraint of appendix˜A. Moreover, if the optimized strategy turns out to be symmetric, the optimum value would also be a legitimate lower bound of appendix˜A. Also, instead of Step 2, one may manually set Bob’s POVM to be the same as that of Alice before running step (3) above [while imposing Eq.˜9] to ensure that the resulting strategy is an SQS. However, with this modification, the iterative algorithm is no longer guaranteed to converge.
Appendix B Miscellaneous Details
B.1 CGLMP Inequalities and their quantum violation
B.1.1 Unitarity of the operators specified in Eq.˜24b
Here, we provide further details to illustrate the unitarity of the operators given in Eq.˜24b, and hence . In the case of , since it is a diagonal matrix having only eigenvalues and , it is unitary.
To see that is unitary, it suffices to show that its rows are orthonormal, which from Eq.˜24b, amounts to demanding
| (67) |
We now give a proof that Eq.˜67 holds.
Proof.
We start by noting the trigonometric identity
| (68) |
which holds for any real and any integer . Eq.˜68 can be easily shown from the more familiar identities:
| (69) |
by setting , , and summing both sides of the first identity over .
Next, let , and substitute these into Eq.˜68. Then, the LHS of Eq.˜68 becomes . The value of this expression remains unchanged if we replace the summation index by , thus giving, modulo a factor of , exactly the LHS of Eq.˜67. Incorporating this factor, we may then rewrite the LHS of Eq.˜67, via Eq.˜68, as
| (70) |
Consider now the case where , then the denominator of Eq.˜70 never vanishes for . However, since is a half-integer multiple of [likewise for ], their cotangent vanishes, and hence the numerator also vanishes, i.e., Eq.˜70, and thus the LHS of Eq.˜67 vanishes whenever . To evaluate Eq.˜70 for the case of , we apply the l’Hôpital rule to obtain
| (71) |
since is a half-integer multiple of . In other words, the LHS of Eq.˜67 indeed becomes unity if . Hence, Eq.˜67 holds as claimed. ∎
B.1.2 Optimal quantum violation and the minimal dimension
In Table˜3, we summarize our results concerning the maximal quantum violation of the CGLMP inequality [20] for to , obtained via the SQS described in Eq.˜24. Notice that these results generalize those presented in [21, 47] for to .
| Quantum value | Difference | Quantum value | Difference | ||
|---|---|---|---|---|---|
| 2 | 2.82842718 | 5.9e-8 | 11 | 3.15549968 | 1.9e-7 |
| 3 | 2.91485425 | 4.1e-8 | 12 | 3.16979224 | 2.6e-7 |
| 4 | 2.97269840 | 1.5e-7 | 13 | 3.18274300 | 2.7e-7 |
| 5 | 3.01571048 | 7.7e-9 | 14 | 3.19456537 | 3e-8 |
| 6 | 3.04970041 | 1.8e-10 | 15 | 3.20542659 | 1.3e-7 |
| 7 | 3.07764831 | 2.4e-9 | 16 | 3.21546005 | 1.34e-6 |
| 8 | 3.10128058 | 1.1e-10 | 17 | 3.22477378 | 2.5e-7 |
| 9 | 3.12168442 | 1e-8 | 18 | 3.23345644 | 4e-8 |
| 10 | 3.13958741 | 1.6e-7 | 19 | 3.24158164 | 9e-8 |
Next, we show in Table˜4 our results for bounding the minimal dimension required to get the maximal violation (or the maximal-violating correlation) of the CGLMP inequality for to .
