Skip to main content
Springer logoLink to Springer
. 2024 Nov 15;405(12):292. doi: 10.1007/s00220-024-05123-2

Spin-Bounded Correlations: Rotation Boxes Within and Beyond Quantum Theory

Albert Aloy 1,2, Thomas D Galley 1,2, Caroline L Jones 1,2, Stefan L Ludescher 1,2, Markus P Müller 1,2,3,
PMCID: PMC11568060  PMID: 39554217

Abstract

How can detector click probabilities respond to spatial rotations around a fixed axis, in any possible physical theory? Here, we give a thorough mathematical analysis of this question in terms of “rotation boxes”, which are analogous to the well-known notion of non-local boxes. We prove that quantum theory admits the most general rotational correlations for spins 0, 1/2, and 1, but we describe a metrological game where beyond-quantum resources of spin 3/2 outperform all quantum resources of the same spin. We prove a multitude of fundamental results about these correlations, including an exact convex characterization of the spin-1 correlations, a Tsirelson-type inequality for spins 3/2 and higher, and a proof that the general spin-J correlations provide an efficient outer SDP approximation to the quantum set. Furthermore, we review and consolidate earlier results that hint at a wealth of applications of this formalism: a theory-agnostic semi-device-independent randomness generator, an exact characterization of the quantum (2, 2, 2)-Bell correlations in terms of local symmetries, and the derivation of multipartite Bell witnesses. Our results illuminate the foundational question of how space constrains the structure of quantum theory, they build a bridge between semi-device-independent quantum information and spacetime physics, and they demonstrate interesting relations to topics such as entanglement witnesses, spectrahedra, and orbitopes.

Introduction

Historically, quantum field theory has been developed by combining the principles of quantum theory with those of special relativity. This development has been a huge success: intersecting both theories turned out to be so constraining that it directly led to a host of novel physical predictions, such as the spin of particles and its relation to statistics, the creation and annihilation of particles, and phenomena such as Unruh radiation.

If, motivated by quantum information theory, we take an operational perspective on this development, then we can describe quantum field theory as the combination of two theories describing different phenomenological aspects of physics: our most successful theory for predicting the probabilities of events (quantum theory), and our most successful theory for describing space and time (special or general relativity). Probabilities have to interplay consistently with spacetime to yield a successful predictive theory.

While it has long been understood that special relativity describes just one possible spacetime geometry among many others, the intuition until recently has been that quantum theory is essentially our only possible choice for describing probabilities of events, except for classical probability theory. Thus, quantum field theory is defined entirely in terms of operator algebras, encompassing both classical and quantum probability theory and their hybrids, and only those.

However, motivated again by quantum information theory and by quantum foundations research, recent years have seen a surge of interest in probabilistic theories that are neither classical nor quantum. One particularly successful direction has been the device-independent (DI) framework [16] for describing quantum information protocols. The main idea is to certify the security of one’s protocols (such as quantum key distribution or randomness generation) by a few simple physical principles only. No assumptions or (in the semi-DI framework [710]) only very mild ones are made on the inner workings of the devices, and the security of the protocol follows from the observed statistics and plausible assumptions such as the no-signalling principle alone.

In this paper, we explore the foundations for studying the interplay of spacetime symmetries with the probabilities of events without assuming the validity of quantum theory. Assuming special relativity, physical systems must react to symmetry transformations (in general, Poincaré transformations) in a consistent way: the symmetry group must act continuously on its state space while preserving its structure. In quantum theory, this means that systems must carry projective representations of this group. Here, we consider more general black boxes (which need not be quantum) yielding statistics which responds to such transformations. Instead of the full Poincaré group, we study the action of one of its simplest nontrivial subgroups: the group of spatial rotations around a fixed axis, SO(2). In an abstract DI language, we study black boxes whose input is given by a spatial rotation around a fixed axis, and which produce one of a finite number of outputs. This specializes, but also greatly extends the framework introduced in [11].

In particular, we consider such “rotation boxes” under the semi-DI assumption that their “spin”, i.e. representation label of SO(2) on the ensemble of boxes, is upper-bounded by some value J. We obtain surprising insights into the structure and possible behavior of such boxes, showing, for example, that for J=0, J=1/2, and J=1, quantum theory describes the most general ways in which any theory could respond to spatial rotations, but that for J3/2, correlations exist which cannot be generated by quantum theory with the same J. We give a Tsirelson-type inequality [12] delineating the quantum correlations from more general ones, and describe a metrological task [13, 14] where post-quantum spin-3/2 systems can outperform all quantum ones. Moreover, rotation boxes can be wired together in Bell experiments, and we review and reinterpret existing work showing that our semi-DI assumption on the maximal spin can be used to certify Bell nonlocality with fewer measurements than otherwise possible, as well as to characterize the quantum-(2, 2, 2) Bell correlations exactly within the set of non-signalling correlations.

Our motivation for studying such boxes and their generalizations is threefold:

  1. Studying how spacetime structure constrains the structure of quantum theory (QT). If we assume that a probabilistic theory “fits into space and time”, does this already imply important structural features of QT? Can we perhaps derive QT from this desideratum? Or how much wiggle room is there in spacetime for probabilistic theories that go beyond quantum theory? A version of this question has been posed and studied for correlations generated by space-like separated parties, where the set of quantum correlations is known to be a strict subset of the general set of no-signalling correlations [12, 1517]. We formulate and solve an analogous question: how can we characterize the set of quantum spin-J correlations in the space of general spin-J correlations?

  2. Novel theory-independent and physically better motivated semi-DI protocols. Assumptions on the response of physical systems to spacetime symmetries can be used directly in semi-DI protocols for certification. In particular, such assumptions are sometimes physically simpler or more meaningful (corresponding to e.g. energy or particle number bounds [10, 18]) than abstract assumptions often made in the field, such as upper bounds on the Hilbert space dimension of the physical system. For example, in [19], some of us have constructed a semi-DI protocol for the generation of random numbers whose security relies on an upper bound of the system’s spin, without assuming the validity of quantum theory.

  3. The study of resource-bounded correlations. What we study in the SO(2)-case in this paper is a special case of analyzing resource-bounded correlations: given some spacetime symmetry, and an upper bound on the symmetry-breaking resources, determine the resulting correlations that quantum theory (or a more general theory) admits. The paradigmatic example is the study of quantum speed limits [2023]: upper-bounding the (expectation value or variance of the) energy constrains how quickly quantum states can become orthogonal. Replacing time-translation symmetry by rotational symmetry leads to the formalism of this paper.

Our article is organized as follows. In Sect. 2, we consider a metrological game to illustrate a gap between the predictions of quantum theory and those of hypothetical, more general theories consistent with rotational symmetry. In Sect. 3, we introduce the conceptual framework and discuss the background assumptions of rotation boxes. More specifically, in Sect. 3.1, we define and analyze the structure of the sets of quantum correlations, when the spin is constrained. In Sect. 3.2, we do so for the corresponding sets of general “rotational correlations”, when boxes are characterized only by their response to rotations (but need not necessarily be quantum). In Sect. 3.3, we discuss how, although defined independently, the rotation set can be interpreted as a relaxation of the quantum set of correlations, and show how this leads to an efficient semidefinite programming (SDP) characterization.

Next, in Sect. 4, we outline our main results, which concern rotation boxes in prepare-and-measure scenarios, and the relation between the quantum and general sets. In Sect. 4.1, we start by analyzing the scenario for the cases J{0,1/2}, for which we show that every rotation box correlation can be generated by a quantum system of the same J. In Sect. 4.2, we consider the J=1 case, and show the equivalence of the rotation and quantum sets of correlations specifically for 2 outputs, based on an exact convex characterization of this set. In Sect. 4.3, we demonstrate that a gap between the sets appears for J3/2. We construct a Tsirelson-like inequality for J=3/2 and provide an explicit correlation of rotation box form that violates the quantum bound. Using the same methodology, we further show that the gap exists for all finite J3/2. In Sect. 4.4, we examine the case where J is unconstrained (i.e. J), in which every rotation correlation can be approximated arbitrarily well by finite-J quantum systems. In Sect. 4.5, we then review our previous results [19], concerning two input rotation boxes, in which we have applied the framework to describe a theory-independent protocol for randomness generation. Finally, in Sect. 4.6, we address how one should understand a “classical” rotation box.

In Sect. 5, we consolidate earlier results concerning Bell setups using our framework. First, in Sect. 5.1, we review and shed some new light on the results of [11], which yield an exact characterization of the (2, 2, 2)-quantum Bell correlations; second, in Sect. 5.2, we clarify the additional assumption of [24] allowing for indirect witnesses of multipartite Bell nonlocality. Next, in Sect. 6, we outline connections to other known results. In particular, in Sect. 6.1, we discuss the conceptual similarity to “almost quantum” Bell correlations [25] in more depth; in Sect. 6.2, we show that the state spaces of rotation boxes are isomorphic to Carathéodory orbitopes [26]; and in Sect. 6.3, we make a connection between the effect space of the rotation GPT system and a family of rebit entanglement witnesses. Finally, we conclude in Sect. 7.

Table 1 gives a brief overview on our notation.

Table 1.

Notation used in the paper

L(V) Space of linear operators on the vector space V
LH(Cn) Space of Hermitian operators on Cn
LS(Rn) Space of symmetric operators on Rn
D(H) Set of density operators on Hilbert space H
E(H) Set of POVM elements on H
LSH(Cn) Space of symmetric Hermitian operators on Cn
Symd(V) Symmetric subspace of Vd
N Natural numbers {1,2,3,4,}
N0 Non-negative integers {0,1,2,3,4}

Invitation: A Spin-Bounded Metrological Task

Consider the following situation, which resembles a typical scenario in quantum metrology. A referee promises to perform a spatial rotation by some angle θ. Before this, we may prepare a physical system in some state, submit it to the rotation, and subsequently measure it to estimate θ. How well can we do this?

If our physical system is a classical gyroscope, we can certainly determine θ perfectly—the challenge lies in the use of microscopic systems. Think of the system as carrying some intrinsic spin J, an integer or half-integer, that responds to rotations. Classical systems correspond to the case of J, supported on an infinite-dimensional Hilbert space with narrowly peaked coherent states, allowing us to resolve the rotation arbitrarily well. Hence, consider a more interesting case: we demand that the system is a quantum spin-J system, where J is small. Concretely, let us choose J=3/2 (the smallest interesting J for this task, as we will see in subsequent sections). That is, we regard the total spin, as represented by the spin quantum number, as a resource, and are constrained in our access to such resources (Fig. 1).

Fig. 1.

Fig. 1

Schematic sketch of the metrological task. An agent holds a physical system of spin J=3/2, in an initial state ρ. She gives it to a referee, who, in a black box with respect to the agent, performs some spatial rotation of angle θ on the system, where θ is chosen according to the distribution function μ(θ) (defined in the main text and shown in Fig. 2). The referee then passes the system back to the agent, who measures it using a two-outcome box in order to determine whether the angle θ is in the range R+ or R- (see also Fig. 2)

Moreover, suppose that our task is not to estimate θ directly. Instead, our task is to guess whether θ is in region R+ or in region R-, as depicted in Fig. 2, corresponding to the sets of angles where the function cos(2θ)+sin(3θ) is either positive or negative. That is, our guess will be a single bit, + or −, and we would like to maximize our probability that this bit equals the sign of cos(2θ)+sin(3θ).

Fig. 2.

Fig. 2

The task is to estimate whether θ is in the range R+ (gray) or in the range R- (white). These ranges are defined according to where the function cos(2θ)+sin(3θ) is either positive or negative. Here we plot its normalized absolute value, which is the probability density that our referee uses to draw the angle θ in our metrological game. The ranges correspond to R+=(0,3π/10)(7π/10,11π/10)(19/10π,2π), R- is the complement R-=(3π/10,7π/10)(11/10π,19/10π)

Let us summarize the task (also sketched in Fig. 1) and specify it some more. First, the referee picks an angle θ, but not uniformly in the interval [0,2π), but according to the distribution function μ(θ):=n-1|cos(2θ)+sin(3θ)|, where n is a constant such that 02πμ(θ)dθ=1 (it turns out that n=535+25). Then, we prepare a spin-3/2 system in some state ρ and send it to the referee, who subsequently applies a rotation by angle θ to it. Finally, we retrieve the system and measure it with a two-outcome POVM (E+,E-). Our task is to produce outcome + if the angle was chosen from R+, and outcome − if the angle was chosen from R-.

This may not be the most obviously relevant task to consider, but it will serve its purpose to demonstrate an in-principle gap between quantum and beyond-quantum resources for metrology.

It turns out that the two events + and − both have probability 1/2, since

R+μ(θ)dθ=R-μ(θ)dθ=12.

But our goal is to improve upon random guessing by preparing and measuring a quantum system used for sensing in the optimal way. By the Born rule, the conditional probability of our measurement outcome is

P(±|θ)=Tr(eiθZρe-iθZE±)=c0±+c1±cosθ+s1±sinθ+c2±cos(2θ)+s2±sin(2θ)+c3±cos(3θ)+s3±sin(3θ), 1

where ρ is some quantum state, Z=diag(3/2,1/2,-1/2,-3/2) is the spin-3/2 representation of the generator of a rotation around a fixed axis, and E±0, E++E-=1 is a measurement operator. The coefficients ci±,si± can be determined from the state and measurement operator. The set of all such probability functions will be called the quantum spin-3/2 correlations, Q3/2. In fact, our construction will be more general than this: we will not define spin-J correlations as those that can be realized on the (2J+1)-dimensional irreducible representation, but on any quantum system where all outcome probabilities are trigonometric polynomials of degree at most 2J. That these correlations can always be realized on C2J+1 is a non-trivial fact which we are going to prove.

The success probability becomes

Psucc=R+P(+|θ)μ(θ)dθ+R-P(-|θ)μ(θ)dθ=R+P(+|θ)μ(θ)dθ+12-R-P(+|θ)μ(θ)dθ=02πP(+|θ)n-1cos(2θ)+sin(3θ)dθ+12=πn(c2++s3+)+12,

where we have used that, by definition, |f(θ)|=±f(θ) for θR±, where f(θ)=cos(2θ)+sin(3θ). To compute the maximum success probability PsuccQ over all spin-3/2 quantum systems, we have to determine the maximum value of c2+s3 on all quantum spin-3/2 correlations. We will do this in Sect. 4.3, showing in Theorem 7 that this maximum equals 1/3. Thus

PsuccQ=maxPQ3/2πn(c2++s3+)+12=12+3π53(5+25)0.8536.

Note that we do not allow the system to start out entangled with another system that is involved in the task. In particular, we are not considering the situation that we keep half of an entangled state and send the other half to the referee that performs the rotation. We leave an analysis of this more general situation for future work.

Now suppose that we drop the assumption that quantum theory applies to the scenario. What if we use a spin-3/2 system for sensing that is not described by quantum physics? In the following sections, we will discuss in detail how such generalized “rotation boxes” can be understood, by considering arbitrary state spaces on which SO(2) acts. In summary, a generalized spin-3/2 correlation (an element of what we denote by R3/2) will be any probability function P(±|θ) that is a trigonometric polynomial of degree three (as the second line of Eq. (1)), but without assuming that it comes from a quantum state and measurement (as in the first line of Eq. (1)).

It turns out that c2+s3 can take larger values for such more general spin-3/2 correlations, and we give an example in Theorem 7. The maximum value turns out to be 5/8. Thus, when allowing more general spin-3/2 rotation boxes, the maximal success probability is

PsuccR=maxPR3/2πn(c2++s3+)+12=12+3π85+250.8828.

Hence, general rotation boxes allow us to succeed in this metrological task with about 3% higher probability.

From a foundational point of view, tasks like the above can be used to analyze the interplay of quantum theory with spacetime structure. For example, we will see that for spins J=0,1/2,1, a gap like the above does not appear, and quantum theory is thus optimal for metrological tasks like the above. From a more practical perspective, the correlation sets RJ are outer approximation to the quantum sets QJ which have characterizations in terms of semidefinite program constraints (in mathematics terminology, the RJ are projected spectrahedra). This allows us to optimize linear functionals (such as the quantity c2+s3 above) over RJ in a computationally efficient way, yielding useful bounds on the possible quantum correlations that are achievable in these scenarios. We will see that general spin-J correlations stand to quantum spin-J correlations in a similar relation as “almost quantum” Bell correlations stand to quantum Bell correlations [25].

In the following section, we will introduce the notions of rotation boxes and spin-J correlation functions in a conceptually and mathematically rigorous way, corroborating the above analysis.

Rotation Boxes Framework

In DI approaches, one often considers quantum networks (such as Bell scenarios) where several black boxes are wired together. As sketched in Fig. 3a, a black box of this kind is typically thought of accepting an abstract input x (for example, a bit, x{0,1}) and yielding an abstract output (for example, a{-1,+1}). In QT, this could describe a measurement, where x denotes the choice of measurement and a its outcome.

Fig. 3.

Fig. 3

Boxes, rotation boxes, and the different ways to think about their physical realization. See the main text for details

In this paper, we consider boxes whose input is given by a spatial rotation around a fixed axis. The input is therefore an angle 0θ<2π. However, we do not just aim at describing generic boxes that accept continuous inputs. The intuition is not that we input a classical description of θ into the box (say, written on a piece of paper or typed on a keyboard), but rather that we physically rotate the box in space (Fig. 3b). That is, we assume that we have a notion of a physical rotation that we can apply to the box, and that this notion is a clear primitive of spatiotemporal physics. This is comparable to a Bell experiment, where we believe that we understand, in a theory-independent way, what it means to “spatially separate two boxes” (say, to transport one of them far away), such that the assumption that no information can travel faster than light enforces the no-signalling condition.

To unpack this idea further, we have to be more specific. A more detailed way to describe black boxes is in terms of a prepare-and-measure scenario: we have a preparation device which generates a physical system in some state, and a measurement device that subsequently receives the physical system and generates a classical outcome. The input x is thought of being supplied to the preparation device such that the resulting state can depend on x. Here, instead, we think of a physical operation being applied to the preparation device:

The input to the rotation box consists of rotating the preparation device by angle θ around a fixed axis, relative to the measurement device, see Fig3c.

Assuming that physics is covariant under rotations about this fixed axis leads to a representation of the SO(2) group on the state space. To see this, we follow similar argumentation to that of [27, Chapter 13]. First, consider an observer O equipped with a coordinate system and holding a k-outcome measurement device, which measures the state ωΩ transmitted by the preparation device (which need not necessarily be described by quantum theory). This produces probability tables, which can be characterized by a function PO:[0,2π)×Ω[0,1]k, such that every pair of angles and states are mapped to valid probability vectors. We assume that the outcome statistics uniquely characterize the state ω, and that Ω is finite-dimensional. Next, consider a different observer O, with their own coordinate system and k-outcome measurement device, related to O by a rotation ϕ of angle α around the fixed axis on which the input angle is defined. This reorientates the coordinate system, which induces a map ϕ^:[0,2π)[0,2π) on the set of inputs, defined by ϕ^(θ):=θ-α, i.e. relating the input angles of O to the input angles of O. According to rotational covariance, this is equivalent to a situation in which the observer O is unchanged but a state ωΩ exists such that

PO(θ,ω)=PO(ϕ^(θ),ω). 2

That is to say, there are no probabilities that could be observed in one frame that could not be observed in another (i.e there are no distinguished frames). Finally, from Eq. (2), a map ϕ¯:ΩΩ can be defined, as ϕ¯(ω):=ω. Now we consider all possible rotations around the fixed axis. This collection of rotations ϕ relating observers is isomorphic to the group SO(2), hence we label them ϕα, where α is the angle of the corresponding SO(2) rotation. From Eq. (2), it follows that

ϕ¯α1ϕ¯α2=ϕ¯α1+α2. 3

Statistical mixing of preparation procedures should be conserved under rotations, therefore every ϕ¯α must be linear (for further details, see Sect. 3.2). Therefore, these maps {ϕ¯α}α define a group representation.

Our mathematical formalism below will not depend on this specific interpretation of the SO(2)-element as a spatial rotation: it will also apply to situations where this group action has a different physical interpretation, for example as some periodic time evolution, or as some abstract transformation without any spacetime interpretation whatsoever. However, the specific scenario of preparation procedures that can be physically rotated in space gives us the clearest and perhaps most theory-independent motivation for believing that our formalism applies to the given situation. This is comparable to the study of non-local boxes [5, 6], where the no-signalling condition is usually motivated by demanding that Alice’s and Bob’s procedures are spacelike separated, but where the probabilistic formalism does not strictly depend on this interpretation. For such boxes, one might also imagine that the procedures are close-by but separated by a screening wall [28], or that the statistics just happen to not be signalling for other reasons. However, the most compelling physical situation in which non-local boxes are realized are those including spacelike separation. Similarly, the most compelling physical realizations of our rotation boxes will be via physical rotations in space.

Note that we do not need to assume a picture that is as specific as depicted in Fig. 3c: there need not literally be a “transmission of some system” from the preparation to the measurement device. We can also think of the preparation as just happening somewhere in space, and the measurement happening at the same place later in time. In this case, any time evolution happening in between the two events will be considered part of the preparation procedure. More generally, the physical transmission of the system to the measurement device can also be considered part of the measurement procedure. Furthermore, what a physical system really “is”, and whether we might want to think of it as some actual object with standalone properties, is irrelevant for our analysis.

We will make one further assumption that is often made in the semi-DI framework: essentially, that there is no preshared entanglement between the preparation and measurement devices. More generally:

The preparation and measurement devices are initially uncorrelated. That is, all correlations between them are established by the preparation procedure.

This has several important consequences, for example the following. Imagine an entangled state of two spin-1/2 particles shared between preparation and measurement devices. Suppose that the preparation device is rotated by 360, i.e. 2π. Then this may introduce a phase factor of (-1) on the preparation subsystem. After transmission to the measurement device, this relative phase can be detected. Thus, a 2π-rotation of the preparation device would induce a transformation on the physical system that does not correspond to the identity. Our assumption above excludes such behavior.

We will be interested in how the probability of the outcome can depend on this spatial rotation, i.e. in the conditional probability P(a|θ). Without any further assumptions, this probability is not constrained at all: we will see that continuity in θ is the unique assumption arising from the standard formalism of quantum theory. We will thus add a simple assumption that has often a natural realization in QT: that the physical systems which are generated by the preparation device admit an upper bound J on their SO(2)-charge, J{0,12,1,32,}. This is an abstract representation-theoretic assumption about how the physical system is allowed to react to spatial rotations. Within QT, it bounds the system’s total angular momentum quantum number relative to the measurement device. If there is no angular momentum, e.g. if we imagine sending a point particle on the axis of rotation to the measurement device as depicted in Fig. 3c, then this becomes a bound on the spin of the system. To save some ink, we will always have this idealized example in mind, and talk about “spin-bounded rotation boxes” in this paper. A more detailed definition and discussion is given in the following subsections.

Since we will only study sets of correlations that arise from upper bounds on the spin, we can always extend our preparation procedure and allow it to prepare an additional spin-0 system (i.e. a system that does not respond to spatial rotations at all) in some random choice of classical basis state. Keeping one copy and transferring the other one to the measurement device will establish shared classical randomness between the two devices, and we can imagine that this happens before the rest of the procedure is accomplished. This shows the following:

All our results remain unchanged if we allow preshared classical randomness between the preparation and measurement devices.

Mathematically, this will be reflected in the fact that all our sets of spin-bounded correlations will be convex.

Let us now turn to the mathematical description of rotation boxes of bounded spin. We will begin by assuming quantum theory, and drop this assumption in the subsequent subsection.

Quantum spin-J correlations QJ

Let us assume that the Hilbert space on which the preparation procedure acts is finite-dimensional. In quantum theory, spacetime symmetries are implemented via projective representations on a corresponding Hilbert space. It is easy to see, and shown by some of us in [19], that this implies that there is some finite set J of, either, integers (JZ={,-2,-1,0,1,2,}) or half-integers (JZ+12={,-5/2,-3/2,-1/2,1/2,3/2,5/2,}) such that the representation is

Uθ=jJ1njeijθ,

where the njN are integers. That is, the rotation by angle θ is represented by a diagonal matrix (in some basis) of complex exponentials, repeating an arbitrary number of times. Only integers or half-integers may appear, which is an instance of the univalence superselection rule which forbids superpositions of bosons and fermions.

Let us begin by writing the above in a canonical form. Setting m:=minJ and M:=maxJ as well as Δ:=(m+M)/2, we can obtain the representation Uθ:=e-iΔθUθ which acts in the same way on density matrices. It is straightforward to see that it has the form

Uθ=j=-JJ1njeijθ, 4

where nj:=nj+Δ (or zero if the latter is undefined) and J:=(M-m)/2. We stipulate that quantum spin-J rotation boxes are those that are described by projective unitary representations of this form. As always in this paper, we have J{0,12,1,32,2,}. We say that Uθ is a proper quantum spin-J rotation box if it is not also a quantum spin-(J-12) box, i.e. if nJ and n-J in (4) are both non-zero.

Quantum spin-J rotation boxes can now be described as follows. The preparation device prepares a fixed quantum state ρ. The spatial rotation of the device by angle θ maps this state to UθρUθ. Finally, the measurement device performs some measurement described by a POVM {Ea}aA, where A is the set of possible outcomes. In this paper, we are only interested in the case that A is a finite set, but this can straightforwardly be generalized.

Definition 1

The set of quantum spin-J correlations with outcome set A, where |A|2, will be denoted QJA, and is defined as follows. It is the collection of all A-tuples of probability functions

θP(a|θ)aA,

such that there exists a Hilbert space with a projective representation of SO(2) of the form (4), some quantum state (i.e. density matrix) ρ, and a POVM {Ea}aA on that Hilbert space such that

P(a|θ)=Tr(UθρUθEa).

The special case of two outcomes, A={-1,+1}, will be denoted QJ (without the A-superscript). Instead of pairs of probability functions, we can equivalently describe this set by the collection of functions P(+1|θ) only, because P(-1|θ)=1-P(+1|θ) follows from it.

Note that the integers nj in Eq. (4) can be arbitrary finite numbers, and so there is no a priori upper bound on the Hilbert space dimension on which the rotation box is represented. We can use this to prove convexity of these sets of correlations:

Lemma 1

The sets QJA are convex.

Proof

Let P,P~QJA, then

P(a|θ)=Tr(EaUθρUθ),P~(a|θ)=Tr(E~aU~θρ~U~θ)

for suitable representations, quantum states, and POVM elements. If 0λ1, we can define the block matrices

Fa:=EaE~a,σ:=λρ(1-λ)ρ~,Vθ:=UθU~θ,

such that the Fa form a POVM, σ is a density matrix, and Vθ is still a representation of the form (4). Then

λP(a|θ)+(1-λ)P~(a|θ)=Tr(FaVθσVθ),

hence λP+(1-λ)P~QJA.

At first sight, it seems as if our choice of terminology conflicts with its usual use in physics: there, a spin-J system is typically meant to describe a spin-J irrep (irreducible representation) of SU(2), living on a (2J+1)-dimensional Hilbert space. Remarkably, we will now show that we can realize all quantum spin-J correlations exactly on such systems:

Theorem 1

Let PQJA be any quantum spin-J correlation. Then there exists a pure state |ψC2J+1 and a POVM {Ea}aA on C2J+1 such that

P(a|θ)=ψ|UθEaUθ|ψ,

where Uθ:=eiθZ, with Z=diag(J,J-1,,-J). Moreover, we can choose |ψ to have real nonnegative entries in any chosen eigenbasis of Z.

In particular, without loss of generality, we can always assume that nj=1 in Eq. (4).

In other words, we can always assume that the SO(2)-rotation is given by rotations around a fixed axis of a spin-J particle in the usual sense, i.e. one that is described by a spin-J irrep of SU(2). We note that two different spin-J correlations P(a|θ) and P(a|θ) may require different orbits Uθ|ψ and Uθ|ψ as well as different POVMs to be generated.

The proof is cumbersome and thus deferred to Appendix B1. A simple consequence of Theorem 1 is that the sets QJA are compact: they arise from the compact sets of |A|-outcome POVMs and quantum states on C2J+1 under a continuous map, mapping the pair ({Ea},ρ) to the function θTr(UθρUθEa). Furthermore, multiplying out the complex exponentials in Uθ=eiθZ shows that these functions are all trigonometric polynomials of degree at most 2J (as in Lemma 5). As we show in the appendix, we can say more:

Lemma 2

The correlation sets QJA are compact convex subsets of full dimension (|A|-1)(4J+1) of the |A|-tuples of trigonometric polynomials of degree 2J or less that sum to one.

This lemma is proven in Appendix B3.

In particular, for A={+1,-1}, the set QJ is a compact subset of the trigonometric polynomials of degree at most 2J, of full dimension 4J+1.

As a simple example, consider the case of two outcomes, A={-1,+1}, and J=1/2. Then Q1/2 is a compact convex set of dimension 3. Its elements are pairs (P(+|θ),P(-|θ)). Since P(-|θ)=1-P(+|θ), we need to specify the functions P(+|θ) only, and can identify Q1/2 with this set of functions. Every such function is a trigonometric polynomial of degree one,

P(+|θ)=c0+c1cosθ+s1sinθ,

and we can depict Q1/2 by plotting the possible values of c0, c1 and s1. The result is shown in Fig. 4. Indeed, as we will show in Sect. 4.1, in this simple case, the only condition for a trigonometric polynomial of degree one to be contained in Q1/2 is that P(+|θ) gives valid probabilities, i.e. that 0P(+|θ)1 for all θ. This simple characterization will, however, break down for larger values of J, as we will see.

Fig. 4.

Fig. 4

The binary quantum spin-1/2 correlations Q1/2, which happens to be the set of trigonometric polynomials P(+|θ)=c0+c1cosθ+s1sinθ with 0P(+|θ)1 for all θ. The two endpoints are the constant zero and one functions, and the other extremal points on the circle correspond to functions θ12+12cos(θ-φ), with φ some fixed angle

Further, as we prove in the Appendix B4, the set of spin-J quantum correlations for any fixed outcome set A grows with increasing J:

Lemma 3

For all J, we have QJAQJ+1/2A.

Since dimQJA<dimQJ+1/2A, this set inclusion is strict.

In the next section, we will drop the requirement that the rotation box—or, rather, the corresponding prepare-and-measure scenario—is described by quantum theory. In order to do so, we will leave the framework of Hilbert spaces, and make use of general state spaces that could describe the scenario. To consider quantum boxes as a special case of a general scenario of this kind, we have to slightly reformulate their description: while it is convenient to consider unitary transformations acting on state vectors, quantum states are actually density matrices, and the rotations act on them by unitary conjugation, ρUθρUθ. The following lemma gives a representation-theoretic characterization of quantum spin-J boxes in terms of the way that spatial rotations act on the density matrices. This reformulation will later on allow us to motivate and derive the generalized definition of rotation boxes beyond quantum theory.

Lemma 4

Let θUθ be any finite-dimensional projective representation of SO(2). Then the following statements are equivalent:

  • (i)

    Up to global phases, the representation can be written in the form (4) with nJn-J0, i.e. it is a representation corresponding to a proper quantum spin-J rotation box.

  • (ii)

    The maximum degree of any trigonometric polynomial θTr(UθρUθE), where ρ is any quantum state and E any POVM element, equals 2J.

  • (iii)
    The associated real representation on the density matrices, θUθUθ, decomposes on the real vector space of Hermitian matrices into
    1m0k=12J1mkcos(kθ)-sin(kθ)sin(kθ)cos(kθ), 5
    where the mk are non-negative integers with m2J0. In the case where nj=1 for all j{-J,...,J}, i.e. when we have the representation on C2J+1 derived in Theorem 1, we obtain mk=2J+1-k.

This lemma is proven in Appendix B5. Let us now drop the assumption that quantum theory holds, and consider more general rotation boxes.

General spin-J correlations RJ

We now introduce the framework of spin-J rotation boxes [11, 19]. Similarly to quantum rotation boxes, a general spin-J rotation box has a preparation procedure that can be rotated by some angle θSO(2) relative to the measurement procedure, which in turn yields some output aA. The behavior of the box is given by the set of probability functions {P(a|θ)}aA, where P(a|θ):RR satisfies 0P(a|θ)1 for all θ and P(a|θ)=P(a|θ+2nπ) for all nZ.

But how can we characterize such boxes without appeal to quantum theory, and how can we say what it even means that such a box has spin at most J? Let us begin with an obvious guess for what the answer to the second question should be, before we justify this by answering the first question.

Our main observation will be that every θP(a|θ) of a quantum spin-J correlation PQJA is a trigonometric polynomial of degree at most 2J. In the characterization of the set QJA, we demand in addition that the resulting probability functions come from a quantum state and POVM together with a unitary representation of SO(2) on a Hilbert space, producing these probabilities via the Born rule. It seems therefore natural to drop the latter condition, and to only demand that the P(a|θ) are trigonometric polynomials of degree at most 2J, giving valid probabilities for all θ. This will be our definition of a general spin-J correlation, to be contrasted with the quantum version in Definition 1:

Definition 2

The set of (general) spin-J correlations with outcome set A, where |A|2, will be denoted RJA, and is defined as follows. It is the collection of all A-tuples of functions

θP(a|θ)aA

such that every one of the functions is a trigonometric polynomial of degree at most 2J in θ, and 0P(a|θ)1 as well as aAP(a|θ)=1 for all θ.

The special case of two outcomes, A={-1,+1}, will be denoted RJ (without the A-superscript). Instead of pairs of probability functions, we can equivalently describe this set by the collection of functions P(+1|θ) only, because P(-1|θ)=1-P(+1|θ) follows from it.

For concreteness, and for later use, let us denote here again what we mean by a trigonometric polynomial of degree at most 2J, and how we typically represent it:

Lemma 5

Suppose that P is a real trigonometric polynomial of degree 2J, and write it as

P(θ)=c0+j=12Jcjcos(jθ)+sjsin(jθ)=k=-2J2Jakeikθ.

Then a-j=aj¯, a0=c0, and for all j1, we have cj=2Re(aj) and sj=-2Im(aj).

This follows from a straightforward calculation.

Clearly, by construction, this notion of spin-J correlations generalizes that of the quantum spin-J correlations:

Lemma 6

Every quantum spin-J correlation is a spin-J correlation. That is, QJARJA.

The comparison of these two sets will be our main question of interest in the following sections. But first, let us return to the question of how to understand rotation boxes without assuming quantum theory, and how to obtain the notion of spin-J correlations in a representation-theoretic manner.

As will be shown, all general rotation box correlations can be generated by an underlying physical system, which may not be quantum. Non-quantum systems can be defined using the framework of Generalized Probabilistic Theories (GPTs). For an introduction to GPTs, see e.g. [2932]. A GPT system A consists of a set of states ΩA which is a convex subset of a real finite-dimensional vector space VA and a convex set of effects EAVA. We assume that ΩA and EA span VA and VA respectively. This assumption is automatically satisfied if the GPT is constructed from an operational theory, defining states as equivalence classes of preparation procedures, and effects as equivalence classes of outcomes of measurement procedures [33, 34]. The natural pairing (e,ω)[0,1] gives the probability of the measurement outcome corresponding to the effect e when the system is in state ω. A measurement is a set of effects {ei}i such that iei=u with u the unit effect, which is the unique effect such that (u,ω)=1 for all ωΩA. A transformation of a GPT system A is given by a linear map T:VAVA which preserves the set of states, T(ΩA)ΩA, and the set of effects, T(EA)EA. The linearity of these maps follows from the assumption that statistical mixtures of preparation procedures must lead to the corresponding statistical mixtures of outcome probabilities, for all possible measurements after the transformation. The set of all transformations of the system A is given by a closed convex subset of the linear space L(VA) of linear maps from VA to itself.

The set of reversible transformations Rev(A) corresponds to those transformations T for which T-1 exists and is also a transformation. It forms a group under composition of linear maps. If there exists a group homomorphism GRev(A) (i.e. a representation of G) for some group G then G is said to be a symmetry of A. In this spirit, the set {ϕ¯α}α of Sect. 3 (or, more precisely, the linear extensions of those maps) are an SO(2) symmetry of the GPT system that describes the scenario. If, given a GPT system (A,ΩA,EA) with an SO(2) symmetry θTθ, with TθRev(A), then the probability distribution P(a|θ)=(ea,Tθω) is a rotation box correlation. In this case, we say that the correlation P(a|θ) can be generated by the GPT system A.

Lemma 7

Consider any finite-dimensional GPT system A=(VA,ΩA,EA), together with a representation of  SO(2), θTθ, such that every Tθ is a reversible transformation. Then the following are equivalent:

  • (i)

    The maximum degree of any trigonometric polynomial θ(e,Tθω), where ωΩA is any state and eEA any effect, equals 2J.

  • (ii)
    The real representation θTθ of SO(2) decomposes on the real vector space A into
    1m0k=12J1mkcos(kθ)-sin(kθ)sin(kθ)cos(kθ), 6
    where the mk are integers with m2J0.

If one of these two equivalent conditions is satisfied, we call the GPT system a spin-J GPT system.

Proof

Since θTθ is a representation of SO(2) on the real vector space VA, it can be decomposed into irreps. In some basis, this gives us the representation Tθ=1m0k=1n1mkcos(kθ)-sin(kθ)sin(kθ)cos(kθ), for some finite integer n, where mn0. Now since ΩA spans VA and EA spans VA, the linear functionals T(e,Tω) span L(VA), where L(VA) is the set of linear operators on VA. In other words, there will be some real numbers αi, effects ei and states ωi such that iαi(ei,Tθωi) yields the component cos(nθ), and this is only possible if θ(e,Tθω) is a trigonometric polynomial of degree at least n for some effect e and state ω. But the degree of this trigonometric polynomial can of course not be higher than n.

This characterization resembles Lemma 4 for the quantum case: it tells us that quantum spin-J rotation boxes are spin-J GPT systems. And it allows us to obtain a justification for our definition of spin-J correlations:

Theorem 2

Let PP(a|θ)aA be an A-tuple of functions in θ. Then the following are equivalent:

  • (i)

    P is a spin-J correlation, i.e. PRJA.

  • (ii)

    There is a spin-J GPT system (VA,ΩA,EA) with a state ωΩA and measurement {ea}aAEA such that P(a|θ)=(ea,Tθω).

Proving the implication (ii)(i) is immediate, given Lemma 7. For the converse implication, we will now show how all correlations in RJA can be reproduced in terms of a single GPT system that we will call RJ:

Definition 3

(Spin-J rotation box system RJ). Let RJ be a GPT system with state space ΩJR4J+1 and effect space EJR4J+1 defined as follows:

ΩJ=conv{ωJ(θ)|θ[0,2π)}, 7

with

ωJ(θ)=1cos(θ)sin(θ)cos(kθ)sin(kθ)cos(2Jθ)sin(2Jθ), 8

and

EJ:={eR4J+1|e·ω[0,1]forallωΩJ}. 9

The unit effect is

u=(1,0,...,0). 10

The system RJ carries a representation SO(2)L(R4J+1), θTθ of SO(2), given by

Tθ=k=02Jγk(θ), 11
γ0(θ)=1, 12
γk(θ)=cos(kθ)-sin(kθ)sin(kθ)cos(kθ),k{1,...,2J}. 13

The system RJ is an unrestricted system by definition. These systems belong to the family of GPT systems with pure states given by the circle S1 and reversible dynamics SO(2); i.e. for J1, they can be interpreted as rebits with modified measurement postulates [35]. The state space ΩJ is the convex hull of an SO(2) orbit of the vector ω(0)R4J+1 and is hence an SO(2) orbitope [36].

The system RJ is canonical in the sense that the SO(2) correlation set it generates is exactly RJA, as shown in the following lemma:

Lemma 8

The set of spin-J correlations RJA can be generated by the system RJ: for every PRJA, there is a measurement {ea}aA on RJ with

P(a|θ)=ea·ωJ(θ).

Conversely, every tuple of probability functions (P(a|θ))aA generated in this way with measurements in RJ is in RJA.

Proof

The set RJ is given by all functions P:θ[0,1] of the form P(θ)=c0+j=12J(cjcos(jθ)+sjsin(jθ)). This can be expressed as

P(θ)=e·ωJ(θ), 14

where ωJ(θ) is defined as in Eq. (8) and e=(c0,c1,s1,...,c2J,s2J). e is an effect on the system RJ since by construction e·ωJ(θ)[0,1], which in turn implies e·ω[0,1] for all ωconv{ωJ(θ)|θ[0,2π)}=ΩJ. This show that any P(θ) can be generated using the orbit of states {ωJ(θ)|θ[0,2π)}.

Given a tuple (P(a|θ))aARJA, we show that it can be generated by a measurement {ea}aA applied to the orbit ΩJ(θ).

P(a|θ) is a function θ[0,1] of the form P(a|θ)=c0a+j=12J(cjacos(jθ)+sjasin(jθ)). The requirement aAP(a|θ)=1 for all θ implies that

ac0a+j=12Jcjacos(jθ)+sjasin(jθ)=1, 15

which in turn entails

ac0a=1,acja=asja=0(1j2J). 16

Every P(a|θ)=ea·ωJ(θ) for ea=(c0a,c1a,s1a,...,c2Ja,s2Ja) which is a valid effect. Moreover, the conditions of Eq. (16) entail that aAea=u with u the unit effect. Hence {ea}aA form a measurement.

Conversely, consider an arbitrary tuple (P(a|θ))aA of SO(2) probability functions generated by RJ:

P(a|θ)=ea·Tθω, 17

where aAea=u and ωΩJ. Since TθL(R4J+1) P(a|θ) is a linear functional, L(R4J+1)R and hence in L(R4J+1). This implies that P(a|θ) is a linear combination of entries in Tθ and therefore a trigonometric polynomial of order at most 2J. Hence P(a|θ)RJ.

The condition aea=u implies

aAP(a|θ)=aea·Tθω=u·Tθω=1 18

Thus (P(a|θ))aARJA.

It follows from the proof of the above lemma that the effect space EJ is isomorphic to RJ as a convex set.

General spin-J correlations as a relaxation of the quantum set

The space of spin-J correlations RJ is defined independently of the quantum formalism, however it can also be interpreted as arising from a relaxation of the quantum formalism.

To see that, we start by noting the Fejér–Riesz theorem [37], which has several important applications for quantum and general rotation boxes:

Theorem 3

(Fejér–Riesz theorem). Suppose that P(θ):=j=-2J2Jajeijθ satisfies P(θ)0 for all θ. Then there is a trigonometric polynomial Q(θ):=j=-JJbjeijθ such that P(θ)=|Q(θ)|2.

From this, we can easily derive the following Lemma:

Lemma 9

Let P(θ)=j=-2J2Jajeijα be a trigonometric polynomial. Then we have P(θ)0 for all θR if and only if there exists a vector b=(b0,b1,,b2J)C2J+1 such that

ak=0j,j+k2Jbj¯bj+k.

Note that necessarily

b2=a0=12π02πP(θ)dθ,

and the matrix Qjk:=bj¯bk is positive semidefinite. Consequently, the following theorem follows from Fejér–Riesz’s theorem:

Theorem 4

If PRJ, then there is a pure quantum state |ψ on C2J+1 and a positive semidefinite matrix E+0 such that

P(+|θ)=ψ|UθE+Uθ|ψ.

We can always choose |ψ as the uniform superposition |ψ:=(2J+1)-1/2j=-JJ|j, Uθ as defined in Theorem 1, and E+=(2J+1)|bb|, where |b is the vector from Lemma 9. Note, however, that E+ is not in general a POVM element, i.e. it will in general have eigenvalues larger than 1.

Proof

Let P(+|θ)=j=-2J2JajeijθRJ, then by Theorem 3:

P(+|θ)=j=-JJb¯je-ijθk=-JJbkeikθ. 19

Now use Uθ as defined in Theorem 1, with orthonormal basis {|j}j=-JJ such that Uθ|j=eijθ|j, and define |b=jbj|j. Then

P(+|θ)=b|Uθj|jkk|Uθ|b=ψ|UθE+Uθ|ψ,

where |ψ:=(2J+1)-1/2j=-JJ|j and E+=(2J+1)|bb|.

Therefore, rotation boxes can be regarded as a relaxation of the quantum formalism: instead of demanding that E+ gives valid probabilities on all states (which would imply 0E+1), the above only demands that it gives valid probabilities on the states of interest, i.e. on the states Uθ|ψ for all θ and some fixed state |ψ. This is strikingly similar to the definition of the so-called almost quantum correlations [25]: for these, one demands that the operators in a Bell experiment commute on the state of interest and not on all quantum states, which gives a relaxation of the set of quantum correlations.

Moreover, Theorem 4 entails that RJ is isomorphic to the linear functionals on conv{Uθ|ψψ|Uθ|θ[0,2π)} giving values in [0, 1]. As discussed in Sect. 6.2, this entails that conv{Uθ|ψψ|Uθ|θ[0,2π)} is isomorphic to the orbitope ΩJ. This isomorphism gives a characterization of ΩJ as a spectrahedron.

That rotation boxes represent a relaxation of the quantum formalism can also be seen by noting the following Lemma which later will be contrasted with its quantum counterpart (Lemma 11):

Lemma 10

Let P(+|θ):=j=-2J2Jajeijθ be a trigonometric polynomial of degree 2J. Then PRJ if and only if there exist positive semidefinite (2J+1)×(2J+1)-matrices Q,S0 such that

  • ak=0j,j+k2JQj,j+k,

  • 1-a0=Tr(S),

  • ak=-0j,j+k2JSj,j+k for all k0.

The first condition implies that 0P(+|θ) for all θR, and the last two constraints guarantee that P(+|θ)1 for all θR. The proof of this lemma is a straightforward application of Lemma 9 and can be found in Appendix B6.

Remarkably, the constraints in Lemma 10 can be adapted into a semidefinite program (SDP) [38]. For instance, imagine we want to find the boundary of the coefficient space of spin-J rotation boxes in some direction nR4J+1 of the trigonometric coefficients space. That is, we want to find the maximal value of f(c,s)=n·(c,s), where c,sR2J+1 are vectors c=(c0,,c2J), s=(s1,,s2J) collecting the trigonometric coefficients leading to valid rotation boxes. Then, one can pose the following SDP:

maxQ,Sf(c,s)s.t.ak=0j,j+k2JQj,j+kfor allk,ak=-0j,j+k2JSj,j+kfor allk0,1-a0=Tr(S),Q,S0, 20

where the entries of QS are labelled from 0 to 2J. For example, for J=1 the first condition above becomes

a-2=Q2,0a-1=Q1,0+Q2,1a0=Q0,0+Q1,1+Q2,2a1=Q0,1+Q1,2a2=Q0,2.

As we show in Appendix B2, the SDP formulation in (20) can be easily generalized to account for an arbitrary finite number of outcomes, i.e. for the analysis of RJA with |A|3. In Sect. 4 we use the SDP methodology in (20) to efficiently derive hyperplanes that bound the set of spin-J rotation boxes (and thus also the set of spin-J quantum boxes). These hyperplanes can be treated as inequalities which, if violated, ensure that the system being probed has spin larger than the J considered.

Suppose now that we are not interested in optimizing some quantity restricted to RJ, but rather we are given a list of coefficients a~ (perhaps by an experimentalist) and we want to know whether these lead to a valid spin-J correlation. Then, one can recast the SDP formulation as a feasibility problem (see, e.g., [39]) by setting the given coefficients as constraints. That is, we are now interested in the following problem:

findQandSs.t.a~k=0j,j+k2JQj,j+kfor allk,a~k=-0j,j+k2JSj,j+kfor allk0,1-a~0=Tr(S),Q,S0, 21

where, contrary to (20), the coefficients a~k are now fixed. If the SDP is feasible, then it will give (2J+1)×(2J+1) matrices Q,S0 certifying that a~ leads to a valid spin-J correlation (c.f. Lemma 10). Conversely, if the SDP is infeasible, then one can obtain a certificate that the given coefficients a~ cannot lead to a valid spin-J correlation (again see, e.g., [39]).

We have already noted above that there is a conceptual similarity between general spin-J correlations (as a relaxation of quantum spin-J correlations) and “almost quantum” Bell correlations [25] (as a relaxation of the quantum Bell correlations). Here we see another aspect of this analogy: the set of almost-quantum Bell correlations has an efficient SDP characterization (derived from the NPA hierarchy [40]), but the set of quantum correlations does not. Similarly, as shown above, general spin-J correlations have an efficient SDP characterization, but we do not know whether quantum spin-J correlations QJA have an SDP characterization, for arbitrary J and A.

In particular, the quantum counterpart of Lemma 10 is the following:

Lemma 11

Let P(+|θ):=j=-2J2Jajeijθ be a trigonometric polynomial of degree 2J. Then PQJ if and only if there exists a positive semidefinite (2J+1)×(2J+1)-matrix Q0 such that

  • ak=0j,j+k2JQj,j+k,

  • Q is the Schur product of a density matrix and a POVM element, i.e. there exist 0E1 and 0ρ with Tr(ρ)=1 such that Qi,j=Ei,jρi,j (denoted Q=Eρ).

The proof follows directly from Theorem 1 and the Born rule, P(+|θ)=Tr(ρUθEUθ). Note that the second condition, the Schur product of ρ,E0, breaks the linearity required for an SDP formulation in the general case where both ρ,E act as free optimizing variables. Nonetheless, for numerical purposes, one may be interested in circumventing this limitation by adopting a see-saw scheme [41, 42] at the cost of introducing local minima in the optimization problem. The see-saw methodology consists in linearizing the problem by fixing one of the free variables and optimizing only over the other free variable. Then, fix the obtained result and optimize over the variable that had been previously fixed. One would iteratively continue this procedure until the objective function converges to a desired numerical accuracy.

For example, in our case, one could start by picking a random quantum state ρ and use an SDP with the conditions in Lemma 11 to find the optimal POVM E for that given ρ. Then, fix the POVM to the new-found E and proceed to optimize using ρ as a free variable in order to update the quantum state to a new more optimal value. One would continue this procedure until eventually the increment gained at each iteration would be negligible. However, as opposed to a general SDP, this approach does not guarantee that a global minimum has been attained due to the possible presence of local minima. To guarantee that a global minimum has been obtained, one has to provide a certificate of optimality (for instance, by means of the complementary slackness theorem [38]).

Rotation Boxes in the Prepare-and-Measure Scenario

So far, we have defined quantum and more general spin-J correlations, QJA and RJA, describing how outcome probabilities can respond to the spatial rotation of the preparation device in a prepare-and-measure scenario. But how are these two sets related? Do they agree or is there a gap? Can all possible continuous functions P(+|θ) be realized for large J? What can we say in the special case of restricting to two possible input angles only, and what is the correct definition of a “classical” rotation box? In this section, we answer all these questions, and we review earlier work by some of us [19], which shows how the results can be applied to construct a theory-agnostic semi-device-independent randomness generator.