| Neg () | Neg() | Neg() | Level | Size | ||
| 2 | 0.5 | 0.5000 | 0.5000 | 2 | (25; 31) | 2 |
| 3 | 0.9836 | 0.9835 | 0.9836 | 2 | (169l; 1,003) | 3 |
| 4 | 1.4561 | 1.4559 | 1.4561 | 2 | (625; 14,797) | 4 |
| 5 | 1.9203 | 1.8905 | 1.8909 | 2 | (1,681; 116,702) | 5 |
| 6 | 2.3778 | 1.7578 | 1.7603 | 1+ | (2,025; 118,155) | 5 |
| 7 | 2.8298 | 1.5896 | 1.5919 | 1+ | (2,809; 97,423) | 5 |
| Inequality | Quantum Bound | Inequality | Quantum Bound | Inequality | Quantum Bound | |||
|---|---|---|---|---|---|---|---|---|
| 2 | 0.4353 | 4 | 0.4677 | 2 | 0.8175 | |||
| 4 | 0.3004 | 4 | 0.7262 | 2 | 0.9763 | |||
| 3 | 0.2879 | 2 | 0.6380 | 2 | 0.8398 | |||
| 2 | 0.6056 | 2 | 1.0130 | 2 | 1.2993 | |||
| 2 | 0.5413 | 2 | 0.6742 | 2 | 1.0648 | |||
| 2 | 0.8785 | 2 | 0.8156 | 2 | 1.1584 | |||
| 2 | 0.4617 | 2 | 0.6402 | 2 | 0.6651 | |||
| 2 | 0.6139 | 2 | 0.8972 | 2 | 1.0742 | |||
| 2 | 0.6383 | 2 | 0.7500 | 2 | 0.9677 | |||
| 2 | 0.6188 | 2 | 1.0246 | 2 | 1.7261 | |||
| 2 | 0.4794 | 2 | 0.5901 | 2 | 0.9457 | |||
| 2 | 0.4142 | 2 | 0.7596 | 2 | 1.0000 | |||
| 2 | 0.6714 | 2 | 0.8814 | 2 | 1.0135 | |||
| 3 | 0.6430 | 2 | 0.5923 | 2 | 0.8704 |
B.2 SQS maximally violating an -like inequality
Consider the Bell inequality from [77, Eq. (27)]
| (72) | ||||
which can be seen, after relabeling, as keeping only the correlation part of the inequality. The optimal quantum strategy presented in [77] is not PPI. However, we can turn it into an SQS by first applying , followed by on Bob’s Hilbert space. After this rotation, the strategy becomes
| (73) | ||||
which is easily verified to give a quantum value of for all . Notice that, by construction, despite its dependence on , the two-qubit state of Eq.˜73 is always maximally entangled.
B.3 Optimal quantum strategy for
An asymmetric two-qubit quantum strategy that gives the maximal violation of the family of Bell inequalities of Section˜V.1 consists of Alice and Bob sharing the two-qubit state
| (74a) | |||
| and measuring the dichotomic observables: | |||
| (74b) | |||
| For , the actual values of these parameters are | |||
| (74c) | |||
| thereby giving , while those for are | |||
| (74d) | |||
B.4 Symmetric bipartite facet-defining Bell inequalities with four binary-outcome measurements
Among the complete list of 175 facet-defining Bell inequalities of the Bell scenario , 55 of them can be cast in a symmetric form. Apart from the positivity facet, the CHSH Bell inequality of Eq.˜6, the (symmetric) Bell inequality [75, 76], the 9 inequalities listed in Table˜1, and the inequality, the remaining 42 symmetric four-setting facet-defining Bell inequalities, listed in Table˜5, show no trade-off between symmetry and dimension.
For completeness, we provide here the explicit form of the Bell inequality:
| (75) | ||||
where we use to refer to the six remaining terms of and s that have to be included to ensure that the Bell inequality is PPI.
For the various SQS (and symmetric-correlation) bounds of those inequalities from Table˜5 with , see Table˜6.
| Inequality | Quantum Bound | |||
|---|---|---|---|---|
| 4 | 0.2990 | 0.2990 | 0.3004 | |
| 3 | 0.2500 | 0.2879 | 0.2879 | |
| 3 | 0.4676 | 0.6430 | 0.6430 | |
| 4 | 0.3056 | 0.3662 (0.4362) | 0.4677 | |
| 4 | 0.6719 | 0.6830 | 0.7262 |
B.5 Quantum violation of
In this Appendix, we shall present a symmetric Bell inequality defined for the Bell scenario such that there is again a trade-off between symmetry and dimension when one maximizes its Bell violation. In particular, even though the maximal quantum violation of can be achieved using a symmetric correlation arising from a symmetric two-qubit state and a mirror-symmetric measurement strategy, the maximal quantum value attainable using a two-qubit SQS is always suboptimal.
Let us first recall the self-testing results shown in [91]. There, the authors considered the following Bell inequality (introduced in [92]):
| (76) | ||||
where, cf. Eq.˜7,
| (77) |
The maximal quantum violation of is easily seen to be , which one can attain by locally measuring the Bell state with the following dichotomic observables:
| (78) |
In fact, it was shown in [91] that the maximal quantum value of self-tests the two-qubit maximally entangled state and, up to complex conjugation, the Pauli observables for and . Now consider the following Bell inequality in the scenario of :
| (79) |
where is obtained from by swapping the roles of Alice and Bob, i.e., . The inequality is symmetric according to Definition˜2.