Q0A=R0A and Q1/2A=R1/2A

In this subsection, we will see that all the spin-J correlations for J=0 and J=1/2 have a quantum realization. That is, for every PR0A (resp. PR1/2A), we can find a spin-0 (resp. spin-1/2) quantum system, a quantum state ρ, and a POVM {Ea}aA such that P(a|θ)=tr(UθρUθEa).

First, we consider J=0. In this case the set of rotation boxes corresponds to all sets with cardinality |A| of constant functions between zero and one summing to one, i.e. PR0A is given by P(a|θ)=ca for all θ[0,2π), where 0ca1 and a=1|A|ca=1. In the quantum case, we consider a representation Uθ of SO(2) consisting of the direct sum of |A| copies of the trivial representation, i.e. Uθ=1|A|. Now, to realize PR0A, we pick an orthonormal basis {ϕa}a=1|A| and construct the state |ψ=a=1|A|ca|ϕa, such that P(a|θ)=|ϕa|Uθψ|2=|ϕa|1|A|ψ|2=ca for every aA and therefore Q0A=R0A.

Next, we will turn our attention to the first non-trivial case, i.e. to J=1/2.

Theorem 5

The correlation set R1/2A is equal to Q1/2A, i.e. Q1/2A=R1/2A.

Proof

We recall (see Definition 3) that the state space of the GPT system R1/2 generating R1/2A is given by

Ω1/2:=conv1cos(θ)sin(θ)|θ[0,2π), 22

and that R1/2 is unrestricted. Next, we will show that the state space Ω1/2 can be identified with the state space of a rebit, which follows from the fact that every pure rebit state ρD(R2)LS(R2), where LS(R2) is the space of real symmetric 2×2- matrices, can be written as

ρ=121+cos(θ)σx+sin(θ)σz, 23

with the Pauli matrices σx and σz. Hence, we define the bijective linear map L:R3LS(R2) by

r0r1r312(r01+r1σx+r3σz). 24

Since R1/2 and the rebit are both unrestricted [43], we can map the effects of R1/2 one to one to the effects of the rebit via the map (L-1):(R3)(LS(R2)). Furthermore, the system R1/2 carries the representation Tθ:

Tθ=1000cos(θ)-sin(θ)0sin(θ)cos(θ). 25

Using the map L again, we can define the SO(2)-representation U on the rebit by U[θ]=LTθL-1. Applied to ρLs(R2), this family of transformations acts as

U[θ](ρ)=UθρUθ, 26

where

Uθ=exp(iθ2σy)=cos(θ2)sin(θ2)-sin(θ2)cos(θ2). 27

Now, let PR1/2A and let ωΩ1/2 and {ea}a=1|A|E1/2 be the state and measurement generating P. We show

P(a|θ)=(ea,Tθω)R3=(ea,L-1LTθL-1Lω)R3=(L-1)ea,LTθL-1LωHS=Ea,U[θ](ω)HS=Tr(EaUθωUθ), 28

where (·,·)R3 and ·,·HS denote the standard inner product in R3 and Hilbert-Schmidt product, respectively, and the Ea=(L-1)ea and ω=Lω are a rebit effect and a rebit state, respectively.

For a characterization of the extreme points of R1/2, see [19] and Fig. 4 above.

The convex structure of R1 and Q1=R1

For clarity, we write the general form of spin-J correlations of Definition 2 in the case J=1. The set R1 of correlations generated by spin-1 rotation boxes consists of all probability distributions P(+|):RR of the following form:

P(+|θ)=c0+c1cosθ+s1sinθ+c2cos(2θ)+s2sin(2θ), 29

where c0,c1,s1,c2,s2R and 0P(+|θ)1 for all θ.

Characterizing the facial structure of R1

We now characterize some of the properties of the convex set R1. Our main goal is to characterize the extreme points of R1, which will then allow us to obtain explicit quantum realizations of these extreme points and hence of all of R1. For θ0[0,2π) we define the following face of R1:

Fθ0:={PR1|P(θ0)=0}. 30

The condition P(θ0)=0 defines a hyperplane in the space of coefficients (c0,c1,s1,c2,s2)R5. Since it is a supporting hyperplane of R1, its intersection with this compact convex set is a face. For some background on convex sets, their faces, and other convex geometry notions used in this section, see e.g. the book by Webster [44].

Lemma 12

The face Fθ0 has dimension dim(Fθ0)3 for every θ0[0,2π).

Proof

For every PFθ0 it must be the case that P(θ0) is a minimum, since P(θ)0. This implies that P(θ0)=ddθP(θ)|θ=θ0=0. Thus, we obtain two linearly independent constraints

P(θ0)=0,P(θ0)=0, 31

and the face Fθ0 is at most three-dimensional.

For θ0,θ1[0,2π), we define the following subsets of Fθ0:

Fθ0,θ1:={PR1|P(θ0)=0,P(θ1)=1}. 32

Every non-empty Fθ0,θ1 is a face of Fθ0 and therefore of R1 (and thus itself compact and convex). Denote the extremal points of a compact convex set C by extC.

Lemma 13

Every non-constant function PextR1 is contained in at least one face Fθ0,θ1.

This lemma is proven in Appendix C1.

If P is extremal in R1, then it is also extremal in every face in which it is contained. Thus, we can determine the extremal points of R1 by determining extFθ0,θ1 (and keeping in mind that the functions which are constant, P(θ)=0 for all θ and P(θ)=1 for all θ, are also extremal in R1).

Next, note that it is sufficient to determine the extremal points in the case that θ0=0. This is because

P(θ)Fθ0,θ1P(θ+θ0)F0,θ1-θ0.

Hence Fθ0,θ1 and F0,θ1-θ0 are related by a linear symmetry Tθ0 of R1, which is defined by

Tθ0(P)(θ):=P(θ+θ0).

That is, Tθ0:R1R1 is a convex-linear map that rotates every rotation box by angle θ0. Since it is a symmetry of R1, it maps extremal points of faces to extremal points of faces. To determine extFθ0,θ1, we only need to “rotate” extF0,θ1-θ0 by θ0.

We now explicitly characterize the faces F0,θ1 by the functions corresponding to their extremal points.

Lemma 14

The faces F0,θ1 for θ1[0,2π) are characterized as follows:

  1. If θ1[0,π2)(3π2,2π), then
    F0,θ1=.
  2. If θ1{π2,3π2}, then F0,θ1 contains a single element:
    F0,π2=F0,3π2=P(θ)=sin2θ.
  3. If θ1π2,3π2\{π}, then F0,θ1 contains exactly two distinct extremal points,
    extF0,θ1={P(θ),P~(θ)},
    where
    P(θ)=c(1-cosθ)(1-cos(θ-θ0)),P~(θ)=1-P(θ1-θ),
    and θ0=2θ1 for θ1(π2,π) and θ0=2(θ1-π) for θ1(π,3π2). The parameter c>0 is uniquely determined by the condition maxθP(θ)=1.
  4. If θ1=π then the face F0,π contains exactly two extremal points, namely
    F0,π={P(θ),P~(θ)},
    where
    P(θ)=sin4θ2,P~(θ)=1-P(θ1-θ)=14(1-cosθ)(3+cosθ).

This lemma is proven in Appendix C2.

In Fig. 5, we plot the face F0 in the coefficients space, illustrating the resulting extremal points from Lemma 14. Note that from the conditions (31) for θ0=0, one has c0=-c1-c2 and s1=-s2, thus dimF0=3.

Fig. 5.

Fig. 5

Different perspectives of the set containing the associated trigonometric coefficients of the face F0 of the binary spin-1 correlations R1, and its extremal points from Lemma 14. The red and yellow lines correspond to the two consecutive extremal points for F0,θ1 with θ1π/2,3π/2, the pink dot corresponds to the case F0,π/2=F0,3π/2, and the green and cyan dots correspond to the two consecutive cases for F0,π

Quantum realizability of R1

Having characterized the facial structure of R1 and its extremal functions, we now ask if this set of correlations can be realized by a quantum spin-1 system.

By Theorem 1, the space Q1 of SO(2)-correlations generated by a quantum spin-1 system is given by the functions P(+|θ)=ψ|UθE+Uθ|ψ, where |ψC3, E+ a POVM element on C3, and Uθ=eiZθ with Z=diag(1,0,-1).

It follows immediately from the convexity of R1 and of Q1 that it is sufficient to show that the extremal points of R1 are quantumly realizable to show that all the correlations in R1 are quantumly realizable.

Lemma 15

δextR1Q1 implies R1=Q1.

This will be used to prove the main result of this subsection:

Theorem 6

(Q1=R1) The correlation set R1 is equal to Q1.

For the proof, see Appendix C3. It follows from constructing explicit quantum spin-1 realizations of all the extremal points of R1 which have been enumerated in Lemma 14.

Although the correlation spaces R1 and Q1 are equal, the J=1 general rotation box system R1 (which generates R1) is not equivalent to a quantum spin-1 system. This can be seen immediately from the fact that R1 is a 5-dimensional GPT system, while a quantum spin-1 system is a 9-dimensional system (since dim(LH(C3))=9).

In the next section, we will see that these two GPT systems, although they generate equivalent SO(2)-correlations, have distinct informational properties.

Inequivalence of spin-1 rotation box system and quantum system

Every PR1A can be decomposed in the following way:

P(a|θ)=c0(a)+c1(a)cos(θ)+s1(a)sin(θ)+c2(a)cos(2θ)+s2(a)sin(2θ)=c0(a),c1(a),s1(a),c2(a),s2(a)·1cos(θ)sin(θ)cos(2θ)sin(2θ)=ea·ω(θ), 33

where ea and ω(θ) are an effect and state of the spin-1 rotation box system R1, as defined in Definition 3 for general spin-J.

We give an explicit definition of the R1=R5,Ω1,E1 GPT system here. The state space Ω1 is given by:

Ω1:=convω(θ)|θ[0,2π)],

where

ω(θ)=1cos(θ)sin(θ)cos(2θ)sin(2θ).

Let VR5 be the real linear span of Ω and V its dual space. The effect space of R1 is

E1:={eV|0(e,ω)1forallωΩ1}.

By definition, R1 is an unrestricted GPT. The state space Ω1 belongs to a family of SO(2)-orbitopes of the form Ca,b:=conv{(1,cos(aθ),sin(aθ),cos(bθ),sin(bθ)|θ[0,2π)} for integers a<b. The facial structure of these orbitopes was studied in [45]. They are a subset of the Carathéodory orbitopes defined in Sect. 6.2. The SO(2) reversible transformations are given by

T(θ)=100000cos(θ)-sin(θ)000sin(θ)cos(θ)00000cos(2θ)-sin(2θ)000sin(2θ)cos(2θ). 34
Lemma 16

The effect space E1 is isomorphic (as a convex set) to R1, i.e. there is an invertible linear map that maps one of these sets onto the other.

Proof

The effect space E1 consists of all (c0,c1,s1,c2,s2)R5 such that Eq. (33) is in [0, 1] for all θ[0,2π). This is equivalent to the condition that P(+|θ)[0,1] for all θ which defines R1 in Eq. (29).

We now describe some informational properties of R1:

Lemma 17

(Properties of R1). The GPT system R1

  1. has three jointly perfectly distinguishable states and no more;

  2. has four pairwise perfectly distinguishable states;

  3. violates bit symmetry.

This lemma is proven in Appendix C4.

Bit symmetry is the property that any pair of perfectly distinguishable pure states (ω0,ω1) of a GPT system can be reversibly mapped to any other pair of perfectly distinguishable pure states (ω0,ω1) of that system [46]. Namely, there exists a reversible transformation T such that (ω0,ω1)=(Tω0,Tω1).

We note that R1 violates bit symmetry not just for the set of SO(2) reversible transformations but for the set of all symmetries. This set is larger than the SO(2) transformations of Eq. (34) and includes the transformation diag(1,1,-1,1,-1) which is not of the form T(θ).

Considering the full set of symmetries is important when contrasting to a qutrit, since the qutrit when restricted to the spin-1 SO(2)-transformations violates bit symmetry, but it obeys bit symmetry when considering the full symmetry group SU(3).

Although the space of correlations R1E1, the GPT system R1 contains additional structure, namely in its state space Ω1. Hence, although every P(+|θ)R1 can be generated using a quantum system Q1, this does not imply that every information-theoretic game carried out using the system R1 can be equally successfully carried out with a spin-1 quantum system. For instance, a game which required one to encode a pair of bits (i,j){0,1}2 in four states of a GPT system such that one could perfectly decode either the first bit or the second bit can be implemented with R1 with 100% success probability, but will necessarily have some error when implemented on a quantum spin-1 system.

A key difference between the the GPT system R1 and the SO(2) quantum spin-1 system Q1 (i.e. a qutrit with dynamics restricted to Uθ=eiZθ) is that inequivalent SO(2)-orbits of pure states of the qutrit are needed to generate R1, whilst a single SO(2)-orbit of states {ω(θ)|θ[0,2π)} of R1 is needed to generate R1.

A formal way to understand the equivalences and inequivalences of RJ and QJ for different values of J is in terms of linear embeddings [47]. We say that a GPT A=(VA,ΩA,EA) can be embedded into a GPT B=(VB,ΩB,EB) if there is a pair of linear maps Φ,Ψ such that Ψ(ΩA)ΩB and Φ(EA)EB which reproduces all probabilities, (Φ(eA),Ψ(ωA))=(eA,ωA) for all eAEA,ωAΩA. As argued in [47], this means that B can simulate the GPT A “univalently”, i.e. in a way that generalizes the concept of noncontextuality for simulations by classical physics.

In the proof that Q1/2A=R1/2A in Sect. 4.1, we have used the fact that the spin-1/2 GPT system R1/2 (the rebit) can be embedded into the qubit Q1/2, seen as a quantum spin-1/2 system. Moreover, it can be done in a way such that the orbit θω(θ) is mapped to an orbit ρ(θ)=Ψ(ω(θ)). That is, the quantum system can reproduce the full probabilistic behavior of the general spin-1/2 system.

However, it is easy to see that no such embedding can exist for the case of J=1. If we had such a pair of linear maps, and if it mapped the orbit ω(θ) to some orbit ρ(θ), then it could not reproduce all probabilities: it would give us four states ρ(0),ρ(π2),ρ(π),ρ(3π2) of the qutrit which are pairwise perfectly distinguishable. But no four pairwise orthogonal states can exist on a qutrit. Clearly, the converse is also true: The spin-1 quantum system Q1 spans the vector space LH(C3)R9 and hence cannot be embedded in the GPT system R1 which spans R5. More generally, we can say the following:

Lemma 18

The spin-1 GPT system R1 cannot be embedded into any finite-dimensional quantum system.

Proof

According to Theorem 2 of [47], all unrestricted GPTs that can be so embedded are special Euclidean Jordan algebras. For all such systems, the numbers of jointly and pairwise perfectly distinguishable states coincide. This can be seen e.g. by noting that perfectly distinguishable pure states in Euclidean Jordan algebras are orthogonal (with respect to the self-dualizing inner product) idempotents (see e.g. [48, Lemma 3.3]), and pairwise orthogonality implies that they are elements of a Jordan frame and hence jointly perfectly distinguishable. But as we have shown in Lemma 17 above, this correspondence does not hold for R1.

Hence, even though the set of spin correlations R1 and Q1 agree, the corresponding GPT systems have genuinely different information-theoretic and physical behaviors. This is also the reason why we do not currently know whether Q1A=R1A for |A|3.

QJRJ for J3/2

Up until now we have seen that for J1 an equivalence holds between the correlation sets QJ and RJ. However, in this section we show that this equivalence breaks for J3/2. We split the analysis in two parts: First, we provide an explicit counterexample of a spin-J correlation outside of the quantum set for J=3/2; Second, we use the same methodology to show that a non-empty gap exists between both sets for any J3/2.

Q3/2R3/2

We start by showing that Q3/2R3/2. Every spin-3/2 correlation can be expressed as a degree-3 trigonometric polynomial:

P(θ)=c0+c1cosθ+s1sinθ+c2cos(2θ)+s2sin(2θ)+c3cos(3θ)+s3sin(3θ), 35

where the coefficients ci and si are suitable real numbers such that 0P(θ)1 for all θ. To show that there exist correlations PR3/2 which are not contained in Q3/2, we construct an inequality that is satisfied by all quantum boxes, but violated by some PR3/2. In particular, we show the following:

Theorem 7

If PQ3/2, then its trigonometric coefficients, as taken from representation (35), satisfy

c2+s3130.5774.

On the other hand, the trigonometric polynomial

P(θ):=25+14sinθ+720cos(2θ)+14sin(3θ)

satisfies 0P(θ)1 for all θ, hence PR3/2, but c2+s3=0.6, i.e. PQ3/2. Therefore, Q3/2R3/2.

Clearly, this also implies that Q3/2AR3/2A for three or more outcomes, k:=|A|3, since P can always appear as the probability of the first of the k outcomes.

In the remainder of this section, we prove this theorem by solving the optimization problem

β:=maxPQ3/2(c2+s3)[P], 36

and show that the quantum bound is β=13. Since (c2+s3)[P]=35, P violates the inequality, thus proving Q3/2R3/2. For the sake of completion, by adapting the SDP in Eq. (20) one can show that the maximal value attainable with rotation boxes is βR=maxPR3/2(c2+s3)[P]=58=0.625, hence β<(c2+s3)[P]<βR. In Fig. 6 we illustrate Theorem 7 by showing the 2D projection of the correlation sets onto the c2-s3 plane and plotting the inequality given by c2+s31/3 as well as the point P violating it.

Fig. 6.

Fig. 6

Spin-3/2 rotation and quantum correlations sets in the c2-s3 plane projection illustrating Q3/2R3/2. The inequality corresponds to the case that saturates Theorem 7, i.e., c2+s3=1/3. The boundary of the 2D projections for the sets Q3/2 (blue) and R3/2 (green) have been numerically obtained using the SDP methodology presented in Appendix D. The quantum inequality (red line) and validity of the rotation box (red dot) PR3/2 but PQ3/2 are analytically proven in the main text

Suppose that there exists a quantum realization PQ3/2, i.e. that there exist a POVM element 0E1 and a quantum state ρ such that P(θ)=Tr(EUθρUθ) (the transpose on E is not necessary, but is used by convention to relate to the Schur product in Lemma 11). Following Lemma 11, then one has

(c2+s3)[P]=2Re(a2[P])-2Im(a3[P])=2Re(Q02+Q13)-2Im(Q03)=2Re(E02ρ02+E13ρ13)-2Im(E03ρ03)=Tr(M[E]ρ), 37

where

M[E]:=00E20-iE30000E31E02000iE03E1300.

Maximizing this over ρ yields the largest eigenvalue of M[E], see e.g. [49]. We determine this eigenvalue in Appendix E1, and the result is as follows:

Lemma 19

The quantum bound of Eq. (36) satisfies

2β2=maxE|E20|2+|E30|2+|E31|2++(|E20|2+|E30|2+|E31|2)2-4|E20|2|E31|2,

where the maximization is over all POVM elements 0E1 or, equivalently, over all orthogonal projectors E=E=E2 on C4.

Matrix entries of orthogonal projectors satisfy certain inequalities as described, for example, in [50]. There, it is shown that |E20|2+|E30|214, |E30|2+|E31|214, and thus

2β2maxx,y,z0,x+y1/4,y+z1/4x+y+z+(x+y+z)2-4xz. 38

The maximum is here over a polytope in three dimensions, and we perform the corresponding optimization in Appendix E2. We find that the maximum equals 2/3, and thus β1/3. In Appendix E2, we also provide an explicit POVM element E and quantum state ρ saturating this bound, hence β=1/3. Furthermore, since β=1/3<(c2+s3)[P]=3/5, we have shown that PR3/2 lies outside of Q3/2 and, therefore, Q3/2R3/2. See Fig. 6, where we plot P for a visual illustration of this result. This proves Theorem 7.

QJRJ for J2

In order to show that QJRJ for any J2, one can easily generalize the inequality from the previous section to the following one:

PQJ(c2J-1+s2J)[P]β=13.

See the proof in Appendix E3.

Therefore, we now want to find a spin-J correlation PJRJ such that this inequality is violated for J2, i.e., (c2J-1+s2J)[PJ]>13. For instance, an educated guess motivated by numerical results is the following trigonometric polynomial:

PJ(θ):=k=-2J2Jakeikθ,

with a-k=ak¯, a0=12, a2J=-i8, and

a2J-1-2m=316-14mm=0,,J-1,a2J-2-2l=-3i32-14ll=0,,J-2.

Indeed, this trigonometric polynomial has s2J+c2J-1=5/8>β, thus violating the quantum bound of the inequality above. Furthermore, in Appendix E4, we show that this trigonometric polynomial satisfies 0PJ(θ)1 for J7/2 and, thus, it is a valid rotation box probability distribution for J7/2 which lies outside of the quantum set. However, for values of J3, the trigonometric polynomial PJ(θ) is not a probability distribution. The way in which we deal with the remaining cases J{2,5/2,3} is to treat them on a case-by-case basis. In particular, in Appendix E4 we provide an explicit example for each case of a PJRJ which is not in QJ. In order to find these examples, we have adapted the SDP in (20) to the following one:

maxQ,Sc2J-1+s2J=2Re(a2J-1)-2Im(a2J)s.t.ak=0j,j+k2JQj,j+kfor allk,ak=-0j,j+k2JSj,j+kfor allk0,1-a0=Tr(S),Q,S0. 39

When the SDP is feasible, it returns some (2J+1)×(2J+1) matrices QS and some complex variables ak with k{0,,2J} that lead to a valid spin-J correlation (c.f. Lemma 10). Indeed, as shown in Appendix E4, the SDP for each of these cases is feasible and, moreover, its solutions are such that c2J-1+s2J>1/3, thus showing that there exist spin-J correlations that go beyond the quantum set for any J2.

QJ approximates all correlations for J

In this section, we will concern ourselves with the case of rotation boxes of unbounded spin (producing correlations which we will denote by R) and their quantum realization. We will see that in this case, we can approximate those boxes arbitrarily well with quantum boxes of finite spin J.

Elements of R are conditional probability distributions θP(+|θ), but we do not make any assumptions on the spin as in the case of RJ. However, one remaining physically motivated assumption is to demand that these outcome probabilities depend continuously on the angle θ. In fact, this is always the case in quantum theory: there, it is typically assumed that representations θUθ are strongly continuous. It is easy to convince oneself that this implies that also the probabilities P(+|θ)=Tr(UθE+Uθρ) are continuous in θ. Thus, we will define

R:={fC(SO(2))|0f(θ)1forallθ[0,2π)}.

Here, C(SO(2)) denotes the continuous real functions on SO(2), which we parametrize by the angle θ. Note that periodicity holds, f(2π)=f(0), by definition of SO(2).

We will now show that every function in R can be approximated to arbitrary precision by quantum spin-J correlations, for large enough J. We are interested in uniform approximation, i.e. if PR, we would like to find some QQJ, where J is finite (but typically large), such that P-Q:=maxθ|P(θ)-Q(θ)| is small. The following theorem makes this claim precise:

Theorem 8

The set of continuous rotational correlations R is the closure of the union of all sets of spin-J quantum boxes QJ with finite J<, i.e.

R=JQJ¯, 40

where the closure is taken with respect to the uniform norm · .

As we will explain at the end of this subsection, this statement holds in completely analogous form for more than two outcomes too, i.e. RA=JQJA¯, with the obvious definition of RA.

Note that the corresponding statement with QJ replaced by RJ is trivially true: it is well-known that every continuous function on the circle can be uniformly approximated by trigonometric polynomials [51]. However, at this point, we do not know whether all probability-valued trigonometric polynomials are contained in some QJ.

Proof

Here, we will only outline the proof idea. The technical details can be found in Appendix F. The proof can be divided into three steps. In the first step, we will use the Hilbert space L2(SO(2)) of equivalence classes of square integrable functions over the circle and construct quantum models for elements of R. To construct a quantum model for any given rotation box correlation θP(+|θ0+θ)R we find an operator P^E(L2(SO(2))) and a sequence of states {[fθ0,n]}nNL2(SO(2)) such that P(+|θ0+θ)=limnU(θ)fθ0,n|P^U(θ)fθ0,n,where U is the regular representation, acting as U(θ)f(θ)=f(θ+θ). In more detail, we define the operator P^ in the following way:

(P^ψ)(θ)=P(+|θ)ψ(θ). 41

The sequence {fθ0,n}n is given by the normalized functions that are constant in the interval [θ0-1n,θ0+1n] and 0 everywhere else. The limit of these sequences can be thought as generalized normalized eigenfunctions |θ0 of P^, and we can write θ|θ0=limnfθ,n|fθ0,n=δ(θ-θ0). It is easy to convince oneself that Uθfθ0,n=f(θ0-θ),n and hence, Uθ|θ0=|θ0-θ, and the claim P(+|θ0+θ)=limnU(θ)fθ0,n|P^U(θ)fθ0,n=θ0+θ|P^|θ0+θ follows. In total, we have seen that by making n larger and larger, the quantum box Pn(+|θ0+θ)=U(θ)fθ0,n|P^U(θ)fθ0,n more and more closely models the behavior of the rotation box P(+|θ0+θ).

In the second step, we will approximate the described quantum box Pn(+|θ0+θ) by a finite-dimensional quantum model. We will start with the same model as before, and then project it on to a finite-dimensional subspace. We recall that for the regular representation, we have a decomposition of the Hilbert space L2(SO(2))=jHj, where Hj is a one-dimensional subspace corresponding to the j-th irrep of SO(2), i.e. U(θ)|ϕj=eijθ|ϕj for every |ϕjHj. Using a basis of L2(SO(2)) that respects this decomposition, we can define the projector ΠJ=j=-JJ|ϕjϕj|. Using this projection, we can define PnJ(+|θ0+θ)=Tr(ΠJP^ΠJU(θ)ΠJ|fθ0,nfθ0,n|ΠJU(θ)), which is an element of QJ. From the Gentle Measurement Lemma [52] and Theorem 9.1. of [53], it follows that if Tr(ΠJ|fθ0,nfθ0,n|)1-ϵ then ϵPn(+|θ0+θ)-PnJ(+|θ0+θ).

In the third and final step, we show that we can make ϵ arbitrarily small by making J larger and larger. This is the case since ΠJ1 strongly for J.

The above theorem can be generalized to N-outcome boxes. We say that an N-outcome rotation box is a family of functions {Pk}k=1N such that every Pk is a non-negative and continuous function on the circle, Pk(θ):=P(ak|θ) for A={a1,,aN}, and k=1NPk(θ)=1 for every θ. For the construction of the quantum model, we use the family of operators {P^k}k=1N defined by

(P^kψ)(θ)=Pk(θ)ψ(θ), 42

and the rest of the extension to N outcomes is straightforward. For the details, see again Appendix F.

Two settings: QJ,α=RJ,α and a theory-independent randomness generator

In previous work [19], some of us have shown that the quantum and rotation sets of correlations are precisely the same for all J, when one considers just two settings (i.e. two possible rotations θ{0,α}). This equivalence is used to describe semi-device-independent protocols for randomness certification, which do not need to assume quantum theory, but instead implement some physical assumption on the response of any transmitted system to rotations.

The setup is as follows (see Fig. 3c for an illustration). The “preparation” box with settings x{1,2} is either left unchanged for x=1, or rotated by some fixed angle α>0 for x=2. The prepared system is then communicated to the “measurement” box, which outputs a{±1}. Like every semi-device-independent protocol, we have to make some assumption about the transmitted systems. Here, we assume that the spin is upper-bounded by some value J.

The statistics of the setup is described by a conditional probability P(a|θx), where θ1=0 and θ2=α. There may be other variables Λ that would admit an improved prediction of the outcome a, such that P(a|θx) is a statistical average over λ,

P(a|θx)=λΛqλP(a|θx,λ),

with some probability distribution qλ. Equivalently, we can describe the statistics with the correlations (E1,E2), where Ex=P(+1|θx)-P(-1|θx). The protocol works by showing that the observation of certain correlations (E1,E2) implies for the conditional entropy

H(A|X,Λ)H>0, 43

which essentially means that the setup produces H random bits, unpredictable even by eavesdroppers holding additional classical information λΛ.

If we assume that quantum theory holds, the set of possible correlations in this scenario is

QJ,α:={(E1,E2)|Ex=P(+1|θx)-P(-1|θx),PQJ},

where θ1=0 and θ2=α. Based on earlier work by other authors [10, 18], we have shown in [19] that this quantum set of correlations QJ,α can be exactly characterized by the inequality

121+E11+E2+1-E11-E2δ, 44

where

δ=cos(Jα)if|Jα|<π20if|Jα|π2. 45

If we do not assume quantum theory, the corresponding set of correlations is

RJ,α:={(E1,E2)|Ex=P(+1|θx)-P(-1|θx),PRJ}.

Using a lemma [54, Thm. 1.1] that constrains the derivative of trigonometric polynomials (also used here for the convex characterization of R1, see Eq. (C16)), we show that rotation box correlations must satisfy precisely the same condition as in the quantum case, i.e.

QJ,α=RJ,α. 46

Thus, for two settings and two outcomes, the possible quantum and general spin-J correlations are identical. For example, statements like “the system must be rotated by at least π/(2J) to obtain a perfectly distinguishable state” are not only true in quantum theory, but in every physical theory:

Lemma 20

Suppose that PRJ with P(+|θ0)=0 and P(+|θ1)=1. Then |θ1-θ0|π/(2J).

This equivalence, Eq. (46), allows us certify randomness independently of the validity of quantum theory. In particular, we characterize the set of “classical” correlations, i.e. for a given set of correlations, the subset containing all those that admit a description as the convex combination of deterministic correlations. This is clearly the same for both quantum and rotation cases, due to the equivalence expressed in Eq. (46). Moreover, for 0<Jα<π/2, the classical set is a strict subset of the quantum and rotation sets: CJ,αQJ,α=RJ,α. Therefore, there exist correlations (predicted by quantum theory) that are incompatible with any deterministic description, even when one allows for post-quantum strategies. Observing such correlations (E1,E2)QJ,α\CJ,α certifies a number H of random bits, as in Eq. (43), which is independent of whether quantum theory holds. That is, even an eavesdropper with arbitrary additional classical information λΛ, as well as access to post-quantum physics, could not anticipate the outputs of the device.

Accordingly, we can conceive of a random number generator whose outputs are provably random irrespective of the validity of quantum theory, with its security instead anchored in the geometry of space. This analysis is further shown to be robust under some probabilistic assumption that allows for experimental error in the spin bound.

What are classical rotation boxes?

Classical rotation box correlations are generated by a classical system with an SO(2) symmetry. For finite-dimensional systems, this entails there is a representation of SO(2) of the form given in Eq. (6) acting on the state space of the classical system. For nN, the finite-dimensional n-level classical system has a state space given by an n-simplex [31, 32]:

Δn={(p1,...,pn)|pi0,i=1npi=1}Rn, 47

and an effect space given by a n-dimensional hypercube

n={(e1,...,en)|0ei1}Rn. 48

The set of symmetries of Δn is Σ(n), which is the symmetric group on n objects. Since SO(2) is not a subgroup of Σ(n), it follows that the only representation of SO(2) which maps Δn to itself is the trivial representation. Thus the set of finite-dimensional classical systems generate the set R0 of trivial spin-0 correlations.

Infinite-dimensional classical systems can carry non-trivial actions of SO(2). Consider a system with configuration space given by the circle S1 which carries the standard action of SO(2).

The circle S1 has a topology induced by the standard topology on R2, and thus a Borel σ-algebra [51]. States of the S1 classical system are probability measures on S1, while effects are given by measureable functions f:S1R that take values between zero and one everywhere, i.e. 0f(θ)1 for all θ. We denote the space of probability measures on S1 by M1+(S1), and the space of measureable functions on S1 by M(S1).

Note that every continuous function f(S1)R is such that the preimage f-1(A) is open if A is open. Since the Borel σ-algebra is the σ-algebra generated by open sets, every fC(S1) is measurable. Since trigonometric polymomials are continuous, every trigonometric polynomial P(a|θ)M(S1).

Denoting by δθ the Dirac measure at the point θ, we have that every element in R can be generated by this infinite-dimensional classical system:

P(a|θ)=θP(a|θ)δθ. 49

We note that the standard action of SO(2) on the circle induces an action on M1+(S1), which acts on the extremal measures as:

δθδθ+θ. 50

The classical system can be thought of as ‘containing’ every spin-J system, since the subspace of M(S1) of trigonometric polynomials of degree 2J or less carries a representation k=02Jγk, where γk is the real representation of SO(2) given in Eqs. (12) and (13). Thus, there is no finite J that characterizes this classical system. Moreover, for any fixed finite value of J, this mathematical subspace cannot be interpreted as an actual standalone physical subsystem in any operationally meaningful way.

Conversely, every classical system has the property that all pure states are perfectly distinguishable. Thus, if the SO(2)-action θTθ acts non-trivially on at least one pure state ω, then ω(θ):=Tθω will be another pure state that is perfectly distinguishable from ω, no matter how small the angle θ>0. But this is incompatible with a finite value of J, as observed in Lemma 20.

The inexistence of any classical finite-spin boxes means that while any rotation box correlation P(a|θ) can be arbitrarily well approximated by a finite-dimensional quantum spin-J system, one always needs an infinite-dimensional SO(2) classical system to approximate or reproduce it, unless P(a|θ) is constant in θ for every a.

Our discussion above has focused on the paradigmatic examples of classical systems described by finite- or infinite-dimensional simplices of probability distributions, but one might instead ask more nuanced and detailed questions about the compatibility of finite spin J and different notions of classicality. For example, how about classical systems with an epistemic restriction [55]? Are systems of finite spin always contextual in the sense that they cannot be linearly embedded into any classical system [34], and if so, how crucial is the assumption of transformation-noncontextuality [56]? We leave the discussion of these interesting questions to future work.

Rotation Boxes in the Bell Scenario

In this section, we consolidate and generalize two earlier results which show how the notion of rotation boxes can be applied in the context of Bell nonlocality: assumptions on the local transformation behavior can be used to characterize the quantum Bell correlations for 2 parties with 2 measurements and 2 outcomes each [11], and they allow us to construct witnesses of Bell nonlocality for N parties [24]. Since many experimental scenarios indeed feature continuous periodic inputs, we think that these are only two examples of a potentially large class of applications of the framework.

Two parties: exact characterization of the quantum (2, 2, 2)-behaviors

One of us and co-authors have shown in [11] that the quantum (2,2,2)-correlations can be characterized exactly in terms of the local transformation behavior with respect to rotations in d-dimensional space, for every d2. Here, we give a stand-alone argument for the special case d=2.

This result contributes to the longstanding research program of characterizing the set of quantum correlations inside the larger set of correlations that satisfy the no-signalling (NS) principle, see [15] for an overview. The no-signalling principle formalizes the idea that information transfer has finite speed in order to constrain the influence between space-like separated events: one party’s choice of measurement cannot instantaneously influence the local statistics of the other. The NS principle, initially introduced in [16], was established as a foundational component of a framework in [17] where the so-called Popescu-Rohrlich correlations (or PR boxes) revealed that non-local correlations beyond those allowed by quantum mechanics are theoretically possible under the constraints of relativistic causality. That is, the set of NS correlations is known to contain the set of quantum correlations as a proper subset. However, while the NS principle has proven useful in several contexts for upper-bounding feasible correlations, characterizing the set of quantum correlations Q via simple physical principles remains an open problem [15].

Suppose that Alice holds a spin-1/2 rotation box, PQ1/2: she can choose her input by performing a spatial rotation by some angle α, and obtain one of two outcomes a{-1,+1}. Furthermore, suppose that the outcome is not only an abstract label, but has an additional geometric interpretation: Alice’s input is a spatial vector n=(cosα,sinα) (say, of a magnetic field), and her output is physically realized by giving her an answer that is either parallel (a=1) or antiparallel (a=-1) to n. Indeed, this situation is realized by a Stern-Gerlach experiment on a spin-1/2 particle in d=3 dimensions; here we restrict ourselves to d=2.

This physical intuition can be expressed as the following expectation:

  • (i)

    If outcome a is obtained on input α, then outcome -a would have been obtained on input α+π.

To make this mathematically rigorous, we have to adapt this (untestable) counterfactual claim to a (testable) statement about probabilities, namely:

  • (ii)

    P(a|α)=P(-a|α+π).

Since we can always write P(+|α)=c0+c1cosα+s1sinα, this is equivalent to the condition c0=12, and it is also equivalent to

  • (iii)

    12π02πP(a|α)dα=12 for a=+1 and a=-1.

That is, on average (over all directions), no outcome is preferred. We say that Alice’s box is unbiased [11] if one of the two (and thus both) equivalent conditions (ii) or (iii) hold. As explained above, this property follows from a geometric interpretation of Alice’s outcome as indicating that she obtains a resulting vector that is either parallel or antiparallel to her input vector.

Now consider a Bell experiment, where both Alice and Bob hold unbiased spin-1/2 boxes. Let us not assume that quantum theory holds; let us only assume that the no-signalling principle is satisfied. In this case, Alice and Bob would choose inputs α and β and obtain outputs a,b{-1,+1} such that the resulting behavior

P(a,b|α,β)

satisfies the no-signalling conditions

aP(a,b|α,β)=aP(a,b|α,β)=:PB(b|β),bP(a,b|α,β)=bP(a,b|α,β)=:PA(a|α).

Let us assume that Alice’s and Bob’s local boxes are always spin-1/2 boxes, and are always unbiased, regardless of what the other party measures. That is, consider the situation in which Bob decides to input angle β into his box, and obtains outcome b, and subsequently communicates this choice and outcome to Alice (say, over the telephone). In this case, Alice would update her probability assignment to

Pb,βA(a|α):=P(a,b|α,β)PB(b|β),

where PB(b|β) is the probability for Bob to obtain outcome b. We will assume that this “conditional box” still produces an unbiased spin-1/2 correlation, for all values of β and b, and we make the analogous assumption if the roles of Alice and Bob are interchanged.

Note that we are not making any assumptions about the global correlations (or their transformation behavior) directly, except that we demand no-signalling.

Surprisingly, the conditions above enforce that the global correlations are quantum (see Appendix G1 for the proof):

Theorem 9

Under the assumptions above, the behavior P is a quantum behavior. That is, there exists a quantum state ρAB on the two-qubit Hilbert space AB and a positive map τ on B with τ(1B)=1B such that

P(a,b|α,β)=TrρABe-iαZ|aa|eiαZτ(e-iβZ|bb|eiβZ),

where Z=12100-1 is half of the Pauli-Z matrix, and |±1=12(|0±|1).

We do not currently know whether the unitary rotation by angle β can be pulled out of the map τ, or whether this positive, but not necessarily completely positive, map can perhaps be dropped completely. This map τ is, however, necessary in the analogous statement for dimension d=3: it is well-known that the quantum singlet state of two spin-1/2 particles leads to perfect anticorrelation between Alice’s and Bob’s binary outcomes [57], but that there is no quantum state that would lead to perfect correlation. Formally, perfect correlation can be obtained by taking the partial transpose of one half of the singlet state, and considering the resulting action on Bob’s local measurement (while leaving the singlet state intact) can be interpreted as a reflection of Bob’s description of spatial geometry relative to Alice’s.

Note that P will be a quantum correlation even if a non-completely positive map τ is necessary: this map cannot be physically implemented, but Bob can still use it to calculate the set of POVM elements that he should use to measure. This way, Alice and Bob can make sure to generate correlations according to P(a,b|α,β).

If Alice and Bob restrict themselves to input one of two angles each, α0,α1 or β0,β1, they generate an instance of what has been called the quantum (2, 2, 2)-behaviors (2 parties, 2 settings and 2 outcomes each):

P(a,b|x,y):=P(a,b|αx,βy)(x,y{0,1}).

The above theorem shows that if Alice’s and Bob’s local conditional boxes are spin-1/2 boxes and unbiased, then P(ab|xy) will be a quantum (2, 2, 2)-behavior. In this case, the mere possibility that Alice and Bob could have input other angles, and that the outcome probabilities would have had to depend linearly on the resulting two-dimensional vectors, constrains these correlations to be quantum.

The results of [11], however, show more: all quantum (2, 2, 2)-behaviors can be obtained in this way, if supplemented with shared randomness:

Theorem 10

The set of quantum (2, 2, 2)-behaviors is exactly the set of non-signalling behaviors that can be obtained in Bell experiments from ensembles of nonlocal boxes that are locally unbiased and locally spin-1/2.

That is, regardless of which theory holds, the resulting behaviors will be quantum. Moreover, all such quantum behaviors can be realized in some theory, namely quantum theory, via random choices among boxes that are locally spin-1/2 and unbiased.

The proof is based on the well-known fact that all extremal quantum (2, 2, 2)-behaviors can be generated on two qubits (and, locally, on the equatorial plane of these qubits, i.e. on two rebits) [12, 25, 58, 59]. To obtain all non-extremal quantum (2, 2, 2)-behaviors, Alice and Bob need additional shared randomness that allows them to select at random between one of several such boxes. See [11] for an explanation of why shared randomness cannot be avoided.

To see that local unbiasedness cannot be removed as a premise of the theorems above, consider the following example. Suppose that Alice and Bob hold local spin-1/2 boxes SA,SBR1/2{0,1}, satisfying

QA(1|α)=12+12cosα,QB(1|β)=12+12cosβ.

What they do is the following. Alice and Bob input their angles into their local boxes, and feed their respective outcomes x,y{0,1} into a PR box

PPR(a,b|x,y)=12δ(1-ab)/2,xy(a,b{-1,+1}).

That is, if the inputs to the PR box are x=y=1, they obtain perfectly anticorrelated outputs, and otherwise, perfectly correlated ones. The result of this procedure defines their non-signalling behavior P. It is not difficult to see that PB(b|β)=12 for all b and all β, and hence

Pb,βA(a|α)=2P(a,b|α,β)=2c,d=01PPR(a,b|c,d)QA(c|α)QB(d|β) 51

is a trigonometric polynomial of degree 1 in α, for every fixed b, β, and a. Similar reasoning applies to Pa,αB(b|β). Hence, all local conditional boxes are spin-1/2 boxes. Set α0=β0:=π and α1=β1:=0, then QA(c|αx)=δcx and QB(d|βy)=δdy, and so

P(a,b|αx,βy)=12c,d=01δ(1-ab)/2,cdQA(c|αx)QB(d|βy)=12δ(1-ab)/2,xy=PPR(a,b|x,y).

Since P can reproduce the PR box correlations for two fixed angles, it is not a quantum behavior. And this is consistent with the theorems above because P is not locally unbiased. To see this, use Eq. (51) and find, for example,

P-1,βA(+1|α)=12+12cosα12+12cosβ.

Treating this as a trigonometric polynomial in α, the coefficient c0 equals 141+cosβ, which is not for all β equal to 12. That is, P is not locally unbiased.

Many parties: witnessing Bell nonlocality

Our framework also helps to clarify and generalize the results of Nagata et al. [24]. In this paper, the authors offer an additional constraint on local realistic models of physical phenomena, which they refer to as rotational invariance, but we shall call spin direction linearity (reasons for which will become clear). This allows for indirect witnesses of Bell nonlocality, for correlations that would otherwise have a local hidden variable description.

They consider an N-party Bell-type scenario, in which every party holds a spin-12 particle. Each party measures the spin component in a chosen direction nj, and outputs a local result rj(nj){±1}. The “Bell” correlation function is introduced as the average of the product of all local results: E(n1,,nN)=r1(n1)rN(nN)avg. Their additional assumption (spin direction linearity) enforces the following structure for any such correlations:

E(n1,,nN)=T^·(n1nN),

where T^ is the correlation tensor Ti1,,iNE(x1(i1),,xN(iN)), where xj(ij),ij{1,2,3} are unit directional vectors of the local coordinate system of the jth party. This is to say that the correlation function is linearly dependent on the unit directions nj along which the spin component is measured, i.e.

E(n1,,nN)=Ti1,,iNni1niN,

with summation over repeated indices.

The three assumptions allow the authors to derive a more restrictive Bell-type inequality, namely:

πNi1,,iN=1,2Ti1,,iN24NTmax,

where Tmax is the maximal possible value of the correlation tensor component, i.e.

Tmax=maxn1,,nNE(n1,,nN).

This would be evaluated by measuring the components Ti1,,iN that compose T^, and then using the tensor to determine the maximum value of E(n1,,nN). Their inequality being strictly less general than Bell’s theorem then allows for the certification of “non-classical” phenomena by observing correlations that would otherwise not violate any Bell inequality. In such an instance, non-classicality is to say that the assumptions of locality, realism and spin direction linearity cannot jointly hold. In particular, the authors of [24] give an example of correlations T that admit a local hidden variable model, but that do not admit such a model if one assumes in addition spin direction linearity.

Although their result is formulated for SO(3), with spin directions defined by vectors nj in three dimensions, the authors use the reparameterization nj(αj)=cos(αj)xj(1)+sin(αj)xj(2), for the plane defined by xj(1),xj(2), such that their main result is stated in terms of just one parameter αj per party. Accordingly, the results hold equally for rotations constrained to a 2D-plane, i.e. SO(2) rather than SO(3). It follows that our framework may be relevant to understand or generalize their results.

In particular, spin direction linearity is not actually about rotational invariance, as is claimed in their paper, but rather captures the assumption that the local systems are spin-12 particles. (Moreover, we will claim that one need only assume that the local systems can be described by a spin-12 box.) The states of a single spin-12 system (a qubit) can be represented by unit vectors on the Bloch ball:

ρ=12(1+n·σ),

which, by measuring in the basis as defined by the jth observer, are mapped via unitary transformations Uθ to states

ρ=12(1+(Rθ·n)·σ).

Local probabilities are linear in states, so are affine-linear in spin direction nj=Rθ·n. The local, conditional boxes P~(rj|nj) (an N-party extension of the conditional boxes introduced in Sect. 5.1) can be written as

P(r1,,rN|n1,,nN)P(r1,,rj-1,rj+1,,rN|n1,,nj-1,nj+1,,nN),

so probabilities P(r1,,rN|n1,,nN) will be affine-linear in spin directions nj, for all 1jN. The constant drops out when going from probabilities to correlations, so then we get spin direction linearity when all subsystems are spin 12.

So far, this demonstrates that the systems being spin-12 is a sufficient condition for E(n1,,nN) to be linear in spin directions. This can also be seen in our framework, by noting that the local systems being spin-12 means that the local conditional boxes P~(rj|nj) should be in R1/2; i.e. they are trigonometric polynomials in αj of degree 1 at most. On the other hand, if the local systems are not spin-12, then the probabilities may contain sin(kαj) or cos(kαj) terms (for k2), in which case spin direction linearity is violated. As such, we can note that the systems being spin-12 is also a necessary condition for spin direction linearity. This is to say, the main result of [24] can be clarified using our framework as an inequality derived from locality, realism and the assumption that the local systems can be characterized as spin-12 boxes. Notably, this reformulation does not rely on the validity of quantum theory (the systems do not need to be quantum spin-12 particles, as in their paper); all three assumptions are theory-independent.

Connection to Other Topics

Almost quantum correlations

As discussed in Sect. 3.3, the set of rotation box correlations bears close resemblance to the set of almost quantum correlations [25]. Indeed, any PRJ can be generated as follows:

P(+|θ)=ψ|UθE+Uθ|ψ, 52

where |ψC2J+1 and E+ is positive semidefinite but not necessarily a POVM element. The only requirement is that E+ gives valid probabilities on the states of interest, i.e. on the states Uθ|ψ for all θ.

This is analogous to almost quantum correlations which are a relaxation of the Bell correlations generated by quantum systems. In standard quantum theory, local separation of the measurement parties (and therefore the no-signalling condition) is implemented by assigning commuting subalgebras to them. For example, consider the case where we have two observers Alice and Bob. We denote Alice’s subalgebra by AC and Bob’s subalgebra by BC, where C can be thought of as a larger global algebra. Here, the commutativity of A and B means that every AA commutes with every BB, i.e. [A,B]=0. When we describe the measurements of Alice and Bob, we equip them with collections of PMs (projective measurements) {Ea|xA}a,xA and {Eb|yB}b,yB respectively, where for every input x, the set {Ea|xA}a is a valid PM (and similarly for Bob). Then, the correlations between Alice and Bob are given by P(a,b|x,y)=ψ|Ea|xAEb|yB|ψ. For “almost quantum” correlations, the assumption that Alice’s and Bob’s collections of PMs are subsets of two commuting subalgebras is relaxed. That is, not all elements of Alice’s collection of PMs have to commute with all elements of Bob’s PM collections, but it is only assumed that they commute on the state of interest for a given setup. In other words, if a given preparation is described by the state |ψ, it is assumed that [Ea|xA,Eb|yB]|ψ=0 for all inputs x and y and outputs a and b. Furthermore, the correlations are still computed by the Born rule p(a,b|x,y)=ψ|Ea|xAEb|yB|ψ. We note that the product Ea|xAEb|yB cannot be considered a bipartite local effect by itself, but only obtains its meaning by combining it with the state |ψ describing the physical situation. This resembles the situation for the rotation boxes, where E+ by itself is not a POVM element, and only the combination of E+ with the states {Uθ|ψ}θ has a physical meaning.

Furthermore, a notable feature both relaxation sets share is that they admit a characterization in terms of semidefinite constraints (as we have seen in 3.3), which allows us to efficiently solve optimization problems within their set by means of SDP in order to bound quantum solutions [60]. This is in contrast to the quantum sets (of Bell resp. spin correlations) which are not known to have characterizations in terms of SDPs.

Orbitopes and spectrahedra

In this section, we show that the state spaces of the spin-J rotation box systems ΩJ are isomorphic to universal Carathéodory orbitopes. Moreover, we show they are isomorphic to spectrahedra. A spectrahedron is the intersection of an affine space with the cone of positive-semidefinite matrices.

Given a list of integers A=(a1,...,an)Nn, the Carathéodory orbitope CA [26] is defined as the convex hull of the following SO(2) orbit in R2n:

{(cos(a1θ),sin(a1θ),...,cos(anθ),sin(anθ))|θ[0,2π)}. 53

The orbitope C(1,...,d) is known as the universal Carathéodory orbitope Cd, and is affinely isomorphic to the state space ΩJ=d2 of the spin-J rotation box system. Similarly C^do, the co-orbitope cone dual to C(1,...,d) is the set of non-negative trigonometric polynomials and is isomorphic to the cone generated by the effect space EJ.