Proposition 9.
The maximal quantum value of cannot be achieved by any symmetric qubit quantum strategy.
Proof.
The maximal quantum value of , namely , can be attained by locally measuring with the observables , specified in Eq.˜78 and their complex conjugation, i.e., for to 9 and for to 3. In particular, any realization achieving the maximal value of also achieves the maximal value of . According to the self-testing result of [91], the underlying state is therefore locally isometric to a two-qubit maximally entangled state, . In particular, the state in any qubit strategy reaching the maximal value of can be written as for some . The identity allows us to consider local unitaries operating on only one side without loss of generality.
Suppose, for contradiction, that there exists some symmetric qubit strategy achieving the maximal value of , i.e., , for . To reach , the three diagonal correlators must all be . Then for each , we have
| (80) |
where the last inequality follows from the Cauchy-Schwarz inequality. Since the inequality is saturated, we must have , where denotes transposition. Applying and on both sides, we infer that , which further implies that
| (81) |
where we have used the fact a qubit observable [and hence ] squares to the identity operator. Equivalently, Eq.˜81 means that there exists some rotation satisfying
| (82) |
where and . Heuristically, it asserts the existence of some rotation that can simultaneously map all s to each of their own reflection .
However, at the maximal value, the observables correspond to three mutually orthogonal Bloch vectors [91] , which cannot be coplanar. Hence, as explained in Proposition 5, there cannot be a rotation that satisfies Eq.˜82, thus leading to a contradiction. Therefore, no symmetric qubit strategies could reach the maximal quantum value of . ∎
What then is the largest value of achievable with a qubit SQS? To this end, the best violation that we have found using a qubit SQS is , where the SQS consists of measuring the Bell state with the observables
| (83) |
and for to . Note that all these observables are real, and thus give rise to measurement (Bloch) vectors lying on the plane.
B.6 Local bound for the Bell inequalities of Eq.˜59
The local upper bound for the Bell expression defined in Eq.˜59 is, for any given pair :
| (84) |
where the expression in each term allows all combinations of signs. The actual bound depends on the octagon slice [spanned by and the vertices of Eq.˜60] to which the point belongs.
For instance, consider the case where belongs to the octagon slice spanned by , , and , with . Since , evaluating the maximum from Eq.˜84 gives
| (85) |
Similarly, if belongs to the octagon slice spanned by , , and , then , thus evaluating the maximum from Eq.˜84 gives
| (86) |
The local upper bound of for belonging to the other six octagon slices can be easily deduced accordingly.
References
- Bell [1964] J. S. Bell, Physics 1, 195 (1964).
- Bell [2004] J. S. Bell, Speakable and Unspeakable in Quantum Mechanics: Collected Papers on Quantum Philosophy, 2nd ed. (Cambridge University Press, 2004).
- Scarani [2012] V. Scarani, The device-independent outlook on quantum physics, Acta Phys. Slovaca 62, 347 (2012).
- Brunner et al. [2014] N. Brunner, D. Cavalcanti, S. Pironio, V. Scarani, and S. Wehner, Bell nonlocality, Rev. Mod. Phys. 86, 419 (2014).
- Ekert [1991] A. K. Ekert, Quantum cryptography based on Bell’s theorem, Phys. Rev. Lett. 67, 661 (1991).
- Mayers and Yao [2004] D. Mayers and A. Yao, Self Testing Quantum Apparatus, Quantum Info. Comput. 4, 273 (2004).
- Barrett et al. [2005a] J. Barrett, L. Hardy, and A. Kent, No Signaling and Quantum Key Distribution, Phys. Rev. Lett. 95, 010503 (2005a).
- Vazirani and Vidick [2014] U. Vazirani and T. Vidick, Fully Device-Independent Quantum Key Distribution, Phys. Rev. Lett. 113, 140501 (2014).
- Pironio et al. [2010] S. Pironio, A. Acín, S. Massar, A. B. d. l. Giroday, D. N. Matsukevich, P. Maunz, S. Olmschenk, D. Hayes, L. Luo, T. A. Manning, and C. Monroe, Random numbers certified by Bell’s theorem, Nature 464, 1021 (2010).