Explicitly, C^do is given by:

{(c0,c1,s1,...,cd,sd)R2d+1|c0+k=1dckcos(kθ)+sksin(kθ)0}. 54

We can characterize Cd in terms of C^do as follows: a point (a1,b1,...,ad,bd)R2d is in the universal Carathéodory orbitope Cd if and only if

c0+k=1dckak+skbk0,(c0,c1,s1,...,cd,sd)C^do. 55

By Theorem 5.2 of [26], the universal Carathéodory orbitope Cd (and therefore ΩJ=d2) is isomorphic to the following spectahedron:

1x1xd-1xdy11xd-2xd-1yd-1yd-21x1ydyd-1y11,, 56

where

xj=aj+ibj, 57
yj=aj-ibj, 58

and (a1,b1,...,ad,bd)R2d is a point in the orbitope Cd. The extremal points occur for ak=cos(kθ) and bk=sin(kθ), thus the orbitope Cd is the convex hull of:

1eiθei(d-1)θeidθe-iθ1ei(d-2)θei(d-1)θe-i(d-1)θe-i(d-2)θ1eiθe-idθe-i(d-1)θe-iθ1. 59

Let us note that this statement is equivalent to Theorem 4. Consider the orbit Uθ|ψψ|Uθ for |ψ and Uθ as defined in Theorem 4:

12J+11eiθei(2J-1)ei2Je-iθ1ei(2J-2)ei(2J-1)e-i(2J-1)e-i(2J-2)1eiθe-i2Je-i(2J-1)e-iθ1. 60

This orbit is isomorphic to the orbit of Eq. (59) for d=2J. According to Theorem 4, every spin-J correlation PRJ can be written

P(+|θ)=Tr(E+Uθ|ψψ|Uθ),

i.e. is a linear functional that takes values in [0, 1] on this orbitope; and, conversely, every such functional is an element of RJ. Therefore, we may say that ΩJ=d2, the state space of the spin-J GPT system RJ, is an orbitope, and moreover, it can be interpreted, due to Theorem 4, as a subset of the quantum state space.

Symmetric entanglement witnesses for rebits

Consider the following orbit of qubit states |ψ(θ)ψ(θ)| in D(C2), where

|ψ(θ)=U(θ)|+=12(eiθ2|0+e-iθ2|1), 61

with

U(θ)=eiθ200e-iθ2,|±=|0±|12.

By writing the orbit in the {|+,|-}-basis

|ψ(θ)=cosθ2|++sinθ2|-, 62

we see that it corresponds to the pure states of a rebit (a qubit in quantum theory over the real numbers R), acted on by a real projective representation of SO(2). The orbit |ψ(θ)ψ(θ)| can thus be viewed as an orbit of rebit states in LS(R2), the symmetric linear operators on R2, or alternatively as an orbit of symmetric qubit states in LSH(C2)LH(C2), where LSH(C2) are the symmetric Hermitian operators, in this case with respect to the |± basis.

Given d rebits with pure states corresponding to rays in (R2)d, the pure symmetric states are those lying in Symd(R2), the symmetric subspace of (R2)d. The set of pure symmetric product states is the set of |ψd, where |ψ is an arbitrary rebit state, and they span the space Symd(R2). The mixed symmetric states are given by the positive unit-trace operators in LS(Symd(R2))LSH(Symd(C2)). This isomorphism follows from the fact that Symd(C2) is the complexification of Symd(R2) and that LS(Rd)LSH(RdC), as shown in Lemmas 30 and 29.

Now consider the orbit of a symmetric two-rebit pure state |ψ(θ)2, where |ψ(θ) defined in Eq. (61). Explicitly, |ψ(θ)ψ(θ)|2LSH(Symd(C2))Sym2(C2)Sym2(C2) is

|ψ(θ)ψ(θ)|2=141eiθeiθei2θe-iθ11eiθe-iθ11eiθe-2iθe-iθe-iθ1. 63

Compare this to the orbit Uθ|ψψ|UθLH(C3) defined in Eq. (60) for J=1, where

Uθ|ψ=13(e-iθ|-1+|0+eiθ|1), 64

and

Uθ|ψψ|Uθ=131eiθe2iθe-iθ1eiθe-2iθe-iθ1. 65

There exists an invertible linear map that maps |ψ(θ)ψ(θ)|2 to Uθ|ψψ|Uθ which can be constructed as follows:

L|ψ(θ)ψ(θ)|2L=Uθ|ψψ|Uθ, 66
L=4310000121200001. 67

The inverse of this map is given by

MUθ|ψψ|UθM=|ψ(θ)ψ(θ)|2, 68
M=34100010010001. 69

This shows that the convex hulls of the two orbits are isomorphic as convex sets. This entails that the space of linear functionals that map every element |ψ(θ)ψ(θ)|2 into the interval [0, 1] is isomorphic to R1. Thus, for every PR1, there exists a linear operator WLSH(Sym2(C2)) and therefore also in LS(Sym2(R2)) such that

P(+|θ)=Tr(W|ψ(θ)ψ(θ)|2). 70

The set of linear operators W such that Tr(W|ψ(θ)ψ(θ)|2)0 for all θ are two-rebit symmetric entanglement witnesses. Thus, the cone generated by R1, is isomorphic to the cone of two-rebit symmetric entanglement witnesses.

The following theorem generalizes the above observation to arbitrary J:

Theorem 11

Every PRJ can be realized as

P(+|θ)=Tr(|ψ(θ)ψ(θ)|2JE+), 71

with E+ an operator in LS(Sym2J(R2)), the symmetric operators on the symmetric subspace of 2J rebits, such that Tr(|ψ(θ)ψ(θ)|2JE+)[0,1].

This theorem is proven in Appendix H2.

The possible operators E+ include positive operators in LS(Sym2J(R2)), which correspond to standard POVM elements on 2J rebits. However, the possible operators E+ also include non-positive operators such as rebit symmetric entanglement witnesses. A d-rebit symmetric entanglement witness WLS(Symd(R2)) is an operator defined as:

ψ|dW|ψd0forallψR2. 72

In typical applications of entanglement witnesses, it is assumed that there exists at least one state ρ such that Tr(ρW)<0, which must then be entangled. Here, however, we are using the notion of an entanglement witness in the generalized sense, such that it also includes W that are non-negative on all symmetric states. Thus, we obtain the following corollary:

Corollary 1

The cone generated by RJ is isomorphic to the set of 2J-rebit symmetric entanglement witnesses.

The fact that Q1=R1, but Q3/2R3/2 can thus be interpreted as follows: all correlations (in θ) generated by two-rebit symmetric entanglement witnesses can also be generated by proper two-rebit measurement operators (and similarly for zero or one rebits, because Q0=R0 and Q1/2=R1/2). However, the analogous statement for three rebits is false.

There is a compelling analogy of this behavior to the study of Bell correlations: all non-signalling correlations on pairs of quantum systems are realizable within quantum theory [61], but this is not true for all non-signalling correlations on triples of quantum systems [62]. The proof of this uses the fact that non-signalling correlations of quantum systems can always be generated by entanglement witnesses, regarded as a generalization of the notion of quantum states, which is yet another similarity to our result above.

Conclusions and Outlook

In this paper, we have introduced a notion of “rotation boxes”, describing all possible ways in which measurement outcome probabilities could respond to spatial rotations around a fixed axis, in any covariant physical theory. We have thoroughly analyzed the resulting notion of spin-bounded correlations, and have demonstrated a variety of interesting results and applications. First, for the prepare-and-measure scenario, we have shown that, for spin J{0,1/2} systems, quantum theory predicts the same observable correlations as the most general physics consistent with the SO(2)-symmetry of the setup. For scenarios with two outcomes, the same is also true for the spin-1 case, although it remains an open questions as to whether this generalizes to any number of outcomes.

However, for spin J3/2, we have demonstrated a gap between quantum and more general predictions; we have derived a Tsirelson-type inequality and constructed an explicit counterexample consistent with general rotation boxes, but inconsistent with quantum rotation boxes. Moreover, we have presented a family of GPT systems that generate these “post-quantum” correlations. On the one hand, this result could hint at possible probabilistic phenomena consistent with spacetime geometry that, if indeed observed, would not be consistent with quantum theory. On the other hand, it is conceivable that the gap closes when we consider the full Lorentz or Poincaré group, which would thus reproduce crucial predictions of quantum theory from spacetime principles alone. For J, we have shown that every continuous rotational correlation can be approximated arbitrarily well by finite-J quantum systems.

Given the theoretical gap between quantum and more general rotational correlations, we have presented a metrological game in which general spin-3/2 resources outperform all quantum ones, demonstrating a post-quantum advantage. We have further applied our framework to Bell scenarios, building on previous results. First, we have demonstrated why the “local unbiasedness” assumption introduced in [11] is crucial to recover the (2, 2, 2)-quantum Bell correlations from the no-signalling set, and that it has a geometric interpretation relating the outputs to the inputs of the box. Second, we have clarified the “rotational invariance” assumption used in [24], from which the authors derive indirect witnesses of multipartite Bell nonlocality. In particular, we argued that their assumption actually expresses the statement that all local subsystems are spin-1/2 (quantum or otherwise), and therefore that is does not rely on the validity of quantum theory.

In addition to addressing foundational questions, our work offers several interesting applications to explore in future work, such as the semi-device-independent analysis of experimental data. For instance, recent experiments have successfully probed Bell nonlocality in many-body systems like Bose–Einstein condensates, using so-called Bell correlation witnesses [63]. These witnesses have the advantage of being experimentally accessible by treating the Bose–Einstein condensate as a single party in which collective observables can be measured. However, a disadvantage of this approach is that it requires additional assumptions compared to a typical Bell test, namely the validity of spin-algebra in quantum mechanics and trust in the measurements, making it device-dependent. Our framework is a suitable candidate for providing weaker assumptions for carrying out semi-device-independent analysis of the observed experimental data, in particular in situations where the experimental parameters are spatiotemporal in nature.

Another interesting application would be to devise self-testing-inspired protocols via rotations. Typical self-testing [6, 64] protocols are tailored to specific pairs of states and measurements, but do not tell us how to operationally implement other valid measurements on the state. It would be interesting to explore whether semi-device-independent self-testing-inspired protocols can be devised where the inputs correspond to directions in physical space (on which the rotation group acts), and the outputs are angular-momentum-valued physical quantities (instead of abstract labels), in order to not only certify a certain state and the implemented measurements, but also certify the state with all other valid measurements in different directions.

A further direction to explore would be whether one can carry a similar study than the one in this manuscript by replacing the local spin bound by a local energy bound (for instance, making use of the Mandelstam-Tamm quantum speed limit [20, 22]). The settings would then not correspond to two different directions in space, but to two different time intervals according to which we let the systems evolve locally. Formally, this would replace the group of rotations SO(2) of this paper by the time translation group (R,+). More generally, it will be a natural next step to consider other groups of interest, such as the full rotation group SO(3) or the Lorentz group, and to see which novel statistical phenomena arise from the non-commutativity and other strutural properties of these groups.

Furthermore, the interplay of entanglement and nonlocality with the group theoretic structure deserves more study. The paradigmatic example is that of spin-1/2 fermions obtaining a (-1) phase on (2π)-rotations, visible in the presence of initial entanglement. This already demonstrates one surprising insight, potentially amongst others still waiting to be discovered, at the intersection of probabilistic and spacetime structure.

Appendices

Appendix A. Background Material

1. Finite-dimensional projective representations of SO(2)

Theorem 16.47 of [65] states that given a compact group G with universal cover G~, a covering map Φ:G~G, and a finite-dimensional projective unitary representation Π:GPU(H), there is a unitary representation Σ:G~U(H) such that ΠΦ=QΣ, where Q is the quotient homomorphism Q:U(H)PU(H), Q:UU/{eiθ} for θR. Any such Σ is irreducible if and only if Π is irreducible.

If G=SO(2), then G~=(R,+). The irreducible unitary representations RU(1) are given by xeitx with tR. These are projective representations of SO(2) and are projectively equivalent to the trivial representation x1. Thus the only irreducible projective representation of SO(2) is the trivial representation. Equivalently, unitary projective irreducible representations are maps SO(2)PU(1), and PU(1) is just the trivial group.

We now characterize reducible projective representations of SO(2).

Lemma 21

Any finite-dimensional projective representation of SO(2) can be written in the form of Eq. (4):

Uθ=j=-JJ1njeijθ, A1

where J{0,12,1,...} and njN0.

Proof

A generic representation RU(H) is of the from

xeidiag(j1,...,jn)x(jiR) A2

in some basis, where there can be repeated entries and, without loss of generality, ikjijk.

The requirement that it is a projective representation of SO(2) entails that

eidiag(j1,...,jn)2π=eiϕ, A3

for some ϕR, which entails

2πji+2πqi=ϕforsomeqiZ, A4
ji+qi=ϕ2π. A5

Thus, ji-jk=qi-qk, and the difference ji-jk is integer-valued for all ij.

Setting j1=ϕ0 and ji=j1+ki with kiN0, the projective representation is of the form:

eiϕ0eidiag(0,k2,...,kn), A6

and can be characterized by a list of non-negative integers {k2,...,kn}. We are however interested in special unitary representations and can transform as follows:

eiϕ0eidiag(0,k2,...,kn)ei(ϕ0+kn2)eidiag(-kn2,k2-kn2,...kn2). A7

Thus, every projective unitary representation can be characterized by a list of integers or half-integers {k1,...,kn}={-kn2,k2-kn2,...kn2}, where k1=-kn.

This lemma entails that any projective representation of SO(2) is characterized by a list {j1,...,jn} of integers or half-integers.

Lemma 22

Projective representations of SO(2) of the form {-J,-J+1,...,J-1,J} with JN are also representations of SO(2), while those with JN/2 are purely projective representations.

Proof

This is because eidiag(-J,...,J)2π equals 1 for integer J and -1 for half-integer J.

2. Real projective representations of SO(2)

Real irreducible representations of SO(2) are labelled by non-negative integers kN0 and are given by the trivial representation for k=0 and by

cos(kθ)-sin(kθ)sin(kθ)cos(kθ) A8

for kN. Thus, a real representation of SO(2) is labelled by a list of non-negative integers {k1,...,kn}. We note that for k a half-integer, Eq. (A8) defines a real irreducible projective representation of SO(2).

Lemma 23

The complexification of the real irreducible projective representation {k} of SO(2) with k0 integer or half-integer is the complex reducible protective representation {k,-k}.

Proof

The real matrix

cos(kθ)-sin(kθ)sin(kθ)cos(kθ), A9

acting on C2 can be diagonalized:

cos(kθ)-sin(kθ)sin(kθ)cos(kθ)eikθ00e-ikθ.

Our general framework of rotation boxes implies that we have real representations of SO(2), because the space of ensembles of boxes (the vector space carrying the GPT system which represents it) will always be a vector space over R. This is also true for projective representations in quantum theory, where SO(2) acts on the vector space of Hermitian matrices that contains the density matrices. However, the following lemma will be useful when discussing quantum theory over the real numbers R:

Lemma 24

Representations of SO(2) {-J,-J+1,...,J-1,J} with integer J are also real representations of SO(2) {0,...,J}, while projective representations SO(2) {-J,-J+1,...,J-1,J} with half-integer J are real projective representations {12,...,J}.

Proof

Consider the following change of basis:

eikθ00e-ikθcos(kθ)-sin(kθ)sin(kθ)cos(kθ). A10

Thus, for integer J:

eidiag(-J,...,J)j=0Jcos(kθ)-sin(kθ)sin(kθ)cos(kθ), A11

which is a real representation of SO(2).

For half-integer J:

eidiag(-J,...,J)j=12Jcos(kθ)-sin(kθ)sin(kθ)cos(kθ), A12

which is a real projective representation of SO(2).

3. Representation-theoretic background

We introduce some necessary representation-theoretic concepts before proceeding with the proofs. Here vector spaces V are isomorphic to Cn, unless otherwise stated. A representation of G is a homomorphism ρ:GGL(V) with the general linear group GL(V) the group of automorphisms on V. We note that we do not require the representation to be faithful (i.e the map is not required to be injective) since we are interested in finite-dimensional unitary representations of (R,+), which is the universal cover of SO(2). The vector space V is the carrier space or representation space of ρ; however, we sometimes call it the representation.

When two representations ρ:GGL(V) and σ:GGL(W) are isomorphic, we write ρσ, or, when the context is clear, VW. An isomorphism of representations is given by an invertible linear map L:VW which is equivariant: σ(g)L(v)=L(ρ(g)v).

Given a representation ρ:GGL(V), we denote by ρ¯:GGL(V¯) the complex conjugate representation and by ρ:GGL(V) the dual representation. For finite-dimensional representations, we have ρ¯ρ.

We denote the space of linear maps from V to W by L(V,W). It carries a representation τ:GGL(L(V,W)) given by (τ(g)(M))(v)=σ(g)M(ρ(g-1)v).

Given a complex vector space V, restricting scalar multiplication from C to R defines the real vector space VR, known as the realification of V, where dimR(VR)=2dimC(V). Given a representation ρ:GGL(V), the space VR carries a real representation ρR:GGL(VR,R) [66].

Given a real vector space W with basis {ei}i, it can be complexified to obtain WC=CRW with basis {1Rei}i. Given a real representation ρ:GGL(W,R), the complexification of the representation ρ is ρC:GGL(WC,C) defined as ρC(g)(1ei)=1ρ(g)(ei) [66].

Definition 4

(Real structure). Given a complex vector space V, a real structure j is an antilinear map j:VV which is an involution: jj=1V. If V carries a representation ρ:GGL(V), then the representation ρ carries the real structure j if j is equivariant: ρ(g)j(v)=j(ρ(g)v).

Given a complex vector space V with a real structure j, an arbitrary vV can be expressed as v=vj=+1+vj=-1 where vj=+1=v+j(v)2 and vj=-1=v-j(v)2. Hence the realification VR decomposes into the direct sum VRVj=1RVj=-1 where Vj=±1:={vV|j(v)=±v}.

Equivariance of j implies that the real subspaces Vj=±1 are closed under the action of ρ(g), and hence ρR decomposes into the direct sum of real representations ρj=+1Rρj=-1, where ρj=+1Rρj=-1 [67, p.95].

Lemma 25

Given a representation ρ:GGL(V) with real structure j, we have (ρj=+1)Cρ.

Proof

Given a complex representation ρ with real structure j, we define the map Π+:(ρ,j)ρj=+1, where ρj=+1 is the real representation defined above.

Given a real representation σ, we define the map Π:σ(σC,k), where σC is the complexification of σ and the real structure k is defined as k(zv)=z¯v.

From [67, p.94], it follows that ΠΠ+ is the identity morphism, which implies

(ρ,j)ΠΠ+(ρ,j)((ρj=1)C,k). A13

Defining the map Γ:(ρ,j)ρ, the claim eRC=r+e+ of [67, Proposition (6.1)] can be expressed in our notation as

ΓΠ(ρ)=ρC. A14

Combining the above two equivalences gives

ρΓ(ρ,j)ΓΠΠ+(ρ,j)Γ((ρj=1)C,k)(ρj=1)C, A15

which proves the lemma.

Lemma 26

Given a representation ρ:GGL(V), the real subspace Sym(VV¯)VV¯ carries the real representation ρ:GGL(Sym(VV¯)), whose complexification is isomorphic to ρ(g)ρ¯(g).

Proof

Given a linear space V and its complex conjugate space V¯, where V¯ has the same elements, but scalar multiplication given by αv=α¯v, we can define the tensor product space WVV¯, where scalar multiplication is defined as

αW(vw)=αvw=vαw=vα¯w. A16

This space carries a representation ρ(g)ρ¯(g). Consider the swap map S:WW, vwwv. This map is anti-linear since

S(αW(vw))=S(αvw)=wαv A17
=wα¯v=α¯WS(vw). A18

Moreover, S is equivariant:

S(ρ(g)ρ¯(g)(vw))=ρ(g)ρ¯(g)S(vw). A19

The existence of an equivariant anti-linear map S:VV¯VV¯ entails that VV¯ has a real structure given by S. The +1 eigenspace of S is Sym(VV¯):={wVV¯|S(w)=w}. By Lemma 25, Sym(VV¯) carries a real representation ρ:GGL(Sym(VV¯)), whose complexification is ρ.

Lemma 27

The real subspace of Hermitian operators on H, LH(H)L(H), carries a real representation of G.

Proof

The Hermitian adjoint of a map ML(H,H) is the map ML(H,H) defined by Mv,wH=v,MwH. The resulting map (“adjoint map”) :L(H,H)L(H,H), MM is anti-linear. Moreover, it is equivariant:

v,(σ(g)Mρ(g-1))wH=σ(g)Mρ(g-1)v,wH=Mρ(g-1)v,σ(g-1)wH=ρ(g-1)v,Mσ(g-1)wH=v,ρ(g)Mσ(g-1)wH.

Thus, the +1 eigenspace LH(H):={M=M|ML(H,H)} carries a real representation of G.

Lemma 28

LH(H)Sym(HH¯) as real representations.

Proof

We define an equivariant invertible linear map Sym(HH¯)LH(H). First, we define the map

L:HH¯L(H,H)HH, A20
eiejeiej A21

which is a group representation isomorphism HH¯HH. We now show it maps Sym(HH¯) to LH(H) and hence is an isomorphism of real representations:

L(S(eiej))=L(ejei)=ejei=(eiej)=(L(eiej)).

Hence L(S(·))=L(·), which implies that for all wSym(HH¯),we have L(w)=L(w). Conversely, for all wHH such that w=w, we have L-1(w)Sym(HH¯).

4. Relevant vector space isomorphisms

Lemma 29

Given a real Hilbert space H and its complexification H, the space LS(H) of symmetric operators on H is isomorphic to the space LSH(H) of symmetric Hermitian operators on H.

Proof

HRn and its complexification HCn. Fixing a basis, an operator OLH(H) is symmetric if and only if its entries are real-valued. Thus the symmetric Hermitian operators on H are given by the n×n real symmetric matrices and therefore isomorphic to LS(H).

Lemma 30

The space Symd(C2) is the complexification of Symd(R2).

Proof

Consider a basis {|0,|1} for both H{R2,C2}. The symmetric group Σd acts on Hd by permuting the tensor factors. A basis for Symd(H) is {|i}0d, with

|i=x{0,1}d|H(x)=i|x, A22

where H(x) is the Hamming weight of the bit string x{0,1}d. Thus, there is a common basis {|i}0d for Symd(R2) and Symd(C2), showing that Symd(C2) is the complexification of Symd(R2).

Corollary 2

The real vector space of symmetric operators LS(Symd(R2)) is isomorphic to the real vector space of symmetric Hermitian operators LSH(Symd(C2)).

5. Relevant SO(2) group representation isomorphisms

In the following, C2 carries the SO(2) projective representation {-12,12}, and R2 carries the real projective representation {12}. WV indicates that the representation of SO(2) on V is isomorphic to the representation of SO(2) on W. We note that the representation {-12,12} is isomorphic to its conjugate and thus to its dual and hence is known as self-dual.

Lemma 31

Symd(C2) carries the representation {-d2,...,d2} of SO(2).

Proof

A basis for Symd(C2) is {|k}k=0d, where

|k=x{0,1}d|H(x)=k|x. A23

The action of diag(eiθ2,e-iθ2)d on this basis is:

|k=x{0,1}d|H(x)=k|xx{0,1}d|H(x)=keik2|x. A24

Thus, Symd(C2) carries the representation {-d2,...,d2}.

Corollary 3

The representation Symd(C2) is self-dual.

Lemma 32

Sym2d(R2) carries the real representation {0,1,...,d} of SO(2).

Proof

The space R2 carries a real irreducible projective representation {12}. The complexification C2 carries a complex irreducible projective representation {-12,12}. Symd(C2) carries the complex projective representation {-d2,-d2+1,...,d2} and Sym2d(C2) carries the complex representation {-d,-d+1,...,d-1,d}. Thus, Sym2d(R2) carries a real representation {0,...,d}.

Appendix B. Proofs for Section 3

1. Proof of Theorem 1

The following lemma has been established in a different context by Miguel Navascués (unpublished). The conditions (ii) and (iii) in this lemma are a priori inequivalent if G has degenerate spectrum, and the distinction of these two cases will become useful in the proof of Lemma 34.

Lemma 33

(Miguel Navascués [68]). Let G=G be an observable on a finite-dimensional Hilbert space, and let P(a|·):RR, θP(a|θ), aA with A some finite set, be real functions. Then the following statements are equivalent:

  • (i)
    There exists a quantum state ρ and a POVM {Ea}aA such that
    P(a|θ)=Tr(eiGθρe-iGθEa).
  • (ii)
    There exists an eigenbasis {|n}n of G, a probability distribution {pn}n over the eigenvectors |n, and positive semidefinite operators {Sa}aA with aASa=npn|nn| such that
    P(a|θ)=+|e-iGθSaeiGθ|+,
    where |+:=n|n (note that this vector is not a normalized state).
  • (iii)
    For every eigenbasis {|n}n of G, there exists a probability distribution {pn}n over the eigenvectors |n, and positive semidefinite operators {Sa}aA with aASa=npn|nn| such that
    P(a|θ)=+|e-iGθSaeiGθ|+.

Moreover, the state ρ in (i) can always be chosen as a pure state, with real non-negative amplitudes in any fixed choice of eigenbasis of G.

Proof

To prove (i)(iii), write ρ=jkρjk|jk| in an arbitrary eigenbasis of G where G|j=gj|j (when G is degenerate, there exist values ij such that gi=gj). Note that

P(a|θ)=jkρjkei(gj-gk)θk|Ea|j=+|e-iGθSaeiGθ|+,

where Sa is defined by its matrix elements

(Sa)kj:=ρjk(Ea)kj.

In other words, Sa=ρEa for the Schur product . Since the Schur product of positive semidefinite matrices is positive semidefinite, so is Sa. Moreover, since aAEa=1, S:=aASa satisfies k|S|j=ρjkδkj, i.e. it is a diagonal matrix with a probability distribution on its diagonal (namely, the diagonal elements of ρ).

The implication (iii)(ii) is trivial. To prove (ii)(i), define |ψ:=npn|n, ρ:=|ψψ| (which is a pure state), and

Ea:=MSaM+1|A|Π0,

where M:=n:pn0pn-1/2|nn| and Π0:=n:pn=0|nn|. Then we have Ea0 and aEa=1. Note that pn=0 implies n|Sa|n=0, and thus m|Sa|n=0 for all m. Hence

Tr(eiGθρe-iGθEa)=+|e-iGθSaeiGθ|+=P(a|θ).

This proves the converse and the claim that ρ can always be chosen pure and with non-negative amplitudes in the eigenbasis {|n}.

Now, for every NN, consider the representation

Uθ(N):=j=-JJ1Neijθ,

and denote by QJ,A(N) the quantum spin-J correlations that can be obtained with a suitable state and measurement on the corresponding Hilbert space. Clearly, every representation of the form (4) is embedded into some Uθ(N) for N large enough, and so

QJANNQJ,A(N).

The next lemma will show that, in fact, all the correlation sets are the same, i.e. QJ,A(N)=QJ,A(1) for all N, and hence QJAQJ,A(1). Since the converse inclusion is trivial, we obtain QJA=QJ,A(1), and Lemma 33 shows that the representing state can always be chosen pure, and with non-negative real amplitudes in the given eigenbasis {|j}. This establishes the validity of Theorem 1.

Lemma 34

We have QJ,A(1)=QJ,A(N) for all NN.

Proof

Let NN be an integer. On the Hilbert space which carries the representation Uα(N), define an eigenbasis {|j,n}-JjJ,1nN such that

Uα(N)|j,n=eijα|j,n,

i.e. where j labels the SO(2) irrep and n labels the multiplicity. For operators X=(j,m),(k,n)X(j,m),(k,n)|j,mk,n| on that Hilbert space, define an associated operator X~ on C2J+1 via X~j,k:=m,nX(j,m),(k,n) (regarding the Hilbert space as a tensor product space AB, where A=C2J+1 and B=CN, this is X~=+|BX|+B, with |+B:=n=1N|nB). Let us first show that X0 implies X~0. To this end, if |ψ=jψj|j is an arbitrary vector, set φj,m:=ψj for all m and |φ:=j,mφj,m|j,m. Then

ψ|X~|ψ=jkψj¯X~j,kψk=jkmnφj,m¯X(j,m),(k,n)φk,n=φ|X|φ0.

It is easy to see that if X=jmpj,m|j,mj,m| with {pj,m} some probability distribution, then X~=jqj|jj|, with {qj} another probability distribution (namely, qj=mpj,m).

Now suppose that PQJ,A(N), i.e. there is a quantum state ρ and a POVM {Ea}aA on the total Hilbert space such that

P(a|θ)=TrUθ(N)ρUθ(N)Ea.

Let G be a generator such that Uθ(N)=eiGθ, and let |+(N):=j,m|j,m. According to Lemma 33, (i)(iii), this implies that there are positive semidefinite matrices Sa and a probability distribution {pj,m} with aASa=j,mpj,m|j,mj,m| such that

P(a|θ)=+(N)(Uθ(N))SaUθ(N)|+(N)=jkmnei(k-j)θ(Sa)(j,m),(k,n)=jkei(k-j)θ(S~a)j,k=+|UθS~aUθ|+.

Thus, due to Lemma 33 (ii)(i), we have PQJ,A(1).

We conclude that QJ,A(N)QJ,A(1). Conversely, QJ,A(1)QJ,A(N) because the former can be trivially embedded into the latter by padding the states with zeroes and the POVM with constants that sum up to one.

2. Generalization of the rotation boxes SDP in Eq. (20) to arbitrary number of outcomes

Here, we generalize the SDP methodology in Eq. (20) to account for an arbitrary finite number of outcomes. Following the notation introduced in 2, let us denote the outcome set with outcomes as A={b1,,bn} with |A|=n and its corresponding set of spin-J correlations as RJA. Then, a generalization of Eq. (20) immediately follows as:

maxQb1,,Qbn-1,Sf(c,s)s.t.aki=0j,j+k2JQj,j+kifor allkandi,a~k=-0j,j+k2JSj,j+kfor allk0,1-a~0=Tr(S),Qb1,,Qbn-1,S0, B1

where the entries of Qb1,,Qbn-1,S are labelled from 0 to 2J, and we have defined a~k=i=1n-1aki. Note that the condition i=1nP(bi|θ)=1 removes one degree of freedom. Consequently, we take i{1,,n-1}, with P(bi|θ)=k=-JJakieikθ. The generalization follows immediately from Eq. (20), which is the specific case for n=2. In particular, the conditions involving Qbi imply 0P(bi|θ) for all i{1,,n-1},θR, and the constraints involving S imply i=1n-1P(bi|θ)1 and, thus, P(bi|θ)1 for all i{1,,n-1},θR. Finally, from i=1nP(bi|θ)=1 one can always find the missing 0P(bn|θ)1.

3. Proof of Lemma 2

The arguments in the main text already demonstrate that every QJA is a compact convex set, and that every P(a|θ) is a trigonometric polynomial of degree at most 2J, i.e. of the form

c0+k=12Jckcos(kθ)+sksin(kθ).

These are 4J+1 parameters. If we have |A| functions of this kind that sum to one, then this tuple is determined by (4J+1)(|A|-1) parameters. All we need to show is that we can generate a set of correlations of this dimension via quantum rotation boxes. Denote the standard basis in C2J+1 by {|j}j=-JJ, such that Uθ|j=eijθ|j. For =1,,2J, define the pair of matrices F(),G() componentwise:

Fkj():=δj-k,+δk-j,,Gkj():=i(δj-k,-δk-j,).

For example, if J=3/2 and =1, then

F()=0100101001010010,G()=i0100-10100-10100-10.

These band matrices are Hermitian. Consider the state |+:=12J+1j=-JJ|j, then

+|UθF()Uθ|+=f,Jcos(θ),+|UθG()Uθ|+=g,Jsin(θ),

where f,J,g,J0 are constants that only depend on and J. Now pick an arbitrary outcome a0A, and define a collection of Hermitian operators in the following way. If aa0, set

Ea:=c0(a)1+=12Jc(a)F()+s(a)G(),

and Ea0:=1-aa0Ea. If we choose the coefficients such that 0<c(a),s(a)c0(a), then every Ea for aa0 will be positive semidefinite, because the matrix c0(a) 1 is contained in the interior of the set of positive semidefinite matrices. Furthermore, if we choose the c0(a) small enough (but still non-zero), then Ea0 will also be positive semidefinite such that we obtain a valid POVM. By construction,

P(a|θ)=+|UθEaUθ|+=c0(a)+=12Jc(a)f,Jcos(θ)+s(a)g,Jsin(θ),

and varying the coefficients c(a),s(a) while respecting the necessary inequalities to have a POVM produces a set of tuples of trigonometric polynomials of full dimension.

4. Proof of Lemma 3

Let PQJA. Then P(a|θ)=Tr(UθρUθEa), with ρ some quantum state and {Ea}aA some POVM on the Hilbert space H=C2J+1=span{|j|-JjJ} (every j is an integer, or every j is a half-integer), while Uθ|j=eijθ|j. Consider the Hilbert space H:=C2J+1=span{|j|-JjJ}, where J:=J+12. Define the isometry W:HH via W|j:=|j+12, then W embeds H isometrically into H, and WW=1H and WW=1H-|-J-J|. Set ρ:=WρW, which is a quantum state on H, and Ea:=1|A||-J-J|+WEaW, then {Ea}aA is a POVM on H. Set Uθ|j:=eijθ|j, then

eiθ/2WUθW=Uθ-e-iJθ|-J-J|,

and hence

Tr(Uθρ(Uθ)Ea)=Tr(UθρUθEa)=P(a|θ),

and so PQJ+1/2A.

5. Proof of Lemma 4

We assume U:SO(2)U(H), U:θUθ, is a finite-dimensional unitary projective representation of SO(2) on the finite-dimensional complex Hilbert space H, and show that this entails the following three propositions:

  • (i): It is proven in Lemma 21 that any finite-dimensional unitary projective representation SO(2)U(V) is of the form given in Eq. (4) (see also Section 1 of the Supplemental Materials of [19]), where J is uniquely defined by the condition nJn-J0.

  • (i)(ii): Using the isomorphism of Lemma 28: LH(V)Sym(VV¯), ρρ and the dual isomorphism LH(V)Sym(VV¯), EE we have:
    Tr(UθρUθE)=E(UU¯)ρ. B2
    By Lemma 26, Sym(VV¯) is closed (as a real vector space) under the action of UU¯. Denoting P the projector P:VV¯Sym(VV¯) we have:
    E(UU¯)ρ=EP(UU¯)Pρ, B3
    where P(UU¯)PLR(Sym(VV¯)) the space of real linear operators on Sym(VV¯). Since ESym(VV¯) and ρSym(VV¯), the map UU¯E(UU¯)ρ can be linearly extended to a functional on LR(Sym(VV¯)) and hence it is a linear combination of the entries in UU¯. As can be seen easily, and is done explicitly below in (iii), these entries are trigonometric polynomials of order at most 2J, which entails Tr(UθρUθE) is a trigonometric polynomial of order at most 2J. And since these maps span LR(Sym(VV¯)) linearly, the degrees of the trigonometric polynomials Tr(UθρUθE) cannot all be strictly smaller than 2J.
  • (i)(iii): Denote the Hilbert space space on which the projective representation acts by HJ. The representation induced on the complex vector space of matrices L(HJ) is given by θUθUθ. Using the isomorphism L(HJ)HJHJHJH¯J with corresponding representations UθUθU(θ)U(θ)U(θ)U¯(θ), we obtain the following decomposition of L(HJ) into irreducible representations:
    U(θ)U¯(θ)=j=-JJ1njeijθk=-JJ1nke-ikθ B4
    =l=-2J2J1mleilθ. B5

The multiplicity ml for a given irreducible representation eilθ is given by

ml=j,k|j-k=lnj×nk. B6

In particular, note that m2J=nJn-J>0. This will imply that the representation on LH(HJ) is generated not by an arbitrary projective unitary representation of SO(2) but one specifically of the form (i), with the specific value of J.

The multiplicity m-l is equal to ml:

m-l=j,k|j-k=-lnj×nk=j,k|k-j=lnj×nk B7
=j,k|k-j=lnk×nj=ml. B8

From the equality

eikθ00e-ikθ=Lcos(kθ)-sin(kθ)sin(kθ)cos(kθ)L-1, B9

where

L=-ii11,L-1=12i1-i1, B10

and from ml=m-l, it follows that

l=-2J2J1mleilθ1m0k=12J1mkcos(kθ)-sin(kθ)sin(kθ)cos(kθ), B11

which is a decomposition of HJH¯J into real irreducible subspaces. By Lemma 26, the real subspace Sym(HJH¯J) carries the real representation of the above form. Thus, so does LH(HJ) due to Lemma 28.

We have thus shown that (i)(ii) and (i)(iii), hence all three statements are equivalent. Finally, we consider the specific case where UθJ:=eiθZJ, with ZJ=diag(J,J-1,,-J) The representation Γθ acts on the linear space spanned by density operators LH(C2J+1) as Γθρ=UθρUθ. Using again the isomorphism Uθ·UθUθU¯θ, Eq. (B6) in the special case nj=1 entails ml=2J+1-l.

6. Proof of Lemma 10

Suppose pRJ, then the Fejér–Riesz theorem implies that there is q(θ)=j=-JJbjeijθ such that p(θ)=|q(θ)|2. Thus

q(θ)¯q(θ)=j,k=-JJei(k-j)θbj¯bk=p(θ),

and hence ak=0j,j+k2Jbj¯bj+k. Define Qjk:=bj¯bk, then Q0 and the first condition in Lemma 10 follows. Similarly, 1-p(θ)0 implies the second and third condition.

Conversely, suppose that the first condition of Lemma 10 is satisfied. Then

p(θ)=k=-2J2Jakeikθ=k=-2J2J0j,j+k2JQj,j+keikθ=k=-2J2J=-2J2JQjei(-j)θ=v|Q|v0,

where vk=eikθ, and where we have used the substitution :=j+k. Similarly, the second and the third condition imply 1-p(θ)0 for all θ.

Appendix C. Proofs for Section 4.2

For clarity, we restate some of the lemmas or theorems before their proofs.

1. Proof of Lemma 13

Lemma 13. Every non-constant function pextR1 is contained in at least one face Fθ0,θ1.

Proof

It is sufficient to show that all pextR1 satisfy minθp(θ)=0 and maxθp(θ)=1. To show that the maximum is unity, let m:=maxθp(θ). Since p is not identically zero by assumption, we have m>0. Suppose that m<1. Then q(θ):=p(θ)/m is itself an element of R1, and p(θ)=m·q(θ)+(1-m)·0. Thus, p is not extremal in R1, which contradicts our assumption that it is. The proof that the minimum is zero is analogous.

2. Proof of Lemma 14

Lemma 14 gives an explicit characterization of the sets F0,θ1 showing for which values of θ1 the set is empty, and for values where F0,θ1 is non-empty, and hence a face of F0, it characterizes the functions in δextF0,θ1.

We first characterize the general form of the functions pF0, which are of interest since F0,θ1F0 for all θ1.

Lemma 35

Let p(θ) be a trigonometric polynomial of degree 2 or less with p(θ)0 for all θ and p(0)=0. Then there are constants c0, φ[0,2π) and 0s1 such that

p(θ)=c(1-cosθ)(1-scos(θ-φ)).
Proof

Due to the Fejér–Riesz theorem, there is a complex polynomial

h(z)=a0+a1z+a2z2

with p(θ)=|h(eiθ)|2; we can choose a2 to be a real number by absorbing complex phases into the definition of h. We have 0=p(0)=h(1), and thus a0=-a1-a2, hence h(z)=(z-1)(a2z+a1+a2). Write -(a1+a2)/a2=reiφ with r0 and φR, then

p(θ)=|h(eiθ)|2=|eiθ-1|2·a2eiθ+a1+a22=2a22(1-cosθ)eiθ-reiφ2=2a22(1-cosθ)(1+r2-2rcos(θ-φ))=c(1-cosθ)1-2r1+r2cos(θ-φ),

where c=2a22(1+r2), and s:=2r/(1+r2)[0,1].

From the previous lemma we can immediately determine the maximal number of roots for functions pR1:

Lemma 36

Every function pR1 reaches value p(θ)=0 at most twice and value p(θ)=1 at most twice

Proof

Consider a function pR1 such that p(θ0)=0. The function p(θ)=p(θ+θ0) is such that p(0)=0 and has the same number of roots as p(θ). Thus we can restrict ourselves to the case of function pR1 such that p(0)=0.

By Lemma 35 these functions have the form:

p(θ)=c(1-cosθ)(1-scos(θ-φ)), C1

which attains value 0 at θ=0 and at θ=φ if the parameter s=1. Thus p(θ) has at most two roots.

Conversely consider a function pR1 which reaches value p(θ1i)=1 for n points {θ11,...,θn}. The function p(θ)=1-p(θ) has n roots p(θ1i)=0 for θ1i{θ11,...,θn}. However, since pR1, n is at most two. Thus, p has at most two points θ1i such that p(θ1i)=1.

Note that compact convex faces have a well-defined dimensionality. We now show that for all faces (i.e. non-empty F0,θ1) the dimensionality is either 0 (i.e the face contains a single point) or 1 (the face is the convex hull of two distinct points).

Lemma 37

Let θ1π. Then either F0,θ1= or dim(F0,θ1)1.

Proof

Let pF0,θ1, then Lemma 35 shows that

p(θ)=c(1-cosθ)(1-scos(θ-φ)),

where c>0 is uniquely determined by the equation p(θ1)=1. Furthermore, θ1 is a local maximum, hence

0=p(θ1)=2ccosθ12-scos32θ1-φsinθ12.

Since 0<θ1<2π, we know that sinθ120, hence

cosθ12-scos32θ1-φ=0.

Suppose that cos32θ1-φ=0, then cosθ12=0, which implies θ1=π, which contradicts the assumptions of the lemma. Hence cos32θ1-φ0, and

s=cosθ12cos32θ1-φ.

But this implies that every pF0,θ1 is uniquely determined by the parameter φ. (Note that not all φ[0,2π) yield valid pF0,θ1, i.e. only a subset of [0,2π) is allowed as possible values for φ, but this observation does not affect the present argumentation.) Hence dim(F0,θ1)1.

Lemma 38

We have dim(F0,π)1.

Proof

Let pF0,π, then Lemma 35 implies

p(θ)=c(1-cosθ)(1-scos(θ-φ)), C2

where 0φ<2π and 0s1. Furthermore,

1=p(π)=2c(1+scosφ),

hence scosφ>-1 and

c=12(1+scosφ).

Substituting this into Eq. (C2), and using that π is a local maximum, the equation 0=p(π) implies

ssinφ=0.

Hence, either s=0 such that p(θ)=12(1-cosθ), or φ=π such that

p(θ)=1-cosθ2(1-s)(1+scosθ), C3

or φ=0 such that

p(θ)=1-cosθ2(1+s)(1-scosθ).

Equation (C3) contains the other two cases via s=0 and s-1, and we conclude that the single parameter -1s<1 determines the element of F0,π uniquely (note that we do not claim that all these values of s give valid functions in the face, just that they are all contained in this family of functions).

A compact convex set of dimension 1 has exactly 2 extremal points. Thus

Corollary 4

Every face F0,θ1 contains either one or two extremal points, depending on whether its dimension is 0 or 1 (in the former case, it contains only a single element).

The faces F0,θ1 contain those functions pR1 such that a global minimum is p(θ0)=0 and a global maximum p(θ1)=1. However, some functions in F0,θ1 can have multiple global maxima and minima, as we shall now see.

Lemma 39

Let θ0θ0[0,2π) be two distinct angles. Then there is a unique pR1 with p(θ0)=p(θ0)=0 and maxθp(θ)=1, and it is of the form

p(θ)=c(1-cos(θ-θ0))(1-cos(θ-θ0)),

with some suitable uniquely determined c>0.

Similarly, if θ1θ1[0,2π) are distinct angles, then there is a unique pR1 with p(θ1)=p(θ1)=1 and minθp(θ)=0, and it is of the form

p(θ)=1-c(1-cos(θ-θ0))(1-cos(θ-θ0)),

with some suitable uniquely determined c>0.

Proof

The latter statement follows from the former by considering q(θ):=1-p(θ). It is thus sufficient to prove the former statement. For symmetry reasons, it is enough to consider the case θ0=0. Due to Lemma 35,

p(θ)=c(1-cosθ)(1-scos(θ-φ)),

where c0, φ[0,2π) and 0s1. Since θ0>0, we have 1-cosθ00, and so p(θ0)=0 implies

1-scos(θ0-φ)=0.

This is only possible if s=1 and cos(θ0-φ)=1, hence φ=θ0, and so

p(θ)=c(1-cosθ)(1-cos(θ-θ0)), C4

and c>0 is uniquely determined by the condition maxθp(θ)=1.

Corollary 5

Every pR1 that either

  • attains the value 0 once and the value 1 twice, or

  • attains the value 1 once and the value 0 twice

is extremal in R1.

Actually, we can easily transform one of these into the other:

Lemma 40

Let pR1 as a (2π)-periodic function on R, and suppose that

p(θ0)=0,p(θ1)=1,p(θ0)=0.

Then the (2π)-periodic function

p~(θ):=1-p(θ0+θ1-θ), C5

is also an element of R1, and it satisfies

p~(θ0)=0,p~(θ1)=1,p~(θ1)=1,

where θ1:=θ0+θ1-θ0.

The proof is very simple and omitted. In general, we can consider the transformation

Tθ0,θ1:pp~,

where p~ is defined by Eq. (C5), which maps R1 onto itself and is linear. Moreover, the lemma above also shows that

Tθ0,θ1(Fθ0,θ1)=Fθ0,θ1,

i.e. it preserves the faces that we are interested in. The idea is that it maps one of the extremal point (with two zeros) to the other extremal point (with two ones).

Let us study whether functions can have more than two global maxima or minima.

Lemma 41

Given a function pR1 with p(θ0)=0 and p(θ1)=1 we have |θ0-θ1|π2.

Proof

This is the special case J=1 of Lemma 20.

From this it follows

Corollary 6

If 0θ1<π2 or if 3π2<θ1<2π then F0,θ1=.

Proof

The set F0,θ1 contains those functions in R1 such that p(θ0)=0 and p(θ1)=1, where θ0=0. For 0θ1<π2, we have |θ0-θ1|<π2. Thus, by Lemma 41, F0,θ1 is empty.

Similarly, since p(2π)=0, it also follows that for 3π2θ1<2π that the face F0,θ1 is empty.

By Lemma 36 a function pδextR1 has at most two global minima and at most two global maxima.

Corollary 7

A non-constant function pδextR1 with two global minima θ0 and θ0 and two global maxima θ1 and θ1 is such that θ0=θ0+π, θ1=θ0+π2 and θ1=θ1+π.

Proof

A function pδextR1 has global minimum p(θ0)=0 by Lemma 13. Thus, if it has two global minima, there is another θ0θ0 such that p(θ0)=0.

Similarly the global maxima of the function are reached for p(θ1)=p(θ1)=1.

By Lemma 41 we have the following relations:

|θ0-θ1|π2,|θ0-θ1|π2, C6
|θ0-θ1|π2,|θ0-θ1|π2. C7

Without loss of generality we assume θ0<θ0 and θ1<θ1.

This implies

|θ0-θ0|π,|θ1-θ1|π. C8

Thus, θ0 and θ0 must lie on antipodal points of the unit circle, and so do θ1 and θ1. Moreover, since θ0 and θ1 must have distance at least π/2, they must have distance exactly π/2, and the four extrema form the corners of a square inside the circle. This proves the claimed equations.

We now show that such a function exists and is unique.

Lemma 42

The only pR1 that have two distinct zeros and two distinct ones are

p(θ)=(1-cos(θ-θ0))(1+cos(θ-θ0)),

with 0θ0<π.

Proof

Since p(θ0)=p(θ0)=0 and maxθp(θ)=1 Lemma 35 implies that p(θ) has the form:

p(θ)=c·(1-cos(θ-θ0))(1-cos(θ-θ0)). C9

By Corollary 7, θ0=θ0+π, hence

p(θ)=c·(1-cos(θ-θ0))(1+cos(θ-θ0)), C10

and maxθp(θ)=1 implies that c=1.

Lemma 43

The unique global maximum of the function fθ0:[0,2π)R,

fθ0(θ):=(1-cosθ)(1-cos(θ-θ0)),

occurs at θ1=θ02+π when θ0(0,π) and at θ1=θ02 when θ0(π,2π).

Proof

Let us find local extrema:

fθ0(θ)=sin(θ)(1-cos(θ0-θ))-(1-cos(θ))sin(θ0-θ)=sin(θ0-2θ)-sin(θ0-θ)+sin(θ). C11

The equation fθ0(θ)=0 has the following solutions in [0,2π):

θ0,θ0,θ02,π+θ02ifθ0[0,π), C12
θ0,θ0,-π+θ02,θ02ifθ0[π,2π). C13

One can check directly that these are zeroes of fθ0. Moreover, since fθ0 is a trigonometric polynomial of degree 2, it has at most 4 zeroes (up to (2π)-periodicity), hence these are the only zeroes. Clearly, fθ0(θ) attains a global minimum for θ=0 and θ=θ0. Let us determine the global maximum:

fθ0θ02=1-cosθ022, C14
fθ0π+θ02=fθ0-π+θ02=1+cosθ022. C15

We see that fθ0θ02<fθ0θ02±π if and only if cosθ02>0. This implies that the unique global maximum occurs at θ1=π+θ02 when θ0(0,π), at θ1=θ02 for θ0(π,2π).

Lemma 44

If θ1=π2 or θ1=3π2, then F0,θ1 contains a single element, namely

F0,π2=F0,3π2=p(θ)=sin2θ.
Proof

Let 0θ1π2, and suppose that pF0,θ1. Consider T(θ):=2p(θ)-1. We have -1T(θ)1 for all θ, thus, we can use the result of [19, Theorem 2]

T(θ)+n2T(θ)2n2, C16

where n is the degree of the trigonometric polynomial (here n=2). Thus,

θ1=01dθ01T(θ)dθ21-T(θ)2=12T(0)T(θ1)dy1-y2=12arcsinT(θ1)-arcsinT(0)=π2.

This is a contradiction if θ1<π2, and so F0,θ1= in this case. On the other hand, to have equality in the case θ1=π2, we must have equality in Eq. (C16) for all 0θπ2, which implies that T(θ)=-cos(2θ). A similar calculation for 3π2θ1<2π proves the claim.

Lemma 45

Let θ1π2,3π2\{π}. Then F0,θ1 contains exactly two distinct extremal points,

extF0,θ1={p(θ),p~(θ)},

namely

p(θ)=c(1-cosθ)(1-cos(θ-θ0)),

and p~ is defined as in Eq. (C5). Here θ0=2θ1 for θ1(π2,π) and θ0=2(θ1-π) for θ1(π,3π2), and c>0 is uniquely determined by the condition maxθp(θ)=1.

Proof

Fix some θ1(π2,π). Then, by Lemma 43, the function fθ0 for θ0=2θ1 is such that fθ0(θ1) is its global maximum. For θ1(π,3π2), the function fθ0 with θ0=2(θ1-π) is such that fθ0(θ1) is its global maximum.