- Colbeck and Kent [2011] R. Colbeck and A. Kent, Private randomness expansion with untrusted devices, J. Phys. A 44, 095305 (2011).
- Bancal et al. [2015] J.-D. Bancal, M. Navascués, V. Scarani, T. Vértesi, and T. H. Yang, Physical characterization of devices from nonlocal correlations, Phys. Rev. A 91, 022115 (2015).
- Liang et al. [2015] Y.-C. Liang, D. Rosset, J.-D. Bancal, G. Pütz, T. J. Barnea, and N. Gisin, Family of Bell-like Inequalities as Device-Independent Witnesses for Entanglement Depth, Phys. Rev. Lett. 114, 190401 (2015).
- Chen et al. [2016] S.-L. Chen, C. Budroni, Y.-C. Liang, and Y.-N. Chen, Natural Framework for Device-Independent Quantification of Quantum Steerability, Measurement Incompatibility, and Self-Testing, Phys. Rev. Lett. 116, 240401 (2016).
- Sekatski et al. [2018] P. Sekatski, J.-D. Bancal, S. Wagner, and N. Sangouard, Certifying the Building Blocks of Quantum Computers from Bell’s Theorem, Phys. Rev. Lett. 121, 180505 (2018).
- Tavakoli et al. [2021] A. Tavakoli, M. Farkas, D. Rosset, J.-D. Bancal, and J. Kaniewski, Mutually unbiased bases and symmetric informationally complete measurements in Bell experiments, Sci. Adv. 7, eabc3847 (2021).
- Wagner et al. [2020] S. Wagner, J.-D. Bancal, N. Sangouard, and P. Sekatski, Device-independent characterization of quantum instruments, Quantum 4, 243 (2020).
- Chen et al. [2021] S.-L. Chen, H.-Y. Ku, W. Zhou, J. Tura, and Y.-N. Chen, Robust self-testing of steerable quantum assemblages and its applications on device-independent quantum certification, Quantum 5, 552 (2021).
- Kaszlikowski et al. [2000] D. Kaszlikowski, P. Gnaciński, M. Żukowski, W. Miklaszewski, and A. Zeilinger, Violations of local realism by two entangled -dimensional systems are stronger than for two qubits, Phys. Rev. Lett. 85, 4418 (2000).
- Durt et al. [2001] T. Durt, D. Kaszlikowski, and M. Żukowski, Violations of local realism with quantum systems described by N-dimensional Hilbert spaces up to , Phys. Rev. A 64, 024101 (2001).
- Collins et al. [2002] D. Collins, N. Gisin, N. Linden, S. Massar, and S. Popescu, Bell Inequalities for Arbitrarily High-Dimensional Systems, Phys. Rev. Lett. 88, 040404 (2002).
- Acín et al. [2002] A. Acín, T. Durt, N. Gisin, and J. I. Latorre, Quantum nonlocality in two three-level systems, Phys. Rev. A 65, 052325 (2002).
- Barrett et al. [2006] J. Barrett, A. Kent, and S. Pironio, Maximally nonlocal and monogamous quantum correlations, Phys. Rev. Lett. 97, 170409 (2006).
- Lee et al. [2007] S.-W. Lee, Y. W. Cheong, and J. Lee, Generalized structure of Bell inequalities for bipartite arbitrary-dimensional systems, Phys. Rev. A 76, 032108 (2007).
- Liang et al. [2009] Y.-C. Liang, C.-W. Lim, and D.-L. Deng, Reexamination of a multisetting Bell inequality for qudits, Phys. Rev. A 80, 052116 (2009).
- Vértesi et al. [2010] T. Vértesi, S. Pironio, and N. Brunner, Closing the detection loophole in Bell experiments using qudits, Phys. Rev. Lett. 104, 060401 (2010).
- Salavrakos et al. [2017] A. Salavrakos, R. Augusiak, J. Tura, P. Wittek, A. Acín, and S. Pironio, Bell inequalities tailored to maximally entangled states, Phys. Rev. Lett. 119, 040402 (2017).
- Tabia et al. [2022] G. N. M. Tabia, V. S. R. Bavana, S.-X. Yang, and Y.-C. Liang, Bell inequality violations with random mutually unbiased bases, Phys. Rev. A 106, 012209 (2022).
- Thew et al. [2004] R. T. Thew, A. Acín, H. Zbinden, and N. Gisin, Bell-type test of energy-time entangled qutrits, Phys. Rev. Lett. 93, 010503 (2004).