Set c:=1/fθ0(θ1) and p(θ):=cfθ0(θ), then p(0)=0, p(θ1)=1=maxθp(θ), and p(θ)0 for all θ, hence pF0,θ1. By Lemma 43, p(θ) reaches value 0 twice at θ=0,θ0, and value 1 at θ1. Hence, due to Corollary 5, p is extremal in R1 and thus also extremal in F0,θ1. Since p does not attain the value 1 twice, we have p~p. Moreover, for the same reason as for p, we have p~extF0,θ1.

We have discovered two distinct extremal points of F0,θ1. Since dimF0,θ11 according to Lemma 37, there cannot be any more extremal points.

The following uses the terminology of Lemma 38.

Lemma 46

The face F0,π contains exactly two extremal points, namely

F0,π={p(θ),p~(θ)},

where p(θ)=sin4θ2, and p~ is defined as in (C5) (concretely, p~(θ)=14(1-cosθ)(3+cosθ)).

Proof

Every pF0,π corresponds to some element of the family of functions ps defined in Eq. (C3), with -1s1. Indeed, the case s=-1 yields a valid function pF0,π, and since it is in the topological boundary of the parameter range, it must correspond to an extremal point of the one-dimensional face. But the reversible transformation T0,π maps extremal points to extremal points, and hence p~:=T0,πp must also be an extremal point of F0,π (in fact, it is the function corresponding to s=13). Since dimF0,π1 according to Lemma 38, these must be the only extremal points. (Note that this also shows that the face corresponds to the parameter range -1s13).

The four statements of Lemma 14 are now proven in Corollary 6, Lemmas 44, 45, and 46, respectively.

3. Proof of Theorem 6

Theorem

(Q1=R1) The correlation set R1 is equal to Q1.

By Lemma 6, we have Q1R1. To show the converse, we will use Lemma 15 and show that all correlations in δextR1 have a quantum spin-1 realization.

Lemma 47

If p(θ)Q1 then p(θ):=p(θ+θ0)Q1.

Proof

The assumption p(θ)Q1 implies that there is a quantum state ρ and a POVM element E such that

p(θ)=Tr(EUθρUθ), C17

hence

p(θ)=p(θ+θ)=Tr(EUθ+θρUθ+θ) C18
=Tr(EUθ(UθρUθ)Uθ)=Tr(EUθρUθ), C19

with ρ=(UθρUθ) a valid quantum state, hence p(θ)Q1.

Thus, we only need to show that the extremal points pδextR1 with p(0)=0 are quantum realizable.

Lemma 48

If p(θ)Q1 with p(θ0)=p(θ0)=0 and p(θ1)=1, then p~(θ):=1-p(θ0+θ1-θ)Q1.

Proof

p(θ)Q1 entails there exists a qutrit state ρ and a qutrit effect E such that

p(θ)=Tr(EUθρUθ), C20

where Uθ=diag(eiθ,1,e-iθ).

Define the effect E=1-Uθ0+θ1EUθ0+θ1, then:

p(θ)=Tr(EU-θρU-θ) C21
=Tr(1ρ)-Tr(EUθ0+θ1-θρUθ0+θ1-θ) C22
=1-p(θ0+θ1-θ)=p~(θ). C23

Since θU-θ is also a quantum spin-1 rotation box, this implies that p~Q1.

The above two lemmas and Lemma 14 imply that R1=Q1 follows from this lemma:

Lemma 49

The following functions are contained in Q1:

  1. p(θ)=sin2θ,

  2. p(θ)=sin4θ2,

  3. p(θ)=c(1-cosθ)(1-cos(θ-θ0)) for θ0(0,2π)\{π}, where c>0 is uniquely determined by the condition maxθp(θ)=1.

Proof

Consider the following SO(2) orbit for a quantum spin-1 system:

|ψ(θ)=12(eiθ|1-e-iθ|-1). C24

For effect E+=|ϕϕ| with |ϕ=12(|1+|-1), we obtain

P(+|θ)=|ϕ|ψ(θ)|2=14(eiθ-e-iθ)2=sin2θ. C25

This proves item 1. To show item 2., consider the following orbit:

|ψ(θ)=12(eiθ|1+2|0+e-iθ|-1), C26

and the effect E+=|ϕϕ| with |ϕ=12(-|1+2|0-|-1). They generate the conditional probability

P(+|θ)=1162-eiθ-e-iθ2=sin4θ2. C27

Finally, let us prove item (3). First, define

θ1:=θ02+πif0<θ0<π,θ02ifπ<θ0<2π.

Note that π2<θ1<3π2. Now define

α:=1-11-cosθ1,β:=12(1-cosθ1),

then |α|2+2|β|2=1. Consider the orbit

|ψ(θ)=α|0+βeiθ|1+βe-iθ|-1,

and the effect E+:=|ψ(θ1)ψ(θ1)|. Then we have

ψ(θ1)|ψ(θ)=cos(θ-θ1)-cosθ11-cosθ1,

and the square of this expression becomes

P(+|θ)=14sin4θ12(1-cosθ)(1-cos(2θ1-θ))=14sin4θ12(1-cosθ)(1-cos(θ-θ0)).

By construction, P(+|θ1)=1, and this is the maximal value over all θ. Thus, we have shown that the family of functions of item 3. is contained in Q1.

4. Proof of Lemma 17

Proof
  1. We first consider a quantum SO(2) rotation box and show that is has three perfectly distinguishable states belonging to a common SO(2) orbit.

    The following three vectors are an orthonormal basis of C3:
    |1=13(|0+|1+|2), C28
    |ω=13(|0+e2πi3|1+e4πi3|2), C29
    |ω2=13(|0+e4πi3|1+e2πi3|2). C30
    It is immediate that these states belong to the following U(1) orbit:
    |ψ(θ)=1000eiθ000ei2θ|1. C31
    Using the measurement {|ωaωa|}a=0,1,2 allows us to perfectly distinguish the three states |ψ(0)=|1,|ψ(2π3)=|ω,|ψ(4π3)=|ω2. By definition, the three probability distributions
    P(a|θ)=|ωa|ψ(θ)|2(a=0,1,2), C32
    are in Q1{0,1,2}R1{0,1,2}. Thus, according to Lemma 8, there is a measurement {ea}a=0,1,2 on R1 such that
    P(a|θ)=ea·ω(θ)(a=0,1,2). C33
    By construction, the measurement {ea}a=0,1,2 perfectly distinguishes the states {ω(0),ω(2π3),ω(4π3)} of R1,i.e.
    ea·ωb·2π3=Pab·2π3=δab(a,b=0,1,2).
    • (b)
      If there are n jointly perfectly distinguishable states, then there are also n jointly perfectly distinguishable pure states ω1,,ωn. In particular, there is an effect en with en·ω1==en·ωn-1=0, but en·ωn=1. Thus, ω1,,ωn-1 are n-1 disjoint pure states in a proper face of Ω1. However, by Theorem 1 of [45], there is no face with three or more pure states (aside from the whole state space), since all proper faces are at most one-dimensional.
  2. Consider the following states:
    ω(0)=11010,ωπ2=101-10, C34
    ω(π)=1-1010,ω3π2=10-1-10. C35
    We define the following effects:
    e±π2=1200120, C36
    e0,π=1212000, C37
    eπ2,3π2=1201200, C38
    One can straightforwadly check that these are indeed valid effects, i.e. they give values in [0, 1] when evaluated on the orbit of pure states ω(θ) (and therefore on the who convex set of states):
    e±π2·ω(θ)=12+cos(2θ)2[0,1], C39
    e0,π·ω(θ)=12+cos(θ)2[0,1], C40
    eπ2,3π2·ω(θ)=12+sin(θ)2[0,1]. C41
    The unit effect is:
    u=10000. C42
    In the following addition is defined mod 2π. The measurement {e±π2,u-e±π2} can be used to perfectly distinguish the state ω(θ) for θ{0,π2,π,3π2} from either of the states ω(θ±π2):
    e±π2·ω(θ)=1,θ{0,π}, C43
    e±π2·ω(θ)=0,θπ2,3π2. C44
    The measurement {e0,π,u-e0,π} can be used to perfectly distinguish ω(0) from ω(π):
    e0,π·ω(0)=1, C45
    e0,π·ω(π)=0. C46
    The measurement {eπ2,3π2,u-eπ2,3π2} can be used to perfectly distinguish ω(π2) from ω(3π2):
    eπ2,3π2·ωπ2=1, C47
    eπ2,3π2·ω3π2=0. C48
    Thus, any pair of states in {ω(0),ω(π2),ω(π),ω(3π2)} can be perfectly distinguished.
  3. From the existence of four pure pairwise perfectly distinguishable states {ω(0),ω(π2),ω(π),ω(3π2)}, violation of bit symmetry follows immediately for reversible transformations T(θ) of the form in Equation (34). Take for example the pairs of perfectly distinguishable states {ω(0),ω(π2)} and {ω(0),ω(π)}, then there is no reversible transformation T(ϕ) mapping one pair to the other, i.e. such that T(ϕ)ω(0)=ω(0) and T(ϕ)ω(π2)=ω(π). However, there exist other transformations T which are symmetries of Ω1 such as T=diag(1,1,-1,1,-1). We now show that bit symmetry is violated for all symmetries of Ω1, not just the SO(2) subgroup {T(θ)|θ[0,2π)}. Let us denote by G the group of all symmetries of Ω1. There exists a group invariant inner product ·,· such that Gx,Gy=x,y for all GG and x,yR5. As for every inner product, there is a positive definite symmetric matrix M>0, M=M, such that x,y=x·My. Group invariance implies that M commutes with all elements of G; in particular, [M,T(θ)]=0 for all θ. A straightforward calculation shows that this implies that M=diag(a,b,b,c,c) for some a,b,c>0. If all pairs of perfectly distinguishable pure states ω1,ω2 were related by a reversible transformation, then their invariant inner products ω1,ω2 would all be identical. But the following are inner products between pairs of perfectly distinguishable pure states:
    ω(0),ω3π2=a-c,ω(0),ω(π)=a-b+c,ω(0),ω2π3=a-12b-12c.
    For these to be identical, we would need to have b=c=0, which contradicts the positive definiteness of M. Thus, bit symmetry cannot hold.

Appendix D. SDP-Based Algorithm to Explore the Correlations Set Boundaries

Here we outline an algorithm to numerically explore and compare the boundary of the correlations sets QJ,RJ which in Sect. 4.3 has led to the derivation of an inequality proving QJRJ for J3/2. The idea is to first choose a plane in some direction of the trigonometric coefficients affine space, and then discretize a circle around its origin to use the SDP-based methodologies in Sect. 3.3 to probe the boundary of the sets QJ,RJ for that particular plane. In other words, we numerically find a 2D projection of the sets QJ,RJ in the trigonometric coefficient space.

In particular, the algorithm goes as follows:

  1. Select two directions v1:=(c1,s1), v2:=(c2,s2) in the (4J+1)-dimensional affine space to define the plane.

  2. Parametrize a direction in the plane p=cos(θ)v1+sin(θ)v2, for some angle θ.

  3. Use the SDP in Eq. (20) to find the boundary of RJ in the direction p and/or the see-saw methodology presented in Sect. 3.3 to approximate the boundary of QJ in the direction p.

  4. Repeat step 3 for all values of θ{0,,2π} to complete a full circle discretized up to desired numerical accuracy.

In Fig. 6 of the main text, we present an example of the final result for J=3/2 in the plane given by the directions v1=(c0,c1,c2,c3,s1,s2,s3)=(0,0,1,0,0,0,0) and v2=(0,0,0,0,0,0,1) (i.e., the c2-s3 plane).

Appendix E. Several Results and Proofs for Section 4.3

1. Proof of Lemma 19

In the following, we will denote the eigenvalues of any self-adjoint n×n matrix A in decreasing order by λ1(A),λ2(A),,λn(A) such that λ1(A)λ2(A)λn(A).

Lemma 50

Consider the 4×4 block matrix

M=0BB0,

where B is a 2×2 matrix. Then its eigenvalues are

λ1(M),λ2(M),λ3(M),λ4(M)=λ1(BB),λ2(BB),-λ2(BB),-λ1(BB).
Proof

We have

M2=BB00BB.

Thus, the squares of the eigenvalues of M are the eigenvalues of BB and BB, which are known to agree. Up to a sign, this determines the eigenvalues of M, and the signs in turn are determined by Tr(M)=0=iλi(M).

Applying this lemma to the matrix M[E], we obtain

λ1(M[E])=λ1(B[E]B[E]),

where B[E]=E20-iE300E31. It is straightforward to compute the eigenvalues of the matrix B[E]B[E], and the result proves Lemma 19.

2. Proof that β=13

The feasible set for the optimization problem in Eq. (38) is given by a polytope R with vertices {(0,0,0),(0,0,14),(0,14,0),(14,0,0),(14,0,14)} (see Fig. 7 for an illustration). Our goal is to compute the maximum of the function

f(x,y,z):=x+y+z+(x+y+z)2-4xz

over all (x,y,z)R. We find that f=0 has no solutions in the topological interior of R, hence the maximum must be attained on one of the lower-dimensional faces of this polytope.

Fig. 7.

Fig. 7

Region R, defined by the constraints x,y,z0,x+y1/4,y+z1/4

There are five two-dimensional faces F1,,F5, but f restricted to face Fi has no stationary points in the relative interior of Fi, for all i. For example, if we define the face F1 by the condition x=0, it is parametrized by 0y14 and 0z14. The function f becomes fF1(y,z)=2(y+z), and (yfF1,zfF1)=(2,2)(0,0) in the relative interior (where 0<y,z<14) of F1, and so f cannot have any local maxima there.

Thus, the global maximum must be attained on one of the eight edges E1,,E8 (one-dimensional faces) or one of the five vertices V1,,V5 (zero-dimensional faces). For seven of the edges, E1,,E7, f has no stationary points in their relative interior, but on one of the edges it does: define E8 as the points in R with x+y=14 and y+z=14, which we can parametrize via 0x14 and y=14-x, z=x, such that

fE8(x)=x+14+x+142-4x2.

Then f8(x)=0 has a solution in the interior 0<x<14, namely x=16, and fE8(16)=23. Indeed, this is the global maximum, since f attains only the values 0 and 12 on the vertices V1,,V5.

We thus find max(x,y,z)Rf(x,y,z)=f(16,112,16)=232β2. This gives the bound β13.

The bound can be attained by a POVM that satisfies |E02|2=|E20|2=16, |E03|2=|E30|2=112 and |E13|2=|E31|2=16. Using semidefinite programming, we found the following possible solution for E:

E=12016-i23012-i231616i23120i2316012,

for which the state ρ would be given by

ρ=1313213213132161613213216161321313213213.

3. Proof that the quantum correlations satisfy c2J-1+s2Jβ=13

Lemma 51

Let PQJ for J32, then its trigonometric coefficients as defined in Lemma 5 satisfy

c2J-1+s2J13.
Proof

The proof follows closely the lines of the J=3/2 case, proven in Sect. 4.3. Here we briefly describe the relevant adaptations. First, we have

(c2J-1+s2J)[p]=2Re(a2J-1[p])-2Im(a2J[p])=2Re(Q0,2J-1+Q1,2J)-2Im(Q0,2J)=2Re(E0,2J-1ρ0,2J-1+E1,2Jρ1,2J)-2Im(E0,2Jρ0,2J)=Tr(M[E]ρ), E1

where now the matrix M[E] is given, in block-matrix notation,

M[E]=02×202×(2J-3)B[E]0(2J-3)×20(2J-3)×(2J-3)0(2J-3)×2B(E)02×(2J-3)02×2,

and B[E]=E2J-1,0-iE2J,00E2J,1. Maximizing Eq. (E1) over all quantum states ρ will again give us the maximal eigenvalue of M[E]. Since

M[E]2=B[E]B[E]02×(2J-3)02×20(2J-3)×20(2J-3)×(2J-3)0(2J-3)×202×202×(2J-3)B[E]B[E],

we obtain again λ1(M[E])=λ1(B[E]B[E]), and this eigenvalue can be bounded exactly as in the (J=3/2)-case by using that

|E2J-1,0|2+|E2J,0|214,|E2J,0|2+|E2J,1|214.

We hence obtain exactly the same upper bound of 1/3.

4. Examples of correlations in RJ\QJ for J2

We begin with the case J7/2.

Lemma 52

For every J7/2, we have QJRJ.

Proof

For β0 and J1, consider the following trigonometric polynomial

pJ,β(θ):=12+14βsin(2Jθ)-34βk=12J-1122J-ksinkπ2+θ-Jπ. E2

This is a trigonometric polynomial (in θ) of degree 2J with s2J=14β and c2J-1=38β, coming from an educated guess based on numerical results. If we can show that it satisfies 0pJ,β(θ)1 for all θ, for some β that is sufficiently close to 1, we have a non-quantum rotation box, since 14+38=0.625. The polynomial has the following closed-form expression

pJ,β(θ)=12+12β(-3)·4-J(cosθ+2sin(Jπ))+fJ(θ)5+4sinθ,

where

fJ(θ)=4cos(θ-2Jθ)-cos(θ+2Jθ)+4sin(2Jθ)=3cosθcos2(Jθ)+(8+10sinθ)cos(Jθ)sin(Jθ)-3cosθsin2(Jθ)=cos(Jθ)sin(Jθ)·3cosθ4+5sinθ4+5sinθ-3cosθ·cos(Jθ)sin(Jθ).

The result must be between the smallest and largest eigenvalues of this matrix, and those eigenvalues turn out to be -5-4sinθ and 5+4sinθ. Thus,

-5-4sinθfJ(θ)5+4sinθ.

Since fJ(θ) is by far the dominant term in the numerator (the other part goes to zero exponentially in J), we have almost shown that pJ,β=1 is a valid rotation box. Now let us be more careful and scale a bit with β<1. Clearly, pJ,βRJ if and only if

β(-3)·4-J(cosθ+2sin(Jπ))+fJ(θ)5+4sinθ1forallθ.

But, due to what we have just shown, the left-hand side is upper-bounded by β(3·4-J(1+2)+1), and hence pJ,β=1/(1+9·4-J)RJ. This establishes a gap if

c2J-1+s2J=0.625β>13,

which is the case for all J7/2.

In what follows we treat the remaining cases J=3/2,2,5/2,3 on a case-by-case basis.

The way we proceed is by finding explicit counterexamples for each remaining J. These counterexamples have been found numerically via the following SDP based on Eq. (20):

maxQ,S,ac2J-1+s2Js.t.ak=0j,j+k2JQj,j+kfor allk,ak=-0j,j+k2JSj,j+kfor allk0,1-a0=Tr(S),Q,S0. E3

When the SDP is feasible, it finds some (2J+1)×(2J+1) matrices QS and some complex variables ak with k{0,,2J} thus obtaining a valid rotation box correlation (c.f. Lemma 10). Then, if these values lead to c2J-1+s2J>13 then the correlation goes beyond the quantum bound and we have the counterexample.

As an example, let us take the case J=3/2 with the coefficients c0=2/5,c1=0,c2=48/125,c3=0, s0=0,s1=6/25,s2=0,s3=32/125. Then, one can check that this forms a valid spin-3/2 correlation since one can define matrices Q,S0 fulfilling Lemma 10 such as

Q3/2:=112516-12i12-16i12i99i1212-9i9-12i16i1212i160,S3/2:=1125242i-1216i-2i27/211i-12-12-11i27/22i-16i-12-2i240.

Finally, observe that for this case we have (c2J-1+s2J)[p]=78125=0.624>13 and thus the point lies outside of Q3/2. The same follows for the remaining cases J=2,5/2,3, for which for the sake of completion we proceed to provide some numerically found examples and their corresponding QJ,SJ certificates.

J = 2. Consider now c0=1/2,c1=-17/250,c2=0,c3=19/50,c4=0, s0=0,s1=0,s2=87/500,s3=0,s4=6/25 such that (c2J-1+s2J)[p]=0.62>13 and to fulfill Lemma 10 define the following matrices:

Q2:=377/2400-35/2438-95/2314i11/782-31/743i19/200-3/25i-35/2438+95/2314i62/811-4/1513+95/2314i-273/9704-3/844i19/20011/782+31/743i-4/1513-95/2314i243/7378-4/1513+95/2314i11/782-31/743i19/200-273/9704+3/844i-4/1513-95/2314i62/811-35/2438-95/2314i3/25i19/20011/782+31/743i-35/2438+95/2314i377/2400,S2:=377/240035/2438-95/2314i11/782+31/743i-19/2003/25i35/2438+95/2314i62/8114/1513+95/2314i-273/9704+3/844i-19/20011/782-31/743i4/1513-95/2314i243/73784/1513+95/2314i11/782+31/743i-19/200-273/9704-3/844i4/1513-95/2314i62/81135/2438-95/2314i-3/25i-19/20011/782-31/743i35/2438+95/2314i377/2400.

J = 5/2. In this case one can take c0=0.5261,c1=0,c2=-0.1044,c3=0,c4=0.3695,c5=0, s0=0,s1=-0.0639,s2=0,s3=0.1926,s4=0,s5=0.2564 such that (c2J-1+s2J)[p]=0.626>13 and to fulfill Lemma 10 define the following matrices

graphic file with name 220_2024_5123_Equ350_HTML.gif

J = 3. Finally, in this case one can take c0=1/2,c1=0.0173,c2=0,c3=-0.0915,c4=0,c5=0.3763,c6=0, s0=0,s1=0,s2=-0.0433,s3=0,s4=0.1864,s5=0,s6=0.2485 such that (c2J-1+s2J)[p]=0.6248>13 and to fulfill Lemma 10 define the following matrices

graphic file with name 220_2024_5123_Equ351_HTML.gif

Appendix F. Proofs for Section 4.4: J

Here we will present the details of the proof of Theorem 8. The first step of the proof in the main text is given by Lemma 53, the second step by Lemma 54 and the final and third step is presented right after the proof of Lemma 54. We will consider the Hilbert space L2(SO(2)), with inner product

f,g=12π02πf(θ)¯g(θ)dθ.

It carries the regular representation of SO(2), defined by (U(θ)f)(θ):=f(θ+θ). As usual, we will pick a representative f of [f]L2(SO(2)) whenever we do concrete calculations. All angle additions (like θ+θ or θ0-1/n) are understood modulo (2π).

Lemma 53

Let PR, then we can write it as a limit of a convergent sequence P(+|θ0+θ)=limnU(θ)fθ0,n|P^U(θ)fθ0,n, where fθ0,nL2(SO(2)) for all nN and θ0, while 0P^1.

Proof

For the choice of PR, we begin by defining an associated operator on L2(SO(2))

(P^ψ)(θ):=P(θ)ψ(θ).

It is easy to see that P^ is a bounded, self-adjoint operator. Furthermore,

ψ|P^|ψ=12π02πψ(θ)¯P(θ)ψ(θ)dθ[0,ψ,ψ],

and so 0P^1, i.e. P^ defines a valid POVM element.

We define

fθ0,n:=πnχ[θ0-1n,θ0+1n], F1

where

χ[θ0-1n,θ0+1n](θ)=1ifθ0-1nθθ0+1n0else, F2

and it is clear that fθ0,nL2(SO(2)). Furthermore, it is easy to show that fθ0,n=1 for all θ,n.

Now, for 0<θ0<2π and n large enough, we calculate

(P^-P(θ0)1)fθ0,n2=12πθ0-1nθ0-1n(P(θ)-P(θ0))2fθ0,n2(θ)dθ(P(Δmax(n))-P(θ0))2fθ0,n2=(P(Δmax(n))-P(θ0))2,

where θ0-1/nΔmax(n)θ0+1/n is chosen such that (P(Δmax(n))-P(θ0))2 is maximal. Since P is continuous, it follows that (P^-P(θ0)1))fθ0,n0 for n. Now we use the Cauchy-Schwarz inequality to show

(P^-P(θ0)1)fθ0,n=fθ0,n·(P^-P(θ0)1)fθ0,nfθ0,n|(P(θ^-P(θ0)1))fθ0,n.

Hence,

limnfθ0,n|(P^-P(θ0)1)fθ0,n=0. F3

We can rewrite this as

limnfθ0,n|P^fθ0,n=P(θ0).

The above is also true for θ=0 if all angles are understood modulo 2π. In a final step, we consider the transformation of fθ0,n under the regular representation U of SO(2):

U(θ)fθ0,n(θ)=fθ0,n(θ+θ)=fθ0-θ,n(θ). F4

Hence we can write

P(+|θ0+θ)=limnU(θ)fθ0,n|P^U(θ)fθ0,n=P(θ0+θ). F5

The claim follows.

Lemma 53 implies that given any PR, we can approximate it arbitrarily well by

Pn(+|θ0+θ)=U(θ)fθ0,n|P^U(θ)fθ0,n. F6

The following standard definitions can be found, for example, in Ref. [53].

Definition 5

For two probability distributions {px} and {qx} we define the classical trace distance by

D~(px,qx)=12x|px-qx|.

We observe that for x{±}, we have

D~(px,qx)=12(|p+-q+|+|p--q-|)=|p+-q+|.

Definition 6

Let AT(H), we define the norm

A1=Tr(|A|),

where |A|=AA.

Definition 7

Let A,BT(H),we define the trace distance

D(A,B)=12A-B1. F7

We will write

σθ0,n=|fθ0,nfθ0,n|. F8

From the Peter-Weyl Theorem [69], we know that L2(SO(2))=jZHj, where

Hj=span{ϕj|ϕj(α)=eijα}.