- Dada et al. [2011] A. C. Dada, J. Leach, G. S. Buller, M. J. Padgett, and E. Andersson, Experimental high-dimensional two-photon entanglement and violations of generalized Bell inequalities, Nat. Phys. 7, 677 (2011).
- Schwarz et al. [2016] S. Schwarz, B. Bessire, A. Stefanov, and Y.-C. Liang, Bipartite Bell inequalities with three ternary-outcome measurements - from theory to experiments, New J. Phys. 18, 035001 (2016).
- Lo et al. [2016] H.-P. Lo, C.-M. Li, A. Yabushita, Y.-N. Chen, C.-W. Luo, and T. Kobayashi, Experimental violation of Bell inequalities for multi-dimensional systems, Sci Rep. 6, 22088 (2016).
- Masanes [2006] L. Masanes, Asymptotic Violation of Bell inequalities and distillability, Phys. Rev. Lett. 97, 050503 (2006).
- Pál and Vértesi [2009] K. F. Pál and T. Vértesi, Quantum bounds on Bell inequalities, Phys. Rev. A 79, 022120 (2009).
- Pál and Vértesi [2010] K. F. Pál and T. Vértesi, Maximal violation of a bipartite three-setting, two-outcome Bell inequality using infinite-dimensional quantum systems, Phys. Rev. A 82, 022116 (2010).
- Panahi et al. [2025] Y. Panahi, M. C. Alañón, D. Centeno, R. J. Costales, L. Mrini, S. Bhattacharyya, and E. Wolfe, Upper bounding Hilbert space dimensions which can realize all the quantum correlations (2025), arXiv:2505.20519 .
- Bancal et al. [2010] J.-D. Bancal, N. Gisin, and S. Pironio, Looking for symmetric Bell inequalities, J. Phys. A 43, 385303 (2010).
- Bancal et al. [2012] J.-D. Bancal, C. Branciard, N. Brunner, N. Gisin, and Y.-C. Liang, A framework for the study of symmetric full-correlation Bell-like inequalities, J. Phys. A 45, 125301 (2012).
- Fadel and Tura [2017] M. Fadel and J. Tura, Bounding the set of classical correlations of a many-body system, Phys. Rev. Lett. 119, 230402 (2017).
- Aloy et al. [2024] A. Aloy, G. Müller-Rigat, J. Tura, and M. Fadel, Deriving Three-Outcome Permutationally Invariant Bell Inequalities, Entropy 26, 816 (2024).
- Moroder et al. [2013] T. Moroder, J.-D. Bancal, Y.-C. Liang, M. Hofmann, and O. Gühne, Device-Independent Entanglement Quantification and Related Applications, Phys. Rev. Lett. 111, 030501 (2013).
- Tavakoli et al. [2024] A. Tavakoli, A. Pozas-Kerstjens, P. Brown, and M. Araújo, Semidefinite programming relaxations for quantum correlations, Rev. Mod. Phys. 96, 045006 (2024).
- Brunner et al. [2008] N. Brunner, S. Pironio, A. Acin, N. Gisin, A. A. Méthot, and V. Scarani, Testing the Dimension of Hilbert Spaces, Phys. Rev. Lett. 100, 210503 (2008).
- Navascués et al. [2014] M. Navascués, G. de la Torre, and T. Vértesi, Characterization of Quantum Correlations with Local Dimension Constraints and Its Device-Independent Applications, Phys. Rev. X 4, 011011 (2014).
- Navascués and Vértesi [2015] M. Navascués and T. Vértesi, Bounding the Set of Finite Dimensional Quantum Correlations, Phys. Rev. Lett. 115, 020501 (2015).
- Navascués et al. [2015] M. Navascués, A. Feix, M. Araújo, and T. Vértesi, Characterizing finite-dimensional quantum behavior, Phys. Rev. A 92, 042117 (2015).
- Navascués et al. [2007] M. Navascués, S. Pironio, and A. Acín, Bounding the Set of Quantum Correlations, Phys. Rev. Lett. 98, 010401 (2007).
- Navascués et al. [2008] M. Navascués, S. Pironio, and A. Acín, A convergent hierarchy of semidefinite programs characterizing the set of quantum correlations, New J. Phys. 10, 073013 (2008).
- Doherty et al. [2008] A. C. Doherty, Y.-C. Liang, B. Toner, and S. Wehner, The Quantum Moment Problem and Bounds on Entangled Multi-prover Games, in 23rd Annu. IEEE Conf. on Comput. Comp, 2008, CCC’08 (Los Alamitos, CA, 2008) pp. 199–210.
- Rosset [2018] D. Rosset, Symdpoly: symmetry-adapted moment relaxations for noncommutative polynomial optimization (2018), arXiv:1808.09598 .
- Tavakoli et al. [2019] A. Tavakoli, D. Rosset, and M.-O. Renou, Enabling Computation of Correlation Bounds for Finite-Dimensional Quantum Systems via Symmetrization, Phys. Rev. Lett. 122, 070501 (2019).
- Ioannou and Rosset [2021] M. Ioannou and D. Rosset, Noncommutative polynomial optimization under symmetry (2021), arXiv:2112.10803 .
- Clauser et al. [1969] J. F. Clauser, M. A. Horne, A. Shimony, and R. A. Holt, Proposed Experiment to Test Local Hidden-Variable Theories, Phys. Rev. Lett. 23, 880 (1969).
- Vidal and Werner [2002] G. Vidal and R. F. Werner, Computable measure of entanglement, Phys. Rev. A 65, 032314 (2002).
- Collins and Gisin [2004] D. Collins and N. Gisin, A relevant two qubit Bell inequality inequivalent to the CHSH inequality, J. Phys. A 37, 1775 (2004).
- Cruzeiro and Gisin [2019] E. Z. Cruzeiro and N. Gisin, Complete list of tight Bell inequalities for two parties with four binary settings, Phys. Rev. A 99, 022104 (2019).
- Grandjean et al. [2012] B. Grandjean, Y.-C. Liang, J.-D. Bancal, N. Brunner, and N. Gisin, Bell inequalities for three systems and arbitrarily many measurement outcomes, Phys. Rev. A 85, 052113 (2012).
- Paulsen et al. [2016] V. I. Paulsen, S. Severini, D. Stahlke, I. G. Todorov, and A. Winter, Estimating quantum chromatic numbers, J. Funct. Anal. 270, 2188 (2016).
- Rodrigues and Lackey [2017] N. Rodrigues and B. Lackey, Nonlocal games, synchronous correlations, and Bell inequalities (2017), arXiv:1707.06200 .
- Cleve et al. [2004] R. Cleve, P. Hoyer, B. Toner, and J. Watrous, Consequences and limits of nonlocal strategies, in Proceedings. 19th IEEE Annual Conference on Computational Complexity, 2004. (2004) pp. 236–249.
- Cirel’son [1980] B. S. Cirel’son, Quantum generalizations of Bell’s inequality, Lett. Math. Phys. 4, 93 (1980).
- Harris and Pandey [2016] S. J. Harris and S. K. Pandey, Entanglement and Non-locality PMATH 990/QIC 890 (2016).
- Renner [2008] R. Renner, Security of Quantum Key Distribution, Int. J. Quantum Inf. 6, 1 (2008).
- Fine [1982] A. Fine, Hidden Variables, Joint Probability, and the Bell inequalities, Phys. Rev. Lett. 48, 291 (1982).
- Wu et al. [2016] X. Wu, J.-D. Bancal, M. McKague, and V. Scarani, Device-independent parallel self-testing of two singlets, Phys. Rev. A 93, 062121 (2016).
- Kaszlikowski et al. [2002] D. Kaszlikowski, L. C. Kwek, J.-L. Chen, M. Żukowski, and C. H. Oh, Clauser-Horne inequality for three-state systems, Phys. Rev. A 65, 032118 (2002).
- Masanes [2003] L. Masanes, Tight Bell inequality for d-outcome measurements correlations, Quantum Info. Comput. 3, 345 (2003).
- Pironio [2005] S. Pironio, Lifting Bell inequalities, J. Math. Phys. 46, 062112 (2005).
- Jebarathinam et al. [2019] C. Jebarathinam, J.-C. Hung, S.-L. Chen, and Y.-C. Liang, Maximal violation of a broad class of Bell inequalities and its implication on self-testing, Phys. Rev. Research 1, 033073 (2019).
- Liang [2008] Y.-C. Liang, Correlations, Bell Inequality Violation & Quantum Entanglement, Ph.D. thesis, University of Queensland (2008).