In the orthonormal basis {ϕj}jZ, we can write

U(θ)=j=-eijθ|ϕjϕj|. F9

Furthermore, we define the projector ΠJ onto the finite-dimensional subspace HJ=j=-JJHj by

ΠJ=j=-JJ|ϕjϕj|.

We write

σθ0,nJ=ΠJσθ0,nΠJTr(ΠJσθ0,n), F10
P^J=ΠJP^ΠJ, F11
UJ(θ)=ΠJU(θ)ΠJ, F12

where σθ0,nJS(HJ),PJ(θ^)E(HJ),UJ(θ)U(HJ) and UJ:SO(2)U(HJ) defined by θUJ(θ) is a representation of SO(2), because UJ(θ)=diag(e-iJθ,e-i(J-1)θ,,ei(J-1)θ,eiJθ). We denote

PnJ(+|θ0+θ)=Tr(P^J(UJ)(θ)σθ0,nJUJ(θ)). F13

By observing that U(θ)ΠJ=ΠJU(θ), we find

PnJ(+|θ0+θ)=Tr(P^J(UJ)(θ)σθ0,nJUJ(θ))=Tr(ΠJP^ΠJ2U(θ)ΠJσθ0,nJΠJU(θ)ΠJ)=Tr(P^ΠJU(θ)σθ0,nJU(θ)ΠJ2)=Tr(P^U(θ)ΠJσθ0,nJΠJU(θ))=Tr(P^U(θ)σθ0,nJU(θ)), F14

where we have used in the third line that the trace is cyclic.

Lemma 54

Suppose that Tr(ΠJσθ0,n)1-ϵ then ϵ|Pn(+|θ0+θ)-PnJ(+|θ0+θ)|.

Proof

The Gentle Measurement Lemma [52] states that if Tr(ΠJσθ0,n)1-ϵ then

σθ0,n-σθ0,nJ12ϵ

holds. Furthermore, we will use Theorem 9.1 from [53], which states

D(ϱ,σ)=max{Em}D~(pm,qm), F15

where the maximization is over all POVMs {Em}, pm=Tr(ϱEm)) and qm=Tr(σEm). We show

ϵD(σθ0,n,σθ0,nJ)=maxEmD~(Tr(σθ0,nEm),Tr(σθ0,nJEm))D~(Pn(θ0+θ),PnJ(θ0+θ))=|Pn(+|θ0+θ)-PnJ(+|θ0+θ)|, F16

where PnJ(θ0+θ) denotes the probability distribution {(PnJ(+|θ0+θ),(PnJ(-|θ0+θ)=1-(PnJ(+|θ0+θ)}, and in the third line we have used that {P+=U(θ)P^U(θ),P-=1-P+} is a POVM.

Let us check that ΠJ1 for J strongly. From the Peter-Weyl Theorem, we know that {ϕj}j=- defines an orthonormal basis of L2(SO(2)) and hence

(ΠJ-1)ψ2=j=-JJϕj|ψϕj-ψ2=|j|>J|ϕj|ψ|2=1-j=-JJ|ϕj|ψ|2J0,

which is true for every ψL2(SO(2)), and thus the claim follows. The last observation implies that we can make ϵ arbitrarily small by making J larger and larger.

Everything said so far in this section can be easily generalized to more than two (say, N) measurement outcomes. Let us define an N-outcome rotational box as N continuous non-negative real functions Pk on the unit circle such that k=1NPk(θ)=1 for every θ[0,2π). Similarly as above, we have associated operators P^k, defining a POVM, and we can project those into the subspaces HJ via P^kJ:=ΠJP^kΠJ. The approximating measurement on this spin-J system will have POVM elements P^1J,,P^N-1J,1-P^1J--P^N-1J. Adaption of all further proof steps from above is straightforward and proves the analogous result for N-outcome rotation boxes.

Appendix G. Proofs for Section 5

1. Proof of Theorem 9

First note that

P(-a,b|α+π,β)=Pb,βA(-a|α+π)PB(b|β)=Pb,βA(a|α)PB(b|β)=P(a,b|α,β),

and similarly, P(a,-b|α,β+π)=P(a,b|α,β). Therefore, P(a,b|α,β) can be determined from the values of acosα, asinα, bcosβ and bsinβ. In particular, we can find a function f, defined on two copies of the circle {(1,x,y)|x2+y2=1}, such that

P(a,b|α,β)=f(1,acosα,asinα),(1,bcosβ,bsinβ).

Let a=1 and α1=0, α2=π/4 and α4:=π/2, and ei:=(1,acosαi,asinαi) for i=1,2,3, then these three vectors are linearly independent and span R3. Let g:R3×R3R be the bilinear form that satisfies g(ei,ej)=f(ei,ej) for i,j=1,2,3.

Now suppose we fix some value of b and of β, then

P(a,b|α,β)=Pb,βA(a|α)PB(b|β)=12+c1acosα+s1asinαPB(b|β),

where c1 and s1 may depend on b and β. For every fixed b and β, this is a linear functional of the vector (1,acosα,asinα). Similar argumentation applies to the roles of A and B exchanged. Thus, f and g must agree on f’s domain of definition, and so

P(a,b|α,β)=g(1,acosα,asinα),(1,bcosβ,bsinβ).

Now, every 2×2 Hermitian matrix MLH(C2) can be parameterized in the form

M=12r0+r3r1-ir2r1+ir2r0-r3

(for r0=1, this is the well-known Bloch representation of quantum states). Define the linear map r(M):=(r0,r1,r2), dropping the r3-component. Finally, define the bilinear form ω:LH(C2)×LH(C2)R via

ω(M,N):=gr(M),r(N).

This bilinear form is unital:

ω(1,1)=g(2,0,0),(2,0,0)=g(1,1,0),(2,0,0)+g(1,-1,0),(2,0,0)=g(1,1,0),(1,1,0)+g(1,1,0),(1,-1,0)+g(1,-1,0),(1,1,0)+g(1,-1,0),(1,-1,0)=P(+1,+1|0,0)+P(+1,-1|0,0)+P(-1,+1|0,0)+P(-1,-1|0,0)=1.

Let us now show that ω(M,N)0 if M and N are positive semidefinite. If M0, then r0=Tr(M)0, and non-negativity of the eigenvalues enforces r12+r22+r32r02, hence r12+r22r02. Hence r(M) lies in the disc of radius r0, and can thus be written as a convex combination of points on the circle. Since g is bilinear, this will give the corresponding convex combination of values, and it is thus sufficient to restrict our attention to the case that r12+r22=r02. In this case, there will be some angle α such that (r1,r2)=(r0cosα,r0sinα). Similar reasoning for the matrix N0 (denoting the first component of r(N) by s0) yields

ω(M,N)=g(r0,r0cosα,r0sinα),(s0,s0cosβ,s0sinβ)=r0s0g(1,cosα,sinα),(1,cosβ,sinβ)=r0s0P(+1,+1|α,β)0.

Set M±(α):=e-iαZ|±±|eiαZ, where |±:=12(|0±|1), and similarly for M±(β), then

ω(Ma(α),Nb(β))=g(1,acosα,asinα),(1,bcosβ,bsinβ)=P(a,b|α,β).

It follows from the results of Barnum et al. [61] (see also Acín et al. [62] for a simplified proof, and Kleinmann et al. [70]) that there is a quantum state ρAB on the two qubits and a positive unital linear map τ:LH(C2)LH(C2) such that

ω(M,N)=TrρABMτ(N).

This completes the proof.

Appendix H. Proofs for Section 6: Connections to Other Topics

1. Background on transitive GPTs

We briefly introduce some necessary background on transitive GPT systems and refer the reader to [35] for a more complete introduction.

A finite-dimensional transitive GPT system is one with pure states X and dynamical group G which is compact (this includes the possibility of finite groups). The space of pure states X is isomorphic to G/H with H the stabilizer subgroup.

To each transitive GPT system is associated a representation of G which we denote ρ:GGL(V). Let us denote its decomposition into irreps by

ViIVi, H1
ρ(g)iIρi(g), H2

where I may contain repeated entries.

By transitivity, the state space can be obtained by applying the representation ρ(g) to a reference pure state ωxV:

ωgx=ρ(g)ωx, H3

which is necessarily invariant under ρ(h) for hHx, the stabilizer of x. The state ωx has support in every irrep Vi for iI (this is in fact not an assumption but follows from what it means for the representations ρ(g) to be associated to the system).

It follows from Theorem 2 of [35] that when (GH) form a Gelfand pair, any two transitive GPT systems with associated representations I and J which are equal as sets (i.e. contain the same irreducible representations ignoring repetitions) are equivalent as GPT systems (assuming that they are effect unrestricted). Two GPT systems (Ω1,E1,Γ1) and (Ω2,E2,Γ2) with associated vector spaces V1 and V2 and with dynamical group G are equivalent if there exists an invertible linear transformation L:V1V2 relating them:

L(Ω1)=Ω2, H4
E1L-1=E2, H5
LΓ1L-1=Γ2. H6

2. Proof of Theorem 11

We will make use of the following lemma:

Lemma 55

The symmetric product states |ψψ|dD(Symd(R2)) have full support in Sym2d(R2) and therefore have support in one copy of every irrep in {0,...,d}.

Proof

The rebit pure states |ψ transform under the real projective irreducible representation 12 of SO(2); a generic rebit state can be written as:

|ψ=cosθ2|++sinθ2|-. H7

If we complexify the vector space this is equal to:

|ψ=12(eiθ2|0+e-iθ2|1). H8

Hence the symmetric product states of d rebits |ψψ|dD(Symd(R2)) are isomorphic (there exists an equivariant invertible linear map) to the product states |ψ(θ)ψ(θ)|dD(Symd(C2)) where |ψ(θ)=12(eiθ2|0+e-iθ2|1). Using the isomorphism L(Symd(C2))Symd(C2)Symd(C2)¯ and Symd(C2))Symd(C2)¯ (by Corollary 3), we have the following isomorphism: L(Symd(C2))Symd(C2)Symd(C2),

|ψ(θ)ψ(θ)|d|ψ(θ)2d. H9

Expanding |ψ(θ)2d gives:

|ψ(θ)2d=12d(eidθ|02d+ei(d-1)θ(|02d-1|1+|02d-2|1|0+...+|1|02d-1)+...+e-idθ|12d)=122dj=02dei(d-j)θPSym|02d-j|1j H10
=12dj=02dx{0,1}2d|H(x)=jei(d-j)θ|x, H11

where PSym|v11...|vdd=σΣd|v1σ-1(1)...|vdσ-1(d), Σd the symmetric group on d elements and H(x) the Hamming weight of the bit string x.

Each |k=d-j=x{0,1}2d|H(x)=j|x belongs to a subspace carrying a projective representation k of SO(2). Thus, |ψ(θ)2d has support on a copy of every complex irrep {d,-d+1,...,d}. The projection of |ψ(θ)2d on the subspace carrying the representation {-k,k} is:

12d(eikθ|k+e-ikθ|-k)=12dcos(kθ)|k+|-k2+sin(kθ)|k+|-k2, H12

implying that |ψ(θ)2d it has support in every real irrep {0,...,d}. This implies |ψ(θ)ψ(θ)|d has suport in every real irrep {0,...,d}, and so |ψψ|d does also.

Hence,

ΩSymd:=conv{(|ψψ|)d||ψPR2}, H13

where PR2 is the set of rebit pure states, is the state space of a transitive GPT system with pure states SO(2) and dynamical group SO(2). It has associated to it the real representation {0,..,d}. The stabilizer subgroup is just the trivial group I={1}.

The set of unrestricted effects on ΩSymd is given by

ESymd:={E|ELS(Symd(R2),0Tr(E|ψψ|d)1}. H14

Since (SO(2),I) forms a Gelfand pair, it follows from [35, Theorem 2 (iii)] that all unrestricted GPT systems generated by applying a real representation {0,..,d} of SO(2) to a reference vector with support in each irrep are equivalent.

Hence the GPT systems (ΩSymd,ESymd) and (Ωd2,Ed2) are equivalent as GPT systems and generate the same SO(2) correlations.

Funding

Open access funding provided by Österreichische Akademie der Wissenschaften We acknowledge support from the Austrian Science Fund (FWF) via project P 33730-N. This research was supported in part by 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. A. A. also acknowledges financial support by the ESQ Discovery programme (Erwin Schrödinger Center for Quantum Science & Technology), hosted by the Austrian Academy of Sciences (ÖAW).

Data availability

Data sharing is not applicable to this article as no data sets were generated or analyzed during the current study.

Declarations

Conflict of interest

The authors have no relevant financial or non-financial interests to disclose.

Footnotes

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

References

  • 1.Mayers, D., Yao, A.: Quantum cryptography with imperfect apparatus. In: Proceedings 39th Annual Symposium on Foundations of Computer Science, pp. 503–509. IEEE (1998)
  • 2.Barrett, J., Hardy, L., Kent, A.: No signaling and quantum key distribution. Phys. Rev. Lett. 95, 010503 (2005) [DOI] [PubMed] [Google Scholar]
  • 3.Acín, A., Brunner, N., Gisin, N., Massar, S., Pironio, S., Scarani, V.: Device-independent security of quantum cryptography against collective attacks. Phys. Rev. Lett. 98, 230501 (2007) [DOI] [PubMed] [Google Scholar]
  • 4.Gallego, R., Brunner, N., Hadley, C., Acín, A.: Device-independent tests of classical and quantum dimensions. Phys. Rev. Lett. 105, 230501 (2010) [DOI] [PubMed] [Google Scholar]
  • 5.Brunner, N., Cavalcanti, D., Pironio, S., Scarani, V., Wehner, S.: Bell nonlocality. Rev. Mod. Phys. 86, 419 (2014) [Google Scholar]
  • 6.Scarani, V.: Bell Nonlocality. Oxford Graduate Texts. OUP, Oxford (2019) [Google Scholar]
  • 7.Pawłowski, M., Brunner, N.: Semi-device-independent security of one-way quantum key distribution. Phys. Rev. A 84, 010302 (2011) [Google Scholar]
  • 8.Liang, Y.C., Vértesi, T., Brunner, N.: Semi-device-independent bounds on entanglement. Phys. Rev. A 83, 022108 (2011) [Google Scholar]
  • 9.Branciard, C., Cavalcanti, E., Walborn, S., Scarani, V., Wiseman, H.M.: One-sided device-independent quantum key distribution: security, feasibility, and the connection with steering. Phys. Rev. A 85, 010301 (2012) [Google Scholar]
  • 10.Van Himbeeck, T., Woodhead, E., Cerf, N.J., García-Patrón, R., Pironio, S.: Semi-device-independent framework based on natural physical assumptions. Quantum 1, 33 (2017) [Google Scholar]
  • 11.Garner, A.J.P., Krumm, M., Müller, M.P.: Semi-device-independent information processing with spatiotemporal degrees of freedom. Phys. Rev. Res. 2, 013112 (2020) [Google Scholar]
  • 12.Cirel’son, B.S.: Quantum generalizations of Bell’s inequality. Lett. Math. Phys. 4, 93–100 (1980) [Google Scholar]
  • 13.Giovannetti, V., Lloyd, S., Maccone, L.: Advances in quantum metrology. Nat. Photon 5, 222–229 (2011) [Google Scholar]
  • 14.Tóth, G., Apellaniz, I.: Quantum metrology from a quantum information science perspective. J. Phys. A Math. Theor. 47(42), 424006 (2014) [Google Scholar]
  • 15.Popescu, S.: Nonlocality beyond quantum mechanics. Nat. Phys. 10, 264 (2014) [Google Scholar]
  • 16.Khalfin, L.A., Tsirelson, B.S.: Quantum and quasi-classical analogs of Bell inequalities. In: Symposium on the Foundations of Modern Physics, vol. 85, p. 441 (1985)
  • 17.Popescu, S., Rohrlich, D.: Quantum nonlocality as an axiom. Found. Phys. 24, 379–385 (1994) [Google Scholar]
  • 18.van Himbeeck, T., Pironio, S.: Correlations and randomness generation based on energy constraints. arXiv:1905.09117 (2019)
  • 19.Jones, C.L., Ludescher, S.L., Aloy, A., Müller, M.P.: Theory-independent randomness generation with spacetime symmetries. arXiv:2210.14811
  • 20.Mandelstam, L., Tamm, I.: The uncertainty relation between energy and time in non-relativistic quantum mechanics. J. Phys. (USSR) 9, 249 (1945) [Google Scholar]
  • 21.Anandan, J., Aharonov, Y.: Geometry of quantum evolution. Phys. Rev. Lett. 65, 1697 (1990) [DOI] [PubMed] [Google Scholar]
  • 22.Hörnedal, N.: Generalizations of the Mandelstam–Tamm quantum speed limit. Master’s thesis, Stockholm University (2021)
  • 23.Marvian, I., Spekkens, R.W., Zanardi, P.: Quantum speed limits, coherence, and asymmetry. Phys. Rev. A 93, 052331 (2016) [Google Scholar]
  • 24.Nagata, K., Laskowski, W., Wieśniak, M., Żukowski, M.: Rotational invariance as an additional constraint on local realism. Phys. Rev. Lett. 93, 230403 (2004) [DOI] [PubMed] [Google Scholar]
  • 25.Navascués, M., Guryanova, Y., Hoban, M.J., Acín, A.: Almost quantum correlations. Nat. Commun. 6, 6288 (2015) [DOI] [PubMed] [Google Scholar]
  • 26.Sanyal, R., Sottile, F., Sturmfels, B.: Orbitopes. Mathematika 57(2), 275–314 (2011) [Google Scholar]
  • 27.Wald, R.M.: General Relativity. University of Chicago Press, Chicago (2010) [Google Scholar]
  • 28.Barrett, J., Colbeck, R., Kent, A.: Memory attacks on device-independent quantum cryptography. Phys. Rev. Lett. 110, 010503 (2013) [DOI] [PubMed] [Google Scholar]
  • 29.Hardy, L.: Quantum theory from five reasonable axioms. arXiv:quant-ph/0101012 (2001)
  • 30.Barrett, J.: Information processing in generalized probabilistic theories. Phys. Rev. A 75, 032304 (2007) [Google Scholar]
  • 31.Müller, M.P.: Probabilistic theories and reconstructions of quantum theory. In: SciPost Physics Lecture Notes, vol. 28 (2021)
  • 32.Plávala, M.: General probabilistic theories: an introduction. Phys. Rep. 1033, 1–64 (2023) [Google Scholar]
  • 33.Holevo, A.: Probabilistic and Statistical Aspects of Quantum Theory. Edizione della Normale. Springer, Pisa (2011) [Google Scholar]
  • 34.Schmid, D., Selby, J.H., Wolfe, E., Kunjwal, R., Spekkens, R.W.: Characterization of noncontextuality in the framework of generalized probabilistic theories. PRX Quantum 2, 010331 (2021) [Google Scholar]
  • 35.Galley, T.D., Masanes, L.: How dynamics constrains probabilities in general probabilistic theories. Quantum 5, 457 (2021) [Google Scholar]
  • 36.Sanyal, R., Sottile, F., Sturmfels, B.: Orbitopes. Mathematika 57(2), 275–314 (2011) [Google Scholar]
  • 37.Fröhner, F.H.: Missing link between probability theory and quantum mechanics: the Riesz–Fejer theorem. Z. Naturforschung A 53(8), 637–654 (1998) [Google Scholar]
  • 38.Blekherman, G., Parrilo, P.A., Thomas, R.R.: Semidefinite optimization and convex algebraic geometry. In: MOS-SIAM Series on Optimization, vol. 13 (2012)
  • 39.Skrzypczyk, P., Cavalcanti, D.: Semidefinite programming in quantum information science. arXiv:2306.11637 (2023) [DOI] [PubMed]
  • 40.Navascués, M., Pironio, S., Acín, A.: A convergent hierarchy of semidefinite programs characterizing the set of quantum correlations. New J. Phys. 10, 073013 (2008) [Google Scholar]
  • 41.Pál, K.F., Vértesi, T.: Maximal violation of a bipartite three-setting, two-outcome Bell inequality using infinite-dimensional quantum systems. Phys. Rev. A 82, 022116 (2010) [Google Scholar]
  • 42.Werner, R.F., Wolf, M.M.: Bell inequalities and entanglement. arXiv:quant-ph/0107093 (2001)
  • 43.Wright, V.J., Weigert, S.: General probabilistic theories with a Gleason-type theorem. Quantum 5, 588 (2021) [Google Scholar]
  • 44.Webster, R.: Convexity. Oxford University Press, Oxford (1994) [Google Scholar]
  • 45.Smilansky, Z.: Convex hulls of generalized moment curves. Isr. J. Math. 52(1–2), 115–128 (1985) [Google Scholar]
  • 46.Müller, M.P., Ududec, C.: Structure of reversible computation determines the self-duality of quantum theory. Phys. Rev. Lett. 108, 130401 (2012) [DOI] [PubMed] [Google Scholar]
  • 47.Müller, M.P., Garner, A.J.P.: Testing quantum theory by generalizing noncontextuality. Phys. Rev. X 13, 041001 (2023) [Google Scholar]
  • 48.Barnum, H., Hilgert, J.: Spectral properties of convex bodies. J. Lie Theory 30, 315–355 (2020) [Google Scholar]
  • 49.Bhatia, R.: Matrix Analysis. Springer, New York (1997) [Google Scholar]
  • 50.Baksalary, O.M., Trenkler, G.: On the entries of orthogonal projection matrices. In: Bapat, R., Kirkland, S., Prasad, K., Puntanen, S. (eds.) Combinatorial Matrix Theory and Generalized Inverses of Matrices. Springer, India (2013) [Google Scholar]
  • 51.Rudin, W.: Real and Complex Analysis. McGraw-Hill, New York (1987) [Google Scholar]
  • 52.Wilde, M.: Quantum Information Theory. Cambridge University Press, Cambridge (2017) [Google Scholar]
  • 53.Nielsen, M., Chuang, I.: Quantum Computation and Quantum Information, 10th Anniversary Cambridge University Press, Cambridge (2010) [Google Scholar]
  • 54.DeVore, R.A., Lorentz, G.G.: Constructive Approximation. Springer, Berlin (1993) [Google Scholar]
  • 55.Spekkens, R.W.: Quasi-quantization: classical statistical theories with an epistemic restriction. In: Chiribella, G., Spekkens, R. (eds.) Quantum Theory: Informational Foundations and Foils, Fundamental Theories of Physics 181. Springer, Dordrecht (2016) [Google Scholar]
  • 56.Spekkens, R.W.: Contextuality for preparations, transformations, and unsharp measurements. Phys. Rev. A 71, 052108 (2005) [Google Scholar]
  • 57.Dakić, B., Brukner, Č: Quantum theory and beyond: is entanglement special? In: Halvorson, H. (ed.) Deep Beauty: Understanding the Quantum World through Mathematical Innovation. Cambridge University Press, Cambridge (2011) [Google Scholar]
  • 58.Masanes, L.: Extremal quantum correlations for N parties with two dichotomic observables per site. arXiv:quant-ph/0512100
  • 59.Toner, B., Verstraete, F.: Monogamy of Bell correlations and Tsirelson’s bound. arXiv:quant-ph/0611001
  • 60.Tavakoli, A., Pozas-Kerstjens, A., Brown, P., Araújo, M.: Semidefinite programming relaxations for quantum correlations. arXiv:2307.02551 (2023)
  • 61.Barnum, H., Beigi, S., Boixo, S., Elliott, M.B., Wehner, S.: Local quantum measurement and no-signaling imply quantum correlations. Phys. Rev. Lett. 104, 140401 (2010) [DOI] [PubMed] [Google Scholar]
  • 62.Acín, A., Augusiak, R., Cavalcanti, D., Hadley, C., Korbicz, J.K., Lewenstein, M., Masanes, L., Piani, M.: Unified framework for correlations in terms of local quantum observables. Phys. Rev. Lett. 104, 140404 (2010) [DOI] [PubMed] [Google Scholar]
  • 63.Schmied, R., Bancal, J.-D., Allard, B., Fadel, M., Scarani, V., Treutlein, P., Sangouard, N.: Bell correlations in a Bose–Einstein condensate. Science 352, 6284 (2016) [DOI] [PubMed] [Google Scholar]
  • 64.Šupić, I., Bowles, J.: Self-testing of quantum systems: a review. Quantum 4, 337 (2020) [Google Scholar]
  • 65.Hall, B.C.: Quantum Theory for Mathematicians. Graduate Texts in Mathematics. Springer, New York (2013) [Google Scholar]
  • 66.Itzkowitz, G., Rothman, S., Strassberg, H.: A note on the real representations of SU (2, C). J. Pure Appl. Algebra 69, 3 (1991)
  • 67.Bröcker, T., Dieck, T.: Representations of Compact Lie Groups. Graduate Texts in Mathematics. Springer, Berlin (2003) [Google Scholar]
  • 68.Navascués, M.: Correlations and symmetries, unpublished note (2021)
  • 69.Robert, A.: Introduction to the Representation Theory of Compact and Locally Compact Groups, vol. 80. Cambridge University Press, Cambridge (1983) [Google Scholar]
  • 70.Kleinmann, M., Osborne, T.J., Scholz, V.B., Werner, A.H.: Typical local measurements in generalized probabilistic theories: emergence of quantum bipartite correlations. Phys. Rev. Lett. 110, 040403 (2013) [DOI] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Data Availability Statement

Data sharing is not applicable to this article as no data sets were generated or analyzed during the current study.


Articles from Communications in Mathematical Physics are provided here courtesy of Springer

RESOURCES