- Popescu and Rohrlich [1994] S. Popescu and D. Rohrlich, Quantum nonlocality as an axiom, Found. Phys. 24, 379 (1994).
- Barrett et al. [2005b] J. Barrett, N. Linden, S. Massar, S. Pironio, S. Popescu, and D. Roberts, Nonlocal correlations as an information-theoretic resource, Phys. Rev. A 71, 022101 (2005b).
- Braunstein et al. [1992] S. L. Braunstein, A. Mann, and M. Revzen, Maximal violation of Bell inequalities for mixed states, Phys. Rev. Lett. 68, 3259 (1992).
- Froissart [1981] M. Froissart, Constructive Generalization of Bell’s Inequalities, Nuov. Cim. B 64, 241 (1981).
- Pitowsky and Svozil [2001] I. Pitowsky and K. Svozil, Optimal tests of quantum nonlocality, Phys. Rev. A 64, 014102 (2001).
- Śliwa [2003] C. Śliwa, Symmetries of the Bell correlation inequalities, Phys. Lett. A 317, 165 (2003).
- Brunner and Gisin [2008] N. Brunner and N. Gisin, Partial list of bipartite Bell inequalities with four binary settings, Phys. Lett. A 372, 3162 (2008).
- Goh et al. [2018] K. T. Goh, J. Kaniewski, E. Wolfe, T. Vértesi, X. Wu, Y. Cai, Y.-C. Liang, and V. Scarani, Geometry of the set of quantum correlations, Phys. Rev. A 97, 022104 (2018).
- Liang and Doherty [2007] Y.-C. Liang and A. C. Doherty, Bounds on quantum correlations in Bell-inequality experiments, Phys. Rev. A 75, 042103 (2007).
- Deza and Dutour Sikirić [2016] M. Deza and M. Dutour Sikirić, Enumeration of the facets of cut polytopes over some highly symmetric graphs, Intl. Trans. in Op. Res. 23, 853 (2016).
- Oudot et al. [2019] E. Oudot, J.-D. Bancal, P. Sekatski, and N. Sangouard, Bipartite nonlocality with a many-body system, New J. Phys. 21, 103043 (2019).
- Šupić and Bowles [2020] I. Šupić and J. Bowles, Self-testing of quantum systems: a review, Quantum 4, 337 (2020).
- Kaniewski [2020] J. Kaniewski, Weak form of self-testing, Phys. Rev. Research 2, 033420 (2020).
- Gigena and Kaniewski [2022] N. Gigena and J. Kaniewski, Quantum value for a family of -like Bell functionals, Phys. Rev. A 106, 012401 (2022).
- Liang et al. [2010] Y.-C. Liang, N. Harrigan, S. D. Bartlett, and T. Rudolph, Nonclassical correlations from randomly chosen local measurements, Phys. Rev. Lett. 104, 050401 (2010).
- Barizien and Bancal [2024] V. Barizien and J.-D. Bancal, Extremal Tsirelson Inequalities, Phys. Rev. Lett. 133, 010201 (2024).
- Liang et al. [2011] Y.-C. Liang, T. Vértesi, and N. Brunner, Semi-device-independent bounds on entanglement, Phys. Rev. A 83, 022108 (2011).
- William et al. [2025] C. William, P. Remy, J.-D. Bancal, Y. Cai, N. Brunner, and A. Pozas-Kerstjens, Symmetric observations without symmetric causal explanations (2025), arXiv:2502.14950 .
- Tura et al. [2014] J. Tura, A. B Sainz, T. Vértesi, A. Acín, M. Lewenstein, and R. Augusiak, Translationally invariant multipartite Bell inequalities involving only two-body correlators, J. Phys. A 47, 424024 (2014).
- Wang et al. [2017] Z. Wang, S. Singh, and M. Navascués, Entanglement and Nonlocality in Infinite 1D Systems, Phys. Rev. Lett. 118, 230401 (2017).
- Machnes [2007] S. Machnes, QLib - A Matlab package for quantum information theory calculations with applications (2007), arXiv:0708.0478 .
- Bowles et al. [2018] J. Bowles, I. Šupić, D. Cavalcanti, and A. Acín, Self-testing of pauli observables for device-independent entanglement certification, Phys. Rev. A 98, 042336 (2018).
- Acín et al. [2016] A. Acín, S. Pironio, T. Vértesi, and P. Wittek, Optimal randomness certification from one entangled bit, Phys. Rev. A 93, 040102 (2016